Next Article in Journal
Effect of Semi-Continuous Anaerobic Digestion on the Substrate Solubilisation of Lignin-Rich Steam-Exploded Ludwigia grandiflora
Next Article in Special Issue
Immunohistochemical Expression Patterns of Tight Junction Proteins, Pro-Apoptotic and Anti-Apoptotic Factors on Progression of Intestinal Mucositis of Onco-Hematological Patients under Epirubicin-Based Chemotherapy
Previous Article in Journal
Occupational Radiation Dose, Especially for Eye Lens: Hp(3), in Medical Staff Members Involved in Computed Tomography Examinations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Bioactive Compounds from Herbal Medicine Targeting Multiple Myeloma

1
Department of Hematology, Faculty of Medicine, Vasile Goldis Western University of Arad, Rebreanu 86, 310414 Arad, Romania
2
“Aurel Ardelean” Institute of Life Sciences, Vasile Godis Western University of Arad, Rebreanu 86, 310414 Arad, Romania
3
Department of Gastroenterology, Faculty of Medicine, Vasile Goldis Western University of Arad, Rebreanu 86, 310414 Arad, Romania
4
Department of Histology, Faculty of Medicine, Vasile Goldis Western University of Arad, Rebreanu 86, 310414 Arad, Romania
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Appl. Sci. 2021, 11(10), 4451; https://doi.org/10.3390/app11104451
Submission received: 4 April 2021 / Revised: 6 May 2021 / Accepted: 11 May 2021 / Published: 13 May 2021
(This article belongs to the Special Issue New Research Trends in Hematology and Cancerous Tumours)

Abstract

:
Multiple myeloma (MM) is one of the most widespread hematological cancers. It is characterized by a clonal proliferation of malignant plasma cells in the bone marrow and by the overproduction of monoclonal proteins. In recent years, the survival rate of patients with multiple myeloma has increased significantly due to the use of transplanted stem cells and of the new therapeutic agents that have significantly increased the survival rate, but it still cannot be completely cured and therefore the development of new therapeutic products is needed. Moreover, many patients have various side effects and face the development of drug resistance to current therapies. The purpose of this review is to highlight the bioactive active compounds (flavonoids) and herbal extracts which target dysregulated signaling pathway in MM, assessed by in vitro and in vivo experiments or clinical studies, in order to explore their healing potential targeting multiple myeloma. Mechanistically, they demonstrated the ability to promote cell cycle blockage and apoptosis or autophagy in cancer cells, as well as inhibition of proliferation/migration/tumor progression, inhibition of angiogenesis in the tumor vascular network. Current research provides valuable new information about the ability of flavonoids to enhance the apoptotic effects of antineoplastic drugs, thus providing viable therapeutic options based on combining conventional and non-conventional therapies in MM therapeutic protocols.

1. Introduction

Multiple myeloma (MM) is one of the most widespread hematological cancers. It is characterized by a clonal proliferation of malignant plasma cells in the bone marrow and by the increased production of M monoclonal protein [1]. MM represents approximately 1% of all cancers [2] and 10% of all hematological tumors [2,3]. Its incidence depends on different factors, such as age, sex, and country [2] and is estimated to 6.2 in 1 × 105 individuals [4]. Moreover, MM affects people over 65 years [5], especially males, from the industrialized regions of Europe, North America, and Australia [6].
Multiple myeloma is heterogeneous cytogenetically and shows numerous genetic and epigenetic changes, in which numerous signaling pathways are involved [7]. It is characterized by chromosomal instability and frequent genetic mutations, including RB1, DIS3, KRAS, MYC, TP53, NRAS, TP53, FAM46C, BRAF, and p18 (CDKN2c) [8,9]. Due to its increased heterogeneity, MM has numerous refractory/recurrent cases. Because the bone marrow is the ideal site for homing and malignant cells progression [10], 70% of patients have bone metastases at the diagnosis, and their percentage increases to 90% during the progression of the malignancy. These patients experience bone lysis caused by inhibition of osteoblast differentiation [11]. Additionally, this condition is associated with hypercalcemia, renal failure, and anemia, which affect the quality of life as well [7,12].
In the recent years, the survival rate of MM patients has increased significantly due to the use of transplanted stem cells or new therapeutic agents, and significantly increased survival rate [13].
The main therapies used in MM are:
Glucocorticoids as dexamethasone and prednisone are steroids used in the treatment of MM [14]. Glucocorticoids indirectly inhibit target genes through inhibitory interactions of glucocorticoid (GC)-glucocorticoid receptor (GR) monomers with NF-κB and activator protein-1 (AP-1) [15], the latter being important factors in the pathogenesis of multiple myeloma [16,17,18].
  • Proteasome inhibitors [19,20,21]. The proteasome is a multicatalytic target responsible for the degradation of the proteins. To date, there are only three proteasome inhibitors approved for clinical: ixazomib [22], bortezomib [23], and carfilzomib [24]. These drugs can stimulate osteogenesis in MM patients [25]. Proteasome modulation can also be achieved by binding lenalidomide to thalidomide and acting as E3-Ligase inhibitors [26].
  • DNA damaging agents. This category includes alkylating drugs, such as melphalan, or other agents, as doxorubicin, as well as histone deacetylase inhibitors such as panobinostat [27].
  • Immunomodulatory drugs, such as pomalidomide and lenalidomide [28,29,30]. MM was one of the first tumors in which the therapeutic efficacy based on immune receptors inhibition, especially of the PD-1 axis, was evaluated in preclinical models [6].
  • Monoclonal antibodies [31,32,33]. There are two monoclonal-based therapy approved for the treatment of MM: daratumumab (CD38 pathway) [34], and elotuzumab (targetSLAMF7 pathway) [35]. The efficacy of monoclonal antibodies is the highest, but this therapy has a high cost [6]. In 2020, the FDA approved a third monoclonal antibody for MM therapy, Sarclisa (isatuximab—irfc) in combination with pomalidomide and dexamethasone [36]. Isatuximab is a monoclonal antibody that targets the transmembrane receptor and the ectoenzyme CD38, a protein overexpressed by malignant hematological cells [37]. Isatuximab is a new MM treatment for patient’s refractory to lenalidomide and proteasome inhibitor [38].
Although various pharmacological strategies for the MM clinical treatments have been developed, in most patients it eventually recurred [39]. Patients have various side effects and face the development of drug resistance [39,40,41]. One of the harmful side effects of MM therapy is neurotoxicity. It is often associated with chemotherapy-induced peripheral neuropathy and has symptoms such as hypersensitivity, dysaesthesia, and paresthesia [42,43]. More than 50% of bortezomib-treated patients experienced this side effect, leading to dose reduction or treatment discontinuation [6]. All these problems highlight the need to research new therapeutic targets to improve the treatment [44,45].
In recent years, there has been a growing interest in the development of drugs to alleviate the side effects of antineoplastic therapy and the emergence of drug resistance, through the use of biologically active compounds with low side effects and that are effective in the treatment of several cancers [46]. Bioactive compounds such as alkaloids, flavonoids, terpenoids or saponins have the ability to prevent the side effects of chemotherapeutic drugs, prolong survival and improve the quality of life of cancer patients [47,48,49,50]. It is estimated that about 70% of cancer drugs are natural products or derived from natural products [51]. However, there are limitations when using non-conventional therapies in oncohematological diseases, because in vitro or in vivo studies predominate, clinical trials are missing, or the results provide non-uniformity in their mechanism of action.
Interest in the study of phenolic compounds has increased over the last decade due to promising biological activities and potential medical applications [52,53,54]. These effects are mainly related to their antioxidant activity. Polyphenols have the ability to eliminate free radicals and chelated metal ions [55]. They can indirectly decrease the production of reactive oxygen species (ROS), either by improving the activity of antioxidant enzymes or by inhibition of pro-oxidant enzymes [56]. Phenolic compounds can induce reversal of carcinogenesis by modulating molecules involved in intracellular signaling and by blocking cancer progression [57]. The property of it to neutralize free radicals (free radical-scavenging properties) helps to prevent chronic pathologies, such as cancer, cardiovascular or neurodegenerative diseases [53]. The anticancer and pro-apoptotic properties of polyphenols have been highlighted in numerous studies [58,59], properties that qualify them to be analyzed as therapeutic options in MM. Currently, there are just a few clinical trials provided by Database ClinicalTrials.gov regarding synergic anti-tumor activity of herbal compounds and MM drugs (curcumin and Bioperine or Lenalidomide, EGCG and monoclonal gamopathy) and therefore it is necessary to focus on other flavonoids or plant extracts as posible options in the therapeutic management of MM.
The aim of this review is to highlight bioactive active compounds (flavonoids) and herbal extracts that target dysregulated signaling pathway in MM, assessed by in vitro and in vivo experiments or clinical studies, in order to explore their therapeutic potential in the treatment of multiple myeloma.

2. Molecular Pathways Involved in Multiple Myeloma Progression

Numerous signaling pathways (WNT, RANK/RANKL/OPG, RAS/MAPK, PI3K/Akt/mTOR, JAK/STAT3, NF-κB) are involved in MM, as well as extracellular mediators (TNFα, IL8, TGFβ1, (IL6, FGF-2, PDGF, IGF, VEGF) [60].
Mitochondrial-mediated apoptosis is an important apoptotic pathway, influenced by pro-apoptotic factors, such as cleaved-caspase-3 and -9, and Bcl-2 family [61]. Bcl-2 family proteins (Bax, Bad, and Bcl-2) are the major regulators of apoptotic processes and play a key role in mitochondrial-mediated apoptosis [62,63].
Activation of Bcl-2 pathways may regulate mitochondrial-mediated apoptosis. Interactions between Bcl-2 proteins can regulate permeabilization of the outer mitochondrial membrane by cytoplasmic releasing of cytochrome C, and plays a critical role in activation of the apoptosis-inherited pathway [64,65]. Moreover, Bax, Bad and Bcl-xs, may facilitate the releasing of cytochrome C or may inhibit anti-apoptotic proteins [66].
Inhibition of the PI3K/Akt/mTOR signaling pathway is crucial for the antiproliferative effect on MM cells [67,68]. Akt is the main downstream target of PI3K [69]. Inhibition of PI3K/Akt/mTOR is beneficial for improving the life quality of MM patients [70]. Phosphorylated Akt binds to PI3K-activated products, and inhibit the pro-apoptotic factors of caspase-3, -9, and Bad [71]. Moreover, MTOR is a significant downstream target of Akt [72].
STAT3 is a transcription factor that may be overexpressed in multiple myeloma. It correlates with enhanced proliferation and apoptosis resistance to therapy. STAT3 activation negatively regulated the T cell-mediated immune response [60]. Aberrant signal transducer activation and transcriptional activator 3 (STAT3) is a key event of the cancerogenesis by inducing anti-apoptosis, angiogenesis, invasion, and metastasis [73,74,75,76]. STAT3 has been shown to be constitutively active in approximately 40 to 60% of MM tumors [77,78,79,80,81].
Decreased expression of different IL6/JAK/STAT signaling pathways by epigenetic silencing may sensitize MM cells to IL-6-regulated proliferation and survival [77,78,82].
Activated STAT3 induced transcription of target genes leads to increased survival, proliferation and drug resistance of MM cells: Src homology containing protein 1 (SHP-1), suppressor of cytokine signaling 1 (SOCS1), estrogen receptor (ER), Janus Associated kinase (JAK), phosphatase of regenerating liver 3 (PRL3), and peroxisome proliferator activated receptors (PPAR) [83].
Abnormal activation of the Wnt signaling mediate MM cells proliferation, survival, and drug resistance [84,85,86]. β-catenin is an essential component of the canonical Wnt pathway regulation [87,88]. Β-catenin activity is strictly controlled by several processes, such as ligand-receptor interactionsprotein stability, nuclear translocation, protein stability, and transcriptional activity [89,90,91].
Nuclear factor kappa B (NF-κB) is an important transcription factor involved in the proliferation and MM cells survival. NF-κB regulates the activation of various molecules involved in apoptosis, as anti-apoptotic proteins or X-linked inhibitor of apoptosis protein). XIAP is the most studied protein that inhibits apoptosis and negatively regulates caspases in MM [92].
Mitogen-activated protein kinase (MAPK) is a cell survival regulator, and Jun N-terminal kinase (JNK) signaling is a component of this pathway. PARP14 is overexpressed in over 80% of MM cells and is regulated by JNK2 [93].
The best strategy for the treatment of MM is a single agent or combination of molecules targeting several dysregulated signaling pathways.

3. Flavonoids in Multiple Myeloma

Flavonoids are ubiquitous in plants and in human nutrition and bring a number of health benefits [94]. Flavonoids are classified into flavonols, flavanones, aurones, flavones, isoflavones, flavan-3-ols, proanthocyanidins, anthocyanidins, and chalcones [95].
They represent a large group of water-soluble antioxidant compounds, which frequently occur in plants as glycosides and consist of two aromatic rings linked by a carbon bond and form a heterocyclic ring [96].
Flavonoids have been extensively studied due to their antitumor activity [97,98,99]. Due to various bioactive activities, flavonoids are considered to be multi-targeting and have multifunctional molecules. They have the ability to kill resistant cancer cells and resensitize conventional anticancer drugs to induce multidrug resistance reversal, which highlights their role in resistant cancer treatment [100].
Flavonoids target membrane lipids and modify their physicochemical properties to exert their bioactive activities [101,102,103,104], most likely due to a membrane bilayer-mediated mechanism [105]. Therefore, it is considered that they act neither as specific regulators of target proteins, but rather as multifunctional agents that negatively regulate key factors involved in multidrug resistance [100].
Moreover, many flavonoids may function as ROS modulators (scavengers or inducers), because they affect the ROS level in cancer cells [106,107]. In order to overcome multidrug resistance, they act as inducers for ROS production, enreaching the toxic threshold for cancer cell apoptosis [106].
Flavonoids involved in management of MM are part of a different class (Table 1).

3.1. Apigenin

Apigenin is a non-mutagenic yellow crystalline solid flavonoid [108]. This is a flavonoid found in oranges, grapefruit, grapefruit, celery, parsley, onions, wheat germ, and chamomile (Figure 1) [109,110,111].
The anti-cancer activity of apigenin was first reported in 1986 by Birt et al. [112].
Apigenin can be used in cancer prevention due to its antioxidant, anti-inflammatory, antigenotoxic effects and its ability to neutralize free radicals [113,114]. Apigenin inhibited prostate, gastrointestinal, bladder, ovarian, colon, and neck cancers, as well as other forms of tumors by inhibiting cancer cell proliferation [115,116,117,118]. Apigenin had a cytotoxic effect on cancer cells [119] and showed proteasome inhibitory activities [120].
Adham et al. [121] evaluated the cytotoxic activity of apigenin in several human MM cell lines, such as MOLP-8, NCI-H929, RPMI-8226, KMS-11, KMS-12 BM, OPM-2, l -363, AMO-I, and JJN-3. Apigenin inhibited cell growth of MM cell lines in a dose-dependent manner, while NCIH-929 cells showed increased sensitivity to apigenin compared to other MM cells. Apigenin induced apoptosis, blocked cell cycle in G2/M phase, increased intracellular ROS, and promoted autophagy in NCI-H929 cells. Induction of autophagy by apigenin was demonstrated by upregulation of the expression of Beclin1 and LC3B-II markers.
Zhao et al. [122] reported cytotoxic effects of apigenin on both MM cell lines and primary MM cells, but not on normal blood mononuclear cells (PBMCs). Apigenin inhibited casein kinase 2 (CK2) activity and blocked the cell cycle in G2/M phase in U266 and RPMI 8226 cells. Additionally, it induced apoptosis and downregulated the expression of anti-apoptotic proteins in MM cells, respectively, such as Mcl-1, XIAP, Bcl-xL, Bcl-2, and survivin. Apigenin inhibited the activation of STAT3, ERK, AKT and NF-κB. Apigenin decreased Cdc37 phosphorylation, disassociated Hsp90/Cdc37/kinase complexes and degraded Hsp90/Cdc37 client proteins. Apigenin induced degradation of multiple kinases, including Src, RIP1, Raf-1, AKT and cyclin dependent kinase 4 (CDK4). In CD138+ primary MM cells, apigenin inhibited CK2 activity, and depleted Cdc37 client kinases.

3.2. Baicalein

Baicalein is one of the main biologically active compounds present in the root of Scutellariae radix [123]. The molecular formula of baicaleine is C15H10O5, as can be seen in Figure 2. The particular structure is the di-orthohydroxyl functional group, located on ring A of its molecular structure [124].
Baicalein has a wide range of biological activities, such as antioxidant activity [125], anti-inflammatory [126], antiviral [127], antitumor activity and lower toxicity [123].
Baicalein has inhibited many types of cancer, for example, gastric cancer, hepatocellular carcinoma, lung cancer, pancreatic cancer, bladder cancer, colorectal cancer, prostate cancer, ovarian and cervical cancer, breast cancer, multiple myeloma, melanoma and osteosarcoma [123].
The antitumor activity of baicalein is due to cell cycle regulation by inhibition of several cyclins/cyclin-dependent kinases [128], attenuating MAPK, Akt and mTOR activities or neutralizing free radicals [129], inducing apoptosis by activating caspase -9, -3 [130] and tumor invasion and metastases inhibition by decreasing MMP-2, -9 expression [131].
Baicalein inhibited the U266 cells proliferation and stimulated their apoptosis. Baicalein also increased cereblon (CRBN) mRNA level and downregulated IKZF1 and IKZF3 transcription factors by proteasomal degradation. MM patients had a lower expression of IKZF3 and IKZF1, providing a better survival rate than those with a higher expression of IKZF1 and IKZF3 [132].
Gu et al. [133] reported inhibitory effects of baicalein inhibited in RPMI 8226 cells and significantly reduced ATP-binding cassette protein expression, G subfamily, ABC -transporter proteins family [133], which is thought to be involved in drug resistance in cancer [134]. Previously, Lin et al. [135] showed that baicalein significantly reduced the RPMI8226 cells by decreasing ABCG2 expression.
Baicalein treatment inhibited the proliferation and migration of MM cells by downregulating the expression of β-catenin, cyclin D1, c-myc, and β7 integrin [136].
Liu et al. [137] demonstrated that baicalein inhibited STAT1 phosphorylation, as well as IL-6-mediated phosphorylation of Jak, MAPK, STAT3, and Akt. Baicalein increased the U266 cell sensitivity to dexamethasone and inhibited IL-6-induced bcl-xl gene expression. Baicalein also showed a strong inhibitory effect on Erk1/2 phosphorylation. Thus, baicalein could be considered an active inhibitor of IL-6-induced protein phosphorylation, and a potential agent in the treatment of MM.
Co-administration of baicalein with dexamethasone inhibited MPC-1-immature myeloma cells by activation of peroxisome proliferator-activated (PPARs) and glucocorticoid receptors, having a synergistic inhibitory effect on NF-κB [138].
Previously, Ma et al. [139] showed in vitro baicalein inhibition of the MM cells, especially MPC-1-immature MM cells. Baicalein inhibited IkB-α phosphorylation, followed by reduced IL-6 and XIAP gene expression and activation of caspase-9 and -3.

3.3. Chrysoeriol

Chrysoeriol (Figure 3) is a methoxy flavone [140,141,142,143], extracted from Rooibos tea (Aspalathus linearis) [144].
This bioactive compound has been shown to have antioxidant, anti-mutagenic, anti-inflammatory, anti-obesity properties [141,143,145,146,147,148], a cytotoxic effect in human lung carcinoma [149], and antitumor activity on breast cancer [150].
Yang et al. [151] studied the anti-myeloma activity of chryseoriol in vitro. Chrysoeriol has been shown to be a selective inhibitor of PI3K-AKT-mTOR pathway. It inhibited the proliferation of KM3 and RPMI 8226 cells, but did not affect the proliferation of the normal peripheral blood mononuclear cells. Chryseoriol blocked the cell cycle of MM cells in the G2/M phase and inhibited the phosphorylation of the AKT protein in MM cells. Chryseoriol significantly decreased p-AKT expression and increased Cyclin B1 and p21 protein expression.

3.4. Luteolin

Luteolin is a flavone found in carrots, peppers, cabbage, broccoli, onion leaves, thyme, parsley, mint, basil, celery, and artichokes [152].
Structurally luteolin has a hydroxyl (-OH) group attached to the 5-, 7-, 3′-, and 4′- positions of the flavone backbone structure. The chemical structure of luteolin is shown in Figure 4 [152]. The antioxidant activity of luteolin is ensured by the ortho-dihydroxy structure in the B-ring and the 2,3-double bond in conjugation with the 4-oxo function of the C ring [153].
Luteolin has shown antioxidant, antimicrobial, anti-inflammatory, chemopreventive, chemotherapeutic, cardioprotective, antidiabetic, neuroprotective and anti-allergic activity [152].
Luteolin has shown anticancer activity in many cancers, such as breast [154,155,156,157], colon [158], prostate [159,160], cervical [161] cancers, as well as glioblastoma [162], oral squamous cell carcinoma [163,164], adenocarcinoma [165], renal carcinoma [166], gastric cancer [167,168,169], hepatocellular carcinoma [170], and pancreatic cancer [171].
Luteolin inhibited the carcinogenesis progression through various pathways, such as kinase inhibition, induction of apoptotic cell death, cell cycle regulation, and decreased the transcription factors [153]. Apoptosis induction involves DNA damage, regulation of redox and protein kinases in inhibiting cancer cell proliferation [172].
Regarding the anti-myeloma activity of luteolin, it has been shown to inhibit RPMI-8226 cell proliferation by apoptosis or autophagy, being interactive or promoting each other [173].

3.5. Scutellarin

Scutellarin (Figure 5), a bioactive flavone isolated from Scutellaria baicalensis [174] is a proteasome inhibitor [175].
Scutellarin showed antioxidant [176], anti-inflammatory [177], and neuroprotective [178] activity. Its anti-cancer activity has been highlighted in various types of tumors, including hepatocellular carcinoma [179], colorectal cancer [180], and squamous cell carcinoma [181].
Scutellarin induced antitumor immunity and attenuated chemoresistance in neoplastic diseases by various signal transduction pathways: ERK/p53, c-met/AKT, AKT/mTOR/4EBP1, and STAT3 [174,175]. Misso et al. [182] highlight the role of the miR-125b/IL-6R/STAT3 feedback loop in modulating miR-34a in MM.
Scutellarin amplified the anti-tumoral effects of bortezomib in a xenograft mouse model. It downregulated the epigenetic modulators HDAC1/3 and amplified the expression of miR-34a. Scutellarin induced a decreased expression of c-Met, p-Akt, mTOR, NF-κB, and XIAP. Co-administration of scutellarin and bortezomib inhibited MM tumor progression. Scutellarin inhibited tumor resistance to bortezomib and induced apoptotic cell death [92].

3.6. Wogonin

Wogonin is one of the active mono-flavones extracted from the root of the Scutellaria baicalensis (Figure 6) [124,183].
It has neuroprotective [184,185], anti-inflammatory, and anti-cancer activity [186,187,188,189,190,191,192]. Wogonin has been shown to induce cytotoxicity to cancer cells in vitro and inhibited tumoral growth in vivo [124,183].
Wogonin targets several signaling pathways to prevent and inhibit the development of cancer both in vitro and in vivo [193,194]. Wogonin is involved in several anti-proliferative processes mediated by ER stress, MAPK, ROS, inhibition of transcription factors, such as NF-κB or activator protein-1 (AP-1), and intracellular suppression Ca2+ signaling [124,194]. Due to its therapeutic potential and a lower toxicity to normal cells [183], this bioactive compound is an adjuvant in chemotherapy by increasing the therapeutic effects and reducing the side effects of conventional drugs [195].
Because MM-induced angiogenesis is an essential process for cancer progession and metastasis, methods of inhibiting angiogenesis have been studied as a new anti-myeloma treatment [196,197]. Fu et al. [198] showed that wogonin inhibited MM-induced angiogenesis in both normoxic and hypoxic conditions by reducing the secretion of proangiogenic factors (VEGF, PDGF and bFGF) in U266 and RPMI 8226 cells and in co-cultures of MM and stromal cells. The antiangiogenic effect of wogonin was previously reported by Song et al. [199] in MCF-7 breast cancer cells.
The c-Myc/HIF-1α plays an essential role in activation of different pro-angiogenic factors (e.g., PDGF, VEGF, and bFGF) [200]. Wogonin induced c-Myc inhibition and stimulated HIF-1α degradation in vitro. This biologically active compound has the ability to inhibit in vivo tumor angiogenesis and growth in an experimental model of BALB/c-nude mice. Wogonin showed synergistic action against MM-dependent angiogenesis in combination with lenalidomide and bortezomib and inhibited expression of c-Myc and HIF-1α in primary MM cells [198].
Zhang et al. [201] reported a cytotoxic effect of wogonin in vitro, as demonstrated by the induction of apoptosis on RPMI8226 cells. Wogonin blocked the cells in the sub-G1 phase. Poly ADP ribose polymerase (PARP) has an important role in mediating the normal cellular response to DNA damage, and enzymatic cleavage is considered to be a marker of apoptosis. Wogonin inhibited PARP in a dose-dependent manner and induced PARP cleavage. Moreover, wogonin increased Bax expression in MM cells, suggesting its ability to stimulate apoptosis via Bax-regulating pathway. Besides, Bcl-2 expression was reduced following wogonin treatment. Wogonin inhibited Akt phosphorylation (Ser 473) in MM cells, the latter being a key event in tumorigenesis and apoptosis. Wogonin treatment did not alter the total Akt protein level. This study suggests that wogonin-induced MM cell apoptosis could be modulated by the Akt pathway, because residual phosphorylation at Ser 473 is required for Akt activation [202].

3.7. Fisetin

Fisetin is a bioactive flavonol that has a diphenylpropane structure [203] containing two aromatic rings bounded by a three carbons and supplemented with four hydroxyl group substitutions and one oxo group [204,205] (Figure 7).
Fisetin is synthesized in strawberries, apples, cucumbers, and onions [206] or plants such as Acacia berlandieri, Acacia greggii, Gleditschia triacanthow, and Rhus cotinus [207]. The average daily intake of fisetin is estimated to be about 0.4 mg in humans [208]. This potent bioactive phytoconstituent acts as a coloring agent in plants [206].
Fisetin has low water solubility and poor intestinal absorption and therefore has low bioavailability [209,210]. Its solubility and bioavailability can be improved by complexing it with cyclodextrins, encapsulating with nanoparticles [211,212,213] and cocrystallization with caffeine, isonicotinamide and nicotinamide [214,215]. These changes aim to improve stability, solubility and biological effects of fisetin. The biological activity of fisetin depends on the presence of hydroxyl and oxo groups. Hydroxyl groups at C-7 and the double bond C2–C3 are essential for its antioxidant activity [216,217].
Fisetin showed antioxidant [218], anti-angiogenic [219,220,221], anti-proliferative [222,223], anti-inflammatory [208,224], anti-aging [225], and antitumor properties [226,227].
Fisetin induced apoptosis, cell cycle blocking and inhibition of CDK activity in various human cancer cell lines. It has also modulated lipid kinase and protein kinase pathways [228,229,230]. The anticancer activity of fisetin has been highlighted by its involvement in various signaling pathways, as blocking the mammalian target of rapamycin (PI3K/Akt/mTOR)/phosphatidylinositol-3-kinase/protein kinase B, mitogen-activated protein kinases)-dependent nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB), and p38 [231,232].
Jang et al. [207] reported that fisetin showed cytotoxic effect and induced apoptosis in U266 cells and blocked the cell cycle in sub-G1 phase. Additionally, it improved the PARP cleavage. Fisetin inhibited the expression of anti-apoptotic Bcl-2 and Mcl-1L, but not Bcl-xL, and increased the expression of pro-apoptotic Bax, Bim and Bad.
AMP-activated protein kinase (AMPK) plays an important role in inducing apoptosis [233,234,235]. Fisetin induced an increase in phosphorylated AMPK and phosphorylated acetyl-CoA carboxylase, as a major substrate of AMPK. Fisetin treatment inhibited mTOR phosphorylation and reduced phosphorylated AKT, the upstream protein of mTOR. ROS have the role of mediating apoptosis in cancers [236,237]. Thus, fisetin induced ROS growth in U266 cells [207].

3.8. Myricetin

Myricetin is a phenolic compound found in berries, grapes, wine, nuts, etc. Myricetin is also known as hydroxy quercetin due to the presence of additional -OH group and exists in two forms: a free form and a glycosidically bound form (Figure 8) [238].
Myricetin has various pharmacological activities, such as antioxidant, anti-inflammatory, analgesic, antitumor, hepatoprotective, and antidiabetic activity [238].
The anti-cancer activity of myricetin was confirmed by a number of in vitro and in vivo cytotoxicity studies. Miricetin acted as an antioxidant, antimetastatic, immunomodulatory agent, apoptosis inducer, inhibitor of angiogenesis, and proliferation and growth promoter [238].
Myricetin induced apoptosis in different cell lines, including colon, pancreatic, esophageal, ovarian, or hepatoma carcinoma cells [239,240,241,242].
Akhtar et al. [243] showed that bulk and nano forms of myrcetin showed an insignificant genotoxicity level in the lymphocytes of MM patients compared to those of healthy ones. This effect was induced by altering the Bcl-2 expressions. The Bcl-2/Bax ratio decreased and the level of P53 protein increased in the lymphocytes of MM patients. Myricetin bulk and nano-treatment induced an increase in intracellular ROS levels, suggesting that modulation of apoptotic proteins triggered by myricetin occurs through P53 and oxidative stress-dependent pathways.

3.9. Quercetin

Quercetin is one of the most important bioflavonoids found in over 20 plants [244], such as berries (blueberries and cranberries), apples, green leafy vegetables, onions, broccoli, cauliflower, cabbage, as well as nuts and seeds [244,245]. It contains two benzene rings joined through a 3-carbon heterocyclic pyrone one (Figure 9). Due to the presence in the structure of quercetin of two antioxidant pharmacophores, it has the ability to easily neutralize free radicals and can bind to metal ions. Catechol with OH group present in the C3 position is an ideal structure that favors the free radical neutralization [246].
The quercetin name derives from the Latin word “Quercetum”, which means oak forest [247]. The name of the International Union of Pure and Applied Chemistry (IUPAC) of quercitin is 2- (3,4-dihydroxyphenyl) -3,5,7-trihydroxychromen-4-one [245].
Quercetin belongs to the flavonols and is not being produced by the body [247]. It is soluble in lipids and alcohol, insoluble in cold water and has a lower solubility in hot water [244,245].
Quercetin is available as a supplement and can be safely orally administrated at a dose of 1 g/ day and absorbed up to 60% [248].
Quercetin exhibits many pharmacological activities, such as antioxidant, anti-inflammatory, anti-diabetic, and anti-proliferative activity [246,249]. The anticancer properties of quercetin have been highlighted in numerous studies [250,251,252], as evidenced by the antioxidant, antiproliferative and growth factor inhibition effect deletion [253]. This flavonol has had the ability to inhibit various cancers, such as breast cancer [254], lung [255], pancreatic [256], colorectal [257], prostate [258], ovarian cancer [259], nasopharyngeal [260], and kidney [261].
Quercetin increased the sensitivity of tumor cells to doxorubicin [246,262], vincristine [263], and 5-fluorouracil [264].
Xu et al. [265] studied the effect of quercetin on multiple myeloma cell line NCI-H929. Quercetin inhibited cell proliferation, stimulated apoptosis and induced cell cycle blockade in G2/M phase. Quercetin activated caspase-3, -8, -9, and PARP apoptosis-associated proteins and inhibited BCL-2 expression. It also stimulated P21, P53, and P27 expression and inhibited CDK4 expression in NCI-H929 cells. Quercetin decreased the p-ERK and p-AKT phosphorylation in NCI-H929 cells.
He et al. [266] demonstrated that quercetin inhibited proliferation in ARP-1, RPMI8226, and MM.1R cell lines by inducing apoptosis and blocking the cell cycle in the G2/M phase. Quercetin upregulated caspase-3, -9, PARP, p21 and downregulated c-myc. The combination of quercetin with dexamethasone showed in vitro and in vivo synergistic inhibitory effect, stimulated apoptosis of MM cells and induced caspase-3 activation. Quercetin also inhibited in vivo tumor growth in vivo (ARP-1 cells in NOD–SCID mouse model).
Quercetin inhibited MM cell proliferation by downregulating the expression of IQ motif-containing GTPase activating protein 1 (IQGAP1) and the activation of extracellular signal-regulated kinase 1/2 (ERK1/2). Quercetin also inhibited activation of MAPK and the interaction between IQGAP1 and ERK1/2 in MM cells [267].

3.10. Epigallocatechin-Gallate

Epigallocatechin-gallate (EGCG) is the common catechin in green tea (Figure 10) [268,269], representing at least 50% of the total catechins in the leaves [269] and is responsible for most biological activities, including angiogenesis [270].
The anti-proliferative, pro-apoptotic, anti-invasive and anti-angiogenic properties of EGCG have been highlighted in numerous studies [271].
EGCG inhibited tumorigenesis by inhibiting carcinogenic activity [272,273]. EGCG decreased tumor proliferation by acting against angiogenesis [274,275,276,277]. EGCG inhibited tumor migration [278,279,280,281] and induced tumor cell death by several mechanisms, such as lysosomal membrane permeabilization, mediated cell death, autophagy, caspase-dependent apoptosis, and caspase-independent apoptosis [268].
EGCG targeted tumor microenviroment components (e.g., fibroblasts, macrophages, fibroblasts, and microvasculature) and reduced oxidative stress, inflammation, oxidative stress, and hypoxia [282].
EGCG has acted synergistically with other natural compounds both in vitro and in vivo in various cancers (e.g., quercetin, curcumin, genistein, and caffeine) [271,283].
EGCG has been tested in combination with currently used chemotherapeutic drugs (e.g., doxorubicin, sunitinib, and cisplatin) [284,285,286,287].
In order to improve the stability and bioavailability of EGCG, new formulations were developed [288,289,290,291].
Apoptosis is the main mechanisms by which EGCG exerts anticancer activity, along with cell proliferation and migration inhibition [292,293]. EGCG affected the expression of anti-apoptotic factors (e.g., Bcl-xl, Bcl-2) and up-regulated pro-apoptotic molecules (e.g., caspase-3, Bax) in several experiments [285,294,295,296].
EGCG inhibited cell growth by inducing apoptosis in MM cell lines or primary MM cells from patients [297,298], by caspase 3 activation and Bcl-2 and Mcl-1 reduction. Additionally, it induced loss of mitochondrial transmembrane potential and intracellular H2O2 and superoxide increasing [298].
EGCG-mediated apoptosis was the result of a direct interaction with 67LR laminin receptor and lipid-shelf clustering, which correlated with increased activation of sphingomyelinase acid (ASM) via protein kinase C delta (PKCδ). This has been demonstrated both in vitro in U266 cells and in cells harvested from patients with MM [299,300] and in vivo [297,300].
EGCG inhibited proliferation and induced apoptosis in U266 cells, a phenomena that was accompanied by the inhibition of the expression enhancer of homologous zest 2 (EZH2) [301], EGCG activated caspase-3, -8, cleaved caspase-9, and PARP and subsequent apoptosis [302].
EGCG has been used in MM clinical practice in combination with phosphodiesterase 5 inhibitor vardenafil [299] or in vitro, together with L-Threo-dihydrosphingosine in vitro [303], and hydrogen sulphide donors [304]. Thus, an attempt was made to reduce the dose of EGCG and hepatic adverse effects [305,306,307,308] and to improve the anti-hepatic effects. MM ale EGCG [299,303,304].
Bae et al. [309] reported that cyclic guanosine monophosphate (cGMP) induced by EGCG activated the PKCδ/ASM singling axis in MM cells. Induction of cGMP was sufficient to induce phosphorylation of PKCδ at Ser664, the major kinase for the induction of ASM activation by EGCG. The inhibitors of the negative regulators of diacylglycerol (DAG) increased the effect of EGCG. EGCG treatment increased phospholipase C (PLC) activity. EGCG-induced ASM activation was completely suppressed by PLC inhibition. Thus, EGCG-induced cGMP activated the cGMP/PLC/PKCδ/ASM signaling axis in MM cells.
The limitations related to the use of this flavonoid in the treatment of patients with MM are represented by the fact that EGCG is active in killing the proliferation of clonal/aberrant plasma cells (aPCs) at concentrations higher than those that can be achieved by drinking tea, and MM cells present a different sensitivity [298,299,300].

3.11. Daidzin

Daidzin (Figure 11) is a phytoestrogen that can be isolated from Pueraria lobate (Fabaceae) [310,311].
Daidzin presented a large variety of pharmacological effects, such as anti-cancer properties [312,313], anti-cholesterol [314], anti-osteoporosis [315], anti-angiocardiopathy [316] and stimulated osteoblast differentiation [317].
Daidzin has shown preventive activities against breast and prostate cancer. This bioactive compound showed anti-proliferative activity against MCF cells by affecting the invasive potential [318,319]. Daidzin has been shown to have an anti-cancer effect in the eary stages of prostate carcinogenesis [313].
Daidzin inhibited the viability of U266 and MM1.S cells, downregulated STAT3 in U266 cells, modulated activation of upstream kinases, inhibited MM cell proliferation, stimulated apoptotic cell death, potentiated anti-inflammatory and anti-cancer properties of bortezomib [320].

3.12. Formononetin

Formononetin (Figure 12), is an isoflavone, isolated from the roots of Trifolium pratense, Astragalus membranaceus, Glycyrrhiza glabra, or Pueraria lobate [321].
It has the ability to influence a number of carcinogenesis pathways in different malignant cells through various molecular mechanisms [322,323,324,325,326]. Formononetin has various pharmacological activities, such as anti-inflammatory, antiviral, antioxidant, neuroprotective activity, and wound healing [325,327,328,329,330].
Formononetin inhibited NF-κB and activated AP-1 in MM cells, attenuated activation of PI3K/AKT and MAPK pathways, decreased tumorigenic protein levels, induced apoptosis, potentiated the antitumor properties of bortezomib, and enhanced its apoptotic effects [321].
Formononetin affected MM growth through the negative regulation of (STAT) 3/5 cascades mediated via oxidative stress. Formononetin inhibited STAT activation in MM cells, reduced the ability to bind STAT3 and STAT5 to DNA, as well as the nuclear pool of p-STAT3 and p-STAT5 in U266 cells. Formononetin inhibited the activation of upstream kinases involved in the STAT3 signaling cascade in MM cells, inhibited IL-6-induced STAT3-dependent reporter gene expression, decreased the expression of proteins involved in anti-apoptosis, angiogenesis, and proliferation (CDK2, CDK4 and Cyclin E), activated caspase-3 and induced cleavage PARP in U266 cells, negatively affected the cell cycle in MM cells, reduced the expression of cyclin D1 and cyclin B1 associated with cell cycle regulation, inhibited cell viability in MM cells, inhibited glutathione reductase expression in U266 cells, induced antitumor effects in the MM model xenograft, inhibited the growth of MM in vivo and inhibited the activation of STAT3 / 5 in tumor tissues, activated caspase-3 and induced cleavage of MM tissular PARP and reduced the expression of apoptotic proteins in tumor tissues [331].
It is also thought to be involved in various signaling pathways, such as NF-κB, phosphatidylinositol 3—kinase/protein kinase B (PI3K/AKT), and AP-1 [332,333].

3.13. Genistein

Genistein belongs to the aglycone subgroup of isoflavones (Figure 13) [334].
Genistein has shown anticancer activity in many cancers, such as prostate cancer [335,336], ovarian cancer [337,338,339], gastric cancer [340], colon cancer [341], hepatocellular carcinoma [342], and neuroblastoma [343].
Genistein decreased cell mitosis, induced apoptosis and increased caspase-3 activity and inhibited NF-κB expression in U266 cells. Genistein treatment inhibited human MM cell proliferation by miR—29b upregulation [344].
Genistein inhibited NF-κB expressed by MM cells, inhibited Akt phosphorylation, decreased the expression of NF-κB-regulated genes, inhibited MM cell proliferation, and induced apoptosis in MM cells [345].

3.14. Chalcones

Chalcones are polyphenolic compounds that belong to flavonoids [346]. They are considered to be precursors of flavonoids and isoflavonoids [347,348].
Chalcone is an α, β-unsaturated ketone [349]. Chemically, chalcones are formed by openchain flavonoids in which the two aromatic rings bind three carbons in an α, -unsaturated system [350]. Chalcones have a diverse structure due to the number and position of different substituents, such as hydroxy or methoxy groups on rings A or B. Chalcones are also C-prenylated and, less frequently, O-prenylated. The most common type of prenylation is the substitution of 3,3-dimethylallyl (prenyl group). In addition, chalcones also have isopentenyl, dimethylchromano, furano, geranyl, and farnesyl groups [351].
Chalcones shows various pharmacological activities such as anticancer [352], cytotoxic, antitumor [350], antimicrobial [353], antituberculosis [354], anti-inflammatory [355], antioxidant [356,357], antidiabetic [358], and antihypertensive [359].
Chalcones act as chemopreventive agents of cancer. They have been shown to have antioxidant properties and an anticancer effect [360]. Compounds containing chalcone moiety have the ability to induce apoptosis and are blocked in the G2/M phase by triggering mitochondrial apoptotic pathway through activation of caspases and blcl-2 family [361].
Zhu et al. [61] noted the ability of 2,4-dihydroxy-30-methoxy-40-ethoxychalcone (DMEC) to inhibit cell proliferation and induce apoptosis in MM via the PI3K/Akt/mTOR pathway. The inhibitory effect of DMEC on MM cells was correlated with mitochondria-mediated apoptosis via upregulation of the cleaved-caspase-3, -9, and cytochrome C, Bcl-2 and PARP downregulation. DMEC showed a pro-apoptotic effect on U266 cells and decreased PI3K expression, p-mammalian target of rapamycin (p-mTOR) and phosphorylated-protein kinase B (Akt). Pretreatment with IGF-1 attenuated DMEC-induced apoptosis in U266 cells.

3.15. Butein

Butein is a chalcone isolated from Toxicodendron vernicifluum, Semecarpus anacardium Butea monosperma, or Dalbergia odorifera [362], with a central core structure that has been shown to give it anti-cancer properties [363]. The IUPAC name of this compound is (E) -1- (2,4-dihydroxyphenyl) -3- (3,4-dihydroxyphenyl) prop-2-en-1-one (Figure 14) [362].
Butein prevented cancer metastasis via MMP-9 and urokinase plasminogen activator (uPA) repression [364]. Butein has been shown to inhibit survival, angiogenesis, proliferation, invasion and metastasis, chemoresistance in different types of cancer, such as breast cancer, leukemia, colorectal cancer, prostate cancer, pancreatic cancer, etc., by modulating several signaling pathways. These studies highlighted that butein acts as a multitargeted agent for both cancer prevention and treatment [362].
Butein inhibited constitutive and interleukin-6-inducible STAT3 activation in MM cells. This was mediated by inhibition of the activation of the upstream kinases c-Src, Janus-like kinase (JAK) 1, and JAK2. Butein induced SHP-1 protein expression and reversed inhibition of STAT3 activation in U266 cells [365].
STAT3 activation regulates the gene expression involved in cell survival and proliferation, as antiapoptotic proteins Bcl-2, Bcl-xL, and Mcl-1 and cyclin D1 [366]. Butein down-regulated the expression of these proteins and led to the suppression of proliferation and induction of apoptosis. Overexpression of STAT3 significantly reduced buteine-induced apoptosis. Butein potentiated the apoptotic effects of thalidomide and Velcade in MM cells [366], which are two drugs used in the clinical therapy of MM [367].

3.16. Cardamonin

Cardamonin, a chalcone, has been isolated from Zingiberaceae species (Figure 15) [368]. It is also found in cardamom and other species, such as Boesenbergia pandurate, Syzygium samarangense, Catimbium speciosum, Alpinia sp., and Elettaria cardamomum [369,370,371].
Cardamonin has shown antioxidant [372], anti-inflammatory [373,374], antineoplastic [375], and hypoglycemic [376,377] activities.
The anticancer properties have been highlighted in different cancers, such as gastric cancer [378,379], nasopharyngeal cancer [380], ovarian cancer [381], breast cancer [382], lung cancer [383], colon cancer [384], prostate cancer [385], colorectal cancer [386,387], and leukemia [388].
The chemotherapeutic properties are mediated by inhibition of oncogenic molecules (c-MYC), transcription factors (NF-κB, Wnt/β-catenin, STAT-3), Bcl-2 anti-apoptotic proteins, plasma proteins (P glycoprotein) and cell cycle drivers (cyclin D1). It induced the activation of pro-apoptotic proteins (Bax) and the release of cytochrome c, autophagy inducers (LC3-II), cysteine proteases (caspase 3, 8, 9), tumor suppressor genes (p53), and thus supressed carcinogenesis, metastasis and drug resistance to conventional therapy [368].
Cardamonin inhibited the growth and proliferation of U266, RPMI 8226 and ARH-77 cells. This calcone induced apoptosis in these cell lines and induced the activation of caspase-3 and PARP. Cardamonin down-regulated the expression of anti-apoptotic gene products in MM cells, inhibited constitutively active NF-κB in MM cells, repressed the expression of IKKs and inhibited phosphorylation of IκBα in MM cells, and down-regulated NF-κB -regulated gene product expression in MM cells [375].

3.17. Isobavachalcone

Isobavachalcone is a prenylated chalcone of the flavonoids (Figure 16). The compound was first isolated in 1968 from Psoralea corylifolia L. [389] (Kuete et Sandjo, 2012). It is found in medicinal plants, such as Broussonetia papyrifera, Angelica keiskei, Psoralea corylifolia, and Maclura tinctoria [390,391].
This naturally occurring chalcone compound has been used in traditional Chinese medicine as an anthelmintic, antibacterial, aphrodisiac, astringent and antiplatelet agent [347,392,393]. Isobavachalcone showed antioxidant, antifungal, antibacterial, and anticancer activity [347].
Isobavachalcone has been shown to have anticancer activity [394,395]. It inhibited skin tumors in vivo [396], induced apoptosis in multiple myeloma, neuroblastoma, prostate cancer, breast cancer, ovarian cancer, and lung cancer [394,395,397,398].
Isobavachalcone showed reduced toxicity to normal cells. Isobavachalcone inhibited the proliferation of MM H929 cells by inducing apoptosis and autophagy. Mitochondrial cell death pathway was involved in its anti-proliferative activity. Protelytic activation of PKCδ was involved in chloroquine plus isobavachalcone-induced cell death. The combination of chloroquine with isobavachalcone showed low toxicity for normal PBMCs [397].

3.18. Isoliquiritigenin

Isoliquiritigenin (20,40,4-trihydroxychalcone), a bioactive compound with a chalcone structure, was isolated from licorice root (Figure 17) [399].
Isoliquiritigenin has various biological activities, such as antioxidant, antiviral, anti-inflammatory, neuroprotective, antidiabetic, chemopreventive and antitumor activity [399].
Isoliquiritigenin has been shown to have anticarcinogenic activity in different cancers, such as breast cancer [400], colon cancer [401], ovarian cancer [402], acute myeloid leukemia [403], and melanoma [404].
Chen et al. [405] studied the effect of isoliquiritigenin on MM cells, both in vitro and in vivo. Isoliquiritigenin inhibited the growth of MM cells and induced their apoptosis. In MM xenograft models, isoliquiritigenin showed significant antitumor activity and potentiated the anti-MM activity of adriamycin. Isoliquiritigenin downregulated IL-6 expression and reduced levels of phosphorylated ERK and STAT3. Isoliquiritigenin inhibited phosphorylation levels of ERK and STAT3 induced by recombinant human IL-6, which are critical signaling proteins in IL-6 signaling regulation networks.

3.19. Xanthohumol

Xanthohumol is a prenylated chalcone found in Humulus lupulus (Figure 18) [406,407].
Xanthohumol has a number of pharmacological activities [408]. It has shown anticancer activity in various cancers, such as breast cancer [409], colorectal cancer [410], cervical cancer [411], non-small cell lung cancer [412], and leukemia [413].
Xanthohumol showed cytostatic and cytotoxic effects in MM cells and slightly affected the viability of PBMCs. Xanthohumol treatment induced cell cycle blockade in the G1 phase, induced cell death in RPMI8226 cells by the apoptotic mechanism, affected the expression of apoptosis-related and cell cycle-associated, induced ROS generation in RPMI8226 cells, and modulated MAP phosphorylation kinases. Xanthohumol-induced apoptosis in RPMI8226 cells involved in the production of ROS and the activation of ERK1/2 and JNK1/2. Xanthohumol decreased sIL6R and VEGF production in RPMI8226 cells [414].

3.20. Bavachin

Bavachin is a prenylated flavonoid [415] isolated from Psoralea corylifolia [416] (Figure 19).
Bavachin has had beneficial effects in cancer [417], inflammation [416], and diabetes [418].
Bavachin reduced the MM cells viability, inhibited NF-κB and STAT3 phosphorylation, increased p53 and NOXA expression, decreased XIAP, survivin, Bcl-xL and Bcl-2 expression, induced apoptosis by activating caspase-3 and -9 and decreased mitochondrial membrane potential [417].

3.21. Icariin

Icariin is a prenylated flavonol glycoside (Figure 20) [419].
Icariin has shown antitumor activity in various cancers, such as gastric cancer [420], liver cancer [421,422], colon cancer [423], breast [424], ovarian [425], or esophageal [426,427].
Jung et al. [428] demonstrated that icariin had the ability to potentiate the anti-proliferative effects of bortezomib by inhibiting STAT3 activation in MM cells. Icariin inhibited STAT3 phosphorylation in U266 and MM.1S cells. Icariin inhibited STAT3 DNA binding activity and nuclear translocation in MM cells, inhibited JAK1, JAK2 and Src, inhibited inducible activation of STAT3 and upstream kinases in MM.1S cells, inhibited the IL-6-induced STAT3-dependent reporter gene, inhibited anti-apoptosis, proliferation, angiogenesis, and metastasis related proteins. Icariin activated caspase-3 and induced PARP cleavage, blocked the cell cycle and stimulated apoptosis in U266 cells, and inhibited the viability of MM cells. This bioactive compound enhanced the cytotoxic effect of bortezomib. Combination treatment of icariin with bortezomib inhibited the activation of STAT3 and its upstream kinases, facilitated the blockade of cell cycle growth in the G0/G1 phase and cellular apoptosis. Icariin showed a synergistic effect with bortezomib in inhibiting the expression of various oncogenic proteins. Inhibition of STAT3 by siRNA reversed the pro-apoptotic effects of icariin.

3.22. Icaritin

Icaritin, a prenylflavonoid, is a hydrolytic product of icariin, extras din Herba Epimedii [429]. Icaritin is a monoprenylated flavonol with 4′-methoxyl (Figure 21) [419].
Icaritin has been able to inhibit different tumors, including breast cancer [430,431,432,433,434,435], hepatocellular carcinoma [436,437,438], lung cancer [439], squamous cell carcinomas [440,441], esophageal cancer [442], ovarian cancer [443], endometrial cancer [444,445], colorectal cancer [446,447], glioblastoma [448,449], and osteosarcoma [449].
Icaritin showed cytotoxic effects and limited tumor progression. It induced inhibition of tumor cell migration, and stem/progenitor cell growth [434,442,450,451]. It also showed a significant inhibitory effect in acute myeloid leukemia (AML), multiple myeloma and lymphoma, and chronic myeloid leukemia (CML) [451,452,453,454,455].
Li et al. [456] showed that icaritin inhibited cell proliferation, induced apoptosis of MM cell line KM3/BTZ and induced the reverse multidrug resistance. The mechanisms involved are thought to be related to down-regulation of heat shock protein 27 (HSP27), P-glycoprotein (P-gp) expression, and up-regulation of Par-4 expression.
Previously, Zhu et al. [452] showed the effect of icaritin on MM cell line (U266) and primary MM cells inhibition, due to S-phase cell cycle blockade by targeting cyclin-related proteins and CDK2 and apoptosis by IL-6/JAK2/STAT3 signaling inhibition. The effects of icaritin on the inhibition of proliferation and induction of apoptosis in MM cells were independent of estrogen receptor blockade. Icaritin showed anti-myeloma activity in vivo, underlined by inhibition of tumor growth, downregulation of p-JAK2, p-STAT3 and VEGF expression and decreased serum levels of IL-6 and IgE. The antimyeloma effect of icaritin in vivo is thought to be mediated by inhibition of the p-JAK2/p-STAT3/VEGF-mediated signaling pathway.

3.23. Icariside II

Icariside II is a flavonoid glycoside isolated from Herba Epimedii (Figure 22) [457].
Icariside II showed anti-cancer activity, including osteosarcoma, prostate and breast tumors [458,459].
Icariside II inhibited the STAT3 signaling pathway in U266 cells, inhibited proliferation, induced apoptosis, and suppressed STAT3-related gene products in U266 cells. Icariside II amplified the effect of bortezomib and thalidomide on the induction of apoptosis in U266 cells [460].
The main molecular mechanisms of flavonoids against multiple myeloma (MM) from in vitro and in vivo experiments, clinical studies, and modulation of proteins and signaling pathways are summarized in Figure 23.

4. Plant Extracts in Multiple Myeloma

In recent years, research has focused on the discovery of new plant extract metabolites that act as antitumor agents in various cancers, including hematological cancers [461]. Plant extracts target different signaling pathways in cancer cells, such as proliferation, differentiation, and apoptosis [462], acting through synergistic and/or additive effects [463,464,465].
Concomitant administration of plant extracts and anti-cancer drugs has increased the therapeutic efficacy of treatment by increasing the sensitivity of cancer cells to drugs and overcoming drug-induced resistance to cancer [466]. Currently in clinical practice a significant number of plant-isolated compounds are used in the treatment of cancer in combination with other drugs [467], against hematological malignancies [468,469,470,471,472].
The main pharmacological effects exerted by herbal extracts are summarized in Table 2.

5. Conclusions and Therapeutic Perspectives

In this review, we presented 21 flavonoids and 16 plant extracts that have shown an antitumor potential in the experimental studies on MM cell lines, preclinical and clinical studies on MM patients, results that could propose them in subsequent therapeutic protocols of MM. Mechanistically, they demonstrated the ability to induce cell cycle blockage and apoptosis or autophagy in cancer cells, as well as the inhibition of proliferation/migration and tumor progression, and inhibition of angiogenesis in the tumor vascular network. However, there are currently ongoing clinical trials only for curcumin and EGCG in combination with conventional drugs. Meanwhile, current research provides valuable new information about the ability of flavonoids to enhance apoptotic effects of antineoplastic drugs (baicalein/quercetin/dexamethasone; EGCG/doxorubicin, cisplatin, sunitinib, phosphodiesterase 5 inhibitor vardenafil, safingol; Butein/thalidomide and Velcade; Icariin/bortezomib), thus providing viable therapeutic options based on combining conventional and non-conventional therapies in MM therapeutic protocols.

Author Contributions

Conceptualization C.C., A.C., A.S., A.H.; methodology C.C., A.C., A.S., E.M.; A.H.; validation C.C., A.C., A.S., A.H.; investigation, C.C., A.C., A.S., A.H.; writing—original draft preparation, C.C., A.C., A.S., A.H.; writing—review and editing, C.C., A.C., A.S., E.M., A.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bergsagel, P.L.; Kuehl, W.M. Chromosome translocations in multiple myeloma. Oncogene 2001, 20, 5611–5622. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Rajkumar, S.V. Multiple myeloma: 2016 update on diagnosis, risk-stratification, and management. Am. J. Hematol. 2016, 91, 719–734. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Spaan, I.; Raymakers, R.A.; Van de Stolpe, A.; Peperzak, V. Wnt signaling in multiple myeloma: A central player in disease with therapeutic potential. J. Hematol. Oncol. 2018, 11, 67. [Google Scholar] [CrossRef] [PubMed]
  4. Fonseca, R.; Abouzaid, S.; Bonafede, M.; Cai, Q.; Parikh, K.; Cosler, L.; Richardson, P. Trends in overall survival and costs of multiple myeloma, 2000–2014. Leukemia 2017, 31, 1915–1921. [Google Scholar] [CrossRef] [Green Version]
  5. Turesson, I.; Bjorkholm, M.; Blimark, C.H.; Kristinsson, S.; Velez, R.; Landgren, O. Rapidly changing myeloma epidemiology in the general population: Increased incidence, older patients, and longer survival. Eur. J. Haematol. 2018. [Google Scholar] [CrossRef] [Green Version]
  6. Malacrida, A.; Cavalloro, V.; Martino, E.; Cassetti, A.; Nicolini, G.; Rigolio, R.; Cavaletti, G.; Mannucci, B.; Vasile, F.; Giacomo, M.D.; et al. Anti-multiple myeloma potential of secondary metabolites from Hibiscus sabdariffa. Molecules 2019, 24, 2500. [Google Scholar] [CrossRef] [Green Version]
  7. Mirzaei, H.; Bagheri, H.; Ghasemi, F.; Khoi, J.M.; Pourhanifeh, M.H.; Heyden, Y.V.; Mortezapour, E.; Nikdasti, A.; Jeandet, P.; Khan, H.; et al. Anti-cancer activity of curcumin on multiple myeloma. Anticancer Agents Med. Chem. 2021, 21, 575–586. [Google Scholar] [CrossRef]
  8. Hu, Y.; Chen, W.; Wang, J. Progress in the identification of gene mutations involved in multiple myeloma. Oncotargets Ther. 2019, 12, 4075–4080. [Google Scholar] [CrossRef] [Green Version]
  9. Vrábel, D.; Pour, L.; Ševčíková, S. The impact of NF-κB signaling on pathogenesis and current treatment strategies in multiple myeloma. Blood Rev. 2019, 34, 56–66. [Google Scholar] [CrossRef]
  10. Moschetta, M.; Kawano, Y.; Sacco, A.; Belotti, A.; Ribolla, R.; Chiarini, M.; Giustini, V.; Bertoli, D.; Sottini, A.; Valotti, M.; et al. Bone marrow stroma and vascular contributions to myeloma bone homing. Curr. Osteoporos. Rep. 2017, 15, 499–506. [Google Scholar] [CrossRef]
  11. Shupp, A.B.; Kolb, A.D.; Mukhopadhyay, D.; Bussard, K.M. Cancer metastases to bone: Concepts, mechanisms, and interactions with bone osteoblasts. Cancers 2018, 10, 182. [Google Scholar] [CrossRef] [Green Version]
  12. Mahindra, A.; Laubach, J.; Raje, N.; Munshi, N.; Richardson, P.G.; Anderson, K. Latest advances and current challenges in the treatment of multiple myeloma. Nat. Rev. Clin. Oncol. 2012, 9, 135–143. [Google Scholar] [CrossRef]
  13. Kumar, S.K.; Dispenzieri, A.; Lacy, M.Q.; Gertz, M.A.; Buadi, F.K.; Pandey, S.; Kapoor, P.; Dingli, D.; Hayman, S.R.; Leung, N.; et al. Continued improvement in survival in multiple myeloma: Changes in early mortality and outcomes in older patients. Leukemia 2014, 28, 1122–1128. [Google Scholar] [CrossRef] [Green Version]
  14. Burwick, N.; Sharma, S. Glucocorticoids in multiple myeloma: Past, present, and future. Ann. Hematol. 2019, 98, 19–28. [Google Scholar] [CrossRef]
  15. Stahn, C.; Buttgereit, F. Genomic and nongenomic effects of glucocorticoids. Nat. Clin. Pract. Rheumatol. 2008, 4, 525–533. [Google Scholar] [CrossRef]
  16. Annunziata, C.M.; Davis, R.E.; Demchenko, Y.; Bellamy, W.; Gabrea, A.; Zhan, F.; Lenz, G.; Hanamura, I.; Wright, G.; Xiao, W.; et al. Frequent engagement of the classical and alternative NF-kappa B pathways by diverse genetic abnormalities in multiple myeloma. Cancer Cell 2007, 12, 115–130. [Google Scholar] [CrossRef] [Green Version]
  17. Keats, J.J.; Fonseca, R.; Chesi, M.; Schop, R.; Baker, A.; Chng, W.J.; Van Wier, S.; Tiedemann, R.; Shi, C.X.; Sebag, M.; et al. Promiscuous mutations activate the noncanonical NF-kappa B pathway in multiple myeloma. Cancer Cell 2007, 12, 131–144. [Google Scholar] [CrossRef] [Green Version]
  18. Fan, F.; Podar, K. The role of AP-1 transcription factors in plasma cell biology and multiple myeloma pathophysiology. Cancers 2021, 13, 2326. [Google Scholar] [CrossRef]
  19. Palumbo, A.; Chanan-Khan, A.; Weisel, K.; Nooka, A.K.; Masszi, T.; Beksac, M.; Spicka, I.; Hungria, V.; Munder, M.; Mateos, M.V.; et al. Daratumumab, bortezomib, and dexamethasone for multiple myeloma. N. Engl. J. Med. 2016, 375, 754–766. [Google Scholar] [CrossRef]
  20. Rajan, A.M.; Kumar, S. New investigational drugs with single-agent activity in multiple myeloma. Blood Cancer J. 2016, 6, e451. [Google Scholar] [CrossRef] [Green Version]
  21. Moreau, P.; Richardson, P.G.; Cavo, M.; Orlowski, R.Z.; San Miguel, J.F.; Palumbo, A.; Harousseau, J.L. Proteasome inhibitors in multiple myeloma: 10 years later. Blood 2012, 120, 947–959. [Google Scholar] [CrossRef] [Green Version]
  22. Shirley, M. Ixazomib: First global approval. Drugs 2016, 76, 405–411. [Google Scholar] [CrossRef]
  23. Guerrero-Garcia, T.A.; Mogollon, R.J.; Castillo, J.J. Bortezomib in plasmablastic lymphoma: A glimpse of hope for a hard-to-treat disease. Leuk. Res. 2017, 62, 12–16. [Google Scholar] [CrossRef]
  24. Ziogas, D.C.; Terpos, E.; Kastritis, E.; Dimopoulos, M.A. An overview of the role of carfilzomib in the treatment of multiple myeloma. Expert Opin. Pharmacother. 2017, 18, 1883–1897. [Google Scholar] [CrossRef]
  25. Pennisi, A.; Li, X.; Ling, W.; Khan, S.; Zangari, M.; Yaccoby, S. The proteasome inhibitor, bortezomib suppresses primary myeloma and stimulates bone formation in myelomatous and nonmyelomatous bones in vivo. Am. J. Hematol. 2009, 84, 6–14. [Google Scholar] [CrossRef] [Green Version]
  26. Driscoll, J. Expression of E3 ubiquitin ligases in multiple myeloma patients after treatment with the proteasome inhibitor bortezomib. Cancer Transl. Med. 2015, 1, 153. [Google Scholar] [CrossRef]
  27. López-Iglesias, A.A.; González-Méndez, L.; San-Segundo, L.; Herrero, A.B.; Hernández-García, S.; Martín-Sánchez, M.; Gutiérrez, N.C.; Paíno, T.; Avilés, P.; Mateos, M.V.; et al. Synergistic DNA-damaging effect in multiple myeloma with the combination of zalypsis, bortezomib and dexamethasone. Haematologica 2017, 102, 168–175. [Google Scholar] [CrossRef] [Green Version]
  28. Costa, F.; Das, R.; Bailur, J.K.; Dhodapkar, K.; Dhodapkar, M.V. Checkpoint inhibition in myeloma: Opportunities and challenges. Front. Immunol. 2018, 9, 2204. [Google Scholar] [CrossRef] [Green Version]
  29. Hoy, S.M. Pomalidomide: A review in relapsed and refractory multiple myeloma. Drugs 2017, 77, 1897–1908. [Google Scholar] [CrossRef] [PubMed]
  30. Kocoglu, M.; Badros, A. The role of immunotherapy in multiple myeloma. Pharmaceuticals 2016, 9, 3. [Google Scholar] [CrossRef] [PubMed]
  31. Ishida, T. Therapeutic antibodies for multiple myeloma. Jpn. J. Clin. Oncol. 2018, 48, 957–963. [Google Scholar] [CrossRef] [PubMed]
  32. Cho, S.F.; Lin, L.; Xing, L.; Yu, T.; Wen, K.; Anderson, K.C.; Tai, Y.T. Monoclonal antibody: A new treatment strategy against multiple myeloma. Antibodies 2017, 6, 18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Sherbenou, D.W.; Mark, T.M.; Forsberg, P. Monoclonal antibodies in multiple myeloma: A new wave of the future. Clin. Lymphoma Myeloma Leuk. 2017, 17, 545–554. [Google Scholar] [CrossRef]
  34. Blair, H.A. Daratumumab: A review in relapsed and/or refractory multiple myeloma. Drugs 2017, 77, 2013–2024. [Google Scholar] [CrossRef]
  35. Passey, C.; Sheng, J.; Mora, J.; Tendolkar, A.; Robbins, M.; Dodge, R.; Roy, A.; Bello, A.; Gupta, M. The clinical pharmacology of elotuzumab. Clin. Pharmacokinet. 2018, 57, 297–313. [Google Scholar] [CrossRef]
  36. US Food and Drug Administration. New Drugs Approval Report 2020. Available online: https://www.fda.gov/media/144982/download (accessed on 20 April 2021).
  37. Zhu, C.; Song, Z.; Wang, A.; Srinivasan, S.; Yang, G.; Greco, R.; Theilhaber, J.; Shehu, E.; Wu, L.; Yang, Z.Y.; et al. Isatuximab acts through Fc-dependent, independent, and direct pathways to kill multiple myeloma cells. Front. Immunol. 2020, 11, 1771. [Google Scholar] [CrossRef]
  38. Attal, M.; Richardson, P.G.; Rajkumar, S.V.; San-Miguel, J.; Beksac, M.; Spicka, I.; Leleu, X.; Schjesvold, F.; Moreau, P.; Dimopoulos, M.A.; et al. Isatuximab plus pomalidomide and low-dose dexamethasone versus pomalidomide and low-dose dexamethasone in patients with relapsed and refractory multiple myeloma (ICARIA-MM): A randomised, multicentre, open-label, phase 3 study. Lancet 2019, 394, 2096–2107. [Google Scholar] [CrossRef]
  39. Kumar, S.K.; Rajkumar, V.; Kyle, R.A.; van Duin, M.; Sonneveld, P.; Mateos, M.V.; Gay, F.; Anderson, K.C. Multiple myeloma. Nat. Rev. Dis. Primers 2017, 3, 17046. [Google Scholar] [CrossRef]
  40. Robak, P.; Drozdz, I.; Szemraj, J.; Robak, T. Drug resistance in multiple myeloma. Cancer Treat. Rev. 2018, 70, 199–208. [Google Scholar] [CrossRef]
  41. Gay, F.; Palumbo, A. Multiple myeloma: Management of adverse events. Med. Oncol. 2010, 27, 646–653. [Google Scholar] [CrossRef]
  42. Meregalli, C. An Overview of bortezomib-induced neurotoxicity. Toxics 2015, 3, 294–303. [Google Scholar] [CrossRef] [PubMed]
  43. Tacchetti, P.; Terragna, C.; Galli, M.; Zamagni, E.; Petrucci, M.T.; Pezzi, A.; Montefusco, V.; Martello, M.; Tosi, P.; Baldini, L.; et al. Bortezomib- and thalidomide-induced peripheral neuropathy in multiple myeloma: Clinical and molecular analyses of a phase 3 study. Am. J. Hematol. 2014, 89, 1085–1091. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Caivano, A.; La Rocca, F.; Laurenzana, I.; Annese, T.; Tamma, R.; Famigliari, U.; Simeon, V.; Trino, S.; De Luca, L.; Villani, O.; et al. Epha3 acts as proangiogenic factor in multiple myeloma. Oncotarget 2017, 8, 34298–34309. [Google Scholar] [CrossRef]
  45. La Rocca, F.; Airoldi, I.; Di Carlo, E.; Marotta, P.; Falco, G.; Simeon, V.; Laurenzana, I.; Trino, S.; De Luca, L.; Todoerti, K.; et al. EphA3 targeting reduces in vitro adhesion and invasion and in vivo growth and angiogenesis of multiple myeloma cells. Cell Oncol. 2017, 40, 483–496. [Google Scholar] [CrossRef]
  46. Aung, T.N.; Qu, Z.; Kortschak, R.D.; Adelson, D.L. Understanding the effectiveness of natural compound mixtures in cancer through their molecular mode of action. Int. J. Mol. Sci. 2017, 18, 656. [Google Scholar] [CrossRef]
  47. Ranaware, A.M.; Banik, K.; Deshpande, V.; Padmavathi, G.; Roy, N.K.; Sethi, G.; Fan, L.; Kumar, A.P.; Kunnumakkara, A.B. Magnolol: A neolignan from the Magnolia family for the prevention and treatment of cancer. Int. J. Mol. Sci. 2018, 19, 2362. [Google Scholar] [CrossRef] [Green Version]
  48. Tewari, D.; Nabavi, S.F.; Nabavi, S.M.; Sureda, A.; Farooqi, A.A.; Atanasov, A.G.; Vacca, R.A.; Sethi, G.; Bishayee, A. Targeting activator protein 1 signaling pathway by bioactive natural agents: Possible therapeutic strategy for cancer prevention and intervention. Pharmacol. Res. 2018, 128, 366–375. [Google Scholar] [CrossRef]
  49. Harsha, C.; Banik, K.; Bordoloi, D.; Kunnumakkara, A.B. Antiulcer properties of fruits and vegetables: A mechanism based perspective. Food Chem. Toxicol. 2017, 108, 104–119. [Google Scholar] [CrossRef]
  50. Deorukhkar, A.; Krishnan, S.; Sethi, G.; Aggarwal, B.B. Back to basics: How natural products can provide the basis for new therapeutics. Expert Opin. Investig. Drugs 2007, 16, 1753–1773. [Google Scholar] [CrossRef]
  51. Karikas, G.A. Anticancer and chemopreventing natural products: Some biochemical and therapeutic aspects. J. BUON 2010, 15, 627–638. [Google Scholar]
  52. Cosme, P.; Rodríguez, A.B.; Espino, J.; Garrido, M. Plant phenolics: Bioavailability as a key determinant of their potential health-promoting applications. Antioxidants 2020, 9, 1263. [Google Scholar] [CrossRef] [PubMed]
  53. Bhuyan, D.J.; Basu, A. Utilisation of bioactive compounds derived from waste in the food industry. In Utilisation of Bioactive Compounds from Agricultural and Food Production Waste; Vuong, Q.V., Ed.; CRC Press: Boca Ratón, FL, USA, 2017; pp. 342–357. [Google Scholar]
  54. Shahidi, F.; Ambigaipalan, P. Phenolics and polyphenolics in foods, beverages and spices: Antioxidant activity and health effects—A review. J. Funct. Foods 2015, 18, 820–897. [Google Scholar] [CrossRef]
  55. Quiñones, M.; Miguel, M.; Aleixandre, A. Los polifenoles, compuestos de origen natural con efectos saludables sobre el sistema cardiovascular. Nutr. Hosp. 2012, 27, 76–89. [Google Scholar] [CrossRef]
  56. Ballard, C.R.; Maróstica, M.R. Health benefits of flavonoids. In Bioactive Compounds: Health Benefits and Potential Applications; Segura-Campos, M.R., Ed.; Elsevier Inc.: Amsterdam, The Netherlands, 2018; pp. 185–201. [Google Scholar]
  57. Gioxari, A.; Kogiannou, D.A.A.; Kalogeropoulos, N.; Kaliora, A.C. Phenolic compounds: Bioavailability and health effects. In Encyclopedia of Food and Health; Caballero, B., Finglas, P.M., Toldrá, F., Eds.; Elsevier: Cambridge, MA, USA, 2016; pp. 339–345. [Google Scholar]
  58. Estrela, J.M.; Mena, S.; Obrador, E.; Benlloch, M.; Castellano, G.; Salvador, R.; Dellinger, R.W. Polyphenolic phytochemicals in cancer prevention and therapy: Bioavailability versus bioefficacy. J. Med. Chem. 2017, 60, 9413–9436. [Google Scholar] [CrossRef]
  59. Kou, X.; Han, L.; Li, X.; Xue, Z.; Zhou, F. Antioxidant and antitumor effects and immunomodulatory activities of crude and purified polyphenol extract from blueberries. Front. Chem. Sci. Eng. 2016, 10, 108–119. [Google Scholar] [CrossRef]
  60. Pojero, F.; Poma, P.; Spanò, V.; Montalbano, A.; Barraja, P.; Notarbartolo, M. Targeting multiple myeloma with natural polyphenols. Eur. J. Med. Chem. 2019, 180, 465–485. [Google Scholar] [CrossRef]
  61. Zhu, F.; Jiang, D.; Zhang, M.; Zhao, B. 2,4-Dihydroxy-3’-methoxy-4’-ethoxychalcone suppresses cell proliferation and induces apoptosis of multiple myeloma via the PI3K/akt/mTOR signaling pathway. Pharm. Biol. 2019, 57, 641–648. [Google Scholar] [CrossRef] [Green Version]
  62. Dai, Y.; Jin, S.; Li, X.; Wang, D. The involvement of Bcl-2 family proteins in AKT-regulated cell survival in cisplatin resistant epithelial ovarian cancer. Oncotarget 2017, 8, 1354–1368. [Google Scholar] [CrossRef] [Green Version]
  63. Brunelle, J.K.; Letai, A. Control of mitochondrial apoptosis by the Bcl-2 family. J. Cell Sci. 2009, 122, 437–441. [Google Scholar] [CrossRef] [Green Version]
  64. Long, F.; Wang, T.; Jia, P.; Wang, H.; Qing, Y.; Xiong, T.; He, M.; Wang, X. Anti-tumor effects of atractylenolide-I on human ovarian cancer cells. Med. Sci. Monit. 2017, 23, 571–579. [Google Scholar] [CrossRef] [Green Version]
  65. Vidya Priyadarsini, R.; Senthil Murugan, R.; Maitreyi, S.; Ramalingam, K.; Karunagaran, D.; Nagini, S. The flavonoid quercetin induces cell cycle arrest and mitochondria-mediated apoptosis in human cervical cancer (HeLa) cells through p53 induction and NF-κB inhibition. Eur. J. Pharmacol. 2010, 649, 84–91. [Google Scholar] [CrossRef]
  66. Jiang, X.; Jiang, H.; Shen, Z.; Wang, X. Activation of mitochondrial protease OMA1 by Bax and Bak promotes cytochrome c release during apoptosis. Proc. Natl. Acad. Sci. USA 2014, 111, 14782–14787. [Google Scholar] [CrossRef] [Green Version]
  67. Porta, C.; Paglino, C.; Mosca, A. Targeting PI3K/Akt/mTOR signaling in cancer. Front. Oncol. 2014, 4, 64. [Google Scholar] [CrossRef] [Green Version]
  68. Ikeda, H.; Hideshima, T.; Fulciniti, M.; Perrone, G.; Miura, N.; Yasui, H.; Okawa, Y.; Kiziltepe, T.; Santo, L.; Vallet, S.; et al. PI3K/p110{delta} is a novel therapeutic target in multiple myeloma. Blood 2010, 116, 1460–1468. [Google Scholar] [CrossRef] [Green Version]
  69. Qu, W.; Pang, C.; Xue, Y.; Zhang, Q.; Wei, X. Dihydroartemisinin inhibits the Raf/ERK/MEK and PI3K/AKT pathways in glioma cells. Oncol. Lett. 2015, 10, 3266–3270. [Google Scholar] [CrossRef] [Green Version]
  70. Liu, T.J.; Koul, D.; LaFortune, T.; Tiao, N.; Shen, R.J.; Maira, S.M.; Garcia-Echevrria, C.; Yung, W.K. NVP-BEZ235, a novel dual phosphatidylinositol 3-kinase/mammalian target of rapamycin inhibitor, elicits multifaceted antitumor activities in human gliomas. Mol. Cancer Ther. 2009, 8, 2204–2210. [Google Scholar] [CrossRef] [Green Version]
  71. Wu, X.Y.; Tian, F.; Su, M.H.; Wu, M.; Huang, Y.; Hu, L.H.; Jin, L.; Zhu, X.J. BF211, a derivative of bufalin, enhances the cytocidal effects in multiple myeloma cells by inhibiting the IL-6/JAK2/STAT3 pathway. Int. Immunopharmacol. 2018, 64, 24–32. [Google Scholar] [CrossRef]
  72. Zhang, L.; Wang, H.; Zhu, J.; Xu, J.; Ding, K. Mollugin induces tumor cell apoptosis and autophagy via the PI3K/AKT/mTOR/p70S6K and ERK signaling pathways. Biochem. Biophys. Res. Commun. 2014, 450, 247–254. [Google Scholar] [CrossRef] [PubMed]
  73. Wong, A.L.A.; Hirpara, J.L.; Pervaiz, S.; Eu, J.Q.; Sethi, G.; Goh, B.C. Do STAT3 inhibitors have potential in the future for cancer therapy? Expert Opin. Investig. Drugs 2017, 26, 883–887. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Chai, E.Z.; Shanmugam, M.K.; Arfuso, F.; Dharmarajan, A.; Wang, C.; Kumar, A.P.; Samy, R.P.; Lim, L.H.; Wang, L.; Goh, B.C.; et al. Targeting transcription factor STAT3 for cancer prevention and therapy. Pharmacol. Ther. 2016, 162, 86–97. [Google Scholar] [CrossRef] [PubMed]
  75. Shanmugam, M.K.; Lee, J.H.; Chai, E.Z.; Kanchi, M.M.; Kar, S.; Arfuso, F.; Dharmarajan, A.; Kumar, A.P.; Ramar, P.S.; Looi, C.Y.; et al. Cancer prevention and therapy through the modulation of transcription factors by bioactive natural compounds. Semin. Cancer Biol. 2016, 40–41, 35–47. [Google Scholar] [CrossRef]
  76. Siveen, K.S.; Sikka, S.; Surana, R.; Dai, X.; Zhang, J.; Kumar, A.P.; Tan, B.K.; Sethi, G.; Bishayee, A. Targeting the STAT3 signaling pathway in cancer: Role of synthetic and natural inhibitors. Biochim. Biophys. Acta 2014, 1845, 136–154. [Google Scholar] [CrossRef] [Green Version]
  77. Sikka, S.; Shanmugam, M.; Kannaiyan, R.; Surana, R.; Shin, E.M.; Kumar, A.; Sethi, G.; Ahn, K. Suppression of essential pro-inflammatory signaling pathways by natural agents for the therapy of multiple myeloma. Phytochem. Rev. 2013, 13, 79–106. [Google Scholar] [CrossRef]
  78. Kannaiyan, R.; Surana, R.; Shin, E.M.; Ramachandran, L.; Sethi, G.; Kumar, A.P. Targeted Inhibition of Multiple Proinflammatory Signalling Pathways for the Prevention and Treatment of Multiple Myeloma, in Multiple Myeloma—An Overview; Gupta, A., Ed.; InTech: Rijeka, Croatia, 2012; pp. 93–128. [Google Scholar] [CrossRef] [Green Version]
  79. Bharti, A.C.; Shishodia, S.; Reuben, J.M.; Weber, D.; Alexanian, R.; Raj-Vadhan, S.; Estrov, Z.; Talpaz, M.; Aggarwal, B.B. Nuclear factor-kappa B and STAT3 are constitutively active in CD138+ cells derived from multiple myeloma patients, and suppression of these transcription factors leads to apoptosis. Blood 2004, 103, 3175–3184. [Google Scholar] [CrossRef]
  80. Quintanilla-Martinez, L.; Kremer, M.; Specht, K.; Calzada-Wack, J.; Nathrath, M.; Schaich, R.; Höfler, H.; Fend, F. Analysis of signal transducer and activator of transcription 3 (Stat 3) pathway in multiple myeloma: Stat 3 activation and cyclin D1 dysregulation are mutually exclusive events. Am. J. Pathol. 2003, 162, 1449–1461. [Google Scholar] [CrossRef]
  81. Catlett-Falcone, R.; Landowski, T.H.; Oshiro, M.M.; Turkson, J.; Levitzki, A.; Savino, R.; Ciliberto, G.; Moscinski, L.; Fernández-Luna, J.L.; Nuñez, G.; et al. Constitutive activation of Stat3 signaling confers resistance to apoptosis in human U266 myeloma cells. Immunity 1999, 10, 105–115. [Google Scholar] [CrossRef] [Green Version]
  82. Galm, O.; Yoshikawa, H.; Esteller, M.; Osieka, R.; Herman, J.G. SOCS-1, a negative regulator of cytokine signaling, is frequently silenced by methylation in multiple myeloma. Blood 2003, 101, 2784–2788. [Google Scholar] [CrossRef] [Green Version]
  83. Chong, P.S.Y.; Chng, W.J.; de Mel, S. STAT3: A promising therapeutic target in multiple myeloma. Cancers 2019, 11, 731. [Google Scholar] [CrossRef] [Green Version]
  84. Savvidou, I.; Khong, T.; Cuddihy, A.; McLean, C.; Horrigan, S.; Spencer, A. β-catenin inhibitor BC2059 is efficacious as monotherapy or in combination with proteasome inhibitor bortezomib in multiple myeloma. Mol. Cancer Ther. 2017, 16, 1765–1778. [Google Scholar] [CrossRef] [Green Version]
  85. Merchant, A.A.; Matsui, W. Targeting Hedgehog—A cancer stem cell pathway. Clin. Cancer Res. 2010, 16, 3130–3140. [Google Scholar] [CrossRef] [Green Version]
  86. Sukhdeo, K.; Mani, M.; Zhang, Y.; Dutta, J.; Yasui, H.; Rooney, M.D.; Carrasco, D.E.; Zheng, M.; He, H.; Tai, Y.T.; et al. Targeting the beta-catenin/TCF transcriptional complex in the treatment of multiple myeloma. Proc. Natl. Acad. Sci. USA 2007, 104, 7516–7521. [Google Scholar] [CrossRef] [Green Version]
  87. Behrens, J.; von Kries, J.P.; Kühl, M.; Bruhn, L.; Wedlich, D.; Grosschedl, R.; Birchmeier, W. Functional interaction of beta-catenin with the transcription factor LEF-1. Nature 1996, 382, 638–642. [Google Scholar] [CrossRef]
  88. Molenaar, M.; van de Wetering, M.; Oosterwegel, M.; Peterson-Maduro, J.; Godsave, S.; Korinek, V.; Roose, J.; Destrée, O.; Clevers, H. XTcf-3 transcription factor mediates beta-catenin-induced axis formation in Xenopus embryos. Cell 1996, 86, 391–399. [Google Scholar] [CrossRef] [Green Version]
  89. Van Andel, H.; Kocemba, K.A.; Spaargaren, M.; Pals, S.T. Aberrant Wnt signaling in multiple myeloma: Molecular mechanisms and targeting options. Leukemia 2019, 33, 1063–1075. [Google Scholar] [CrossRef]
  90. Yang, Y.; Shi, J.; Tolomelli, G.; Xu, H.; Xia, J.; Wang, H.; Zhou, W.; Zhou, Y.; Das, S.; Gu, Z.; et al. RARα2 expression confers myeloma stem cell features. Blood 2013, 122, 1437–1447. [Google Scholar] [CrossRef]
  91. Kramps, T.; Peter, O.; Brunner, E.; Nellen, D.; Froesch, B.; Chatterjee, S.; Murone, M.; Züllig, S.; Basler, K. Wnt/wingless signaling requires BCL9/legless-mediated recruitment of pygopus to the nuclear beta-catenin-TCF complex. Cell 2002, 109, 47–60. [Google Scholar] [CrossRef] [Green Version]
  92. Li, L.; Zheng, Y.; Zhang, W.; Hou, L.; Gao, Y. Scutellarin circumvents chemoresistance, promotes apoptosis, and represses tumor growth by HDAC/miR-34a-mediated down-modulation of Akt/mTOR and NF-κB-orchestrated signaling pathways in multiple myeloma. Int. J. Clin. Exp. Pathol. 2020, 13, 212–219. [Google Scholar]
  93. Qin, W.; Wu, H.J.; Cao, L.Q.; Li, H.J.; He, C.X.; Zhao, D.; Xing, L.; Li, P.Q.; Jin, X.; Cao, H.L. Research progress on PARP14 as a drug target. Front. Pharmacol. 2019, 10, 172. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Li, N.; Liu, J.H.; Zhang, J.; Yu, B.Y. Comparative evaluation of cytotoxicity and antioxidative activity of 20 flavonoids. J. Agric. Food Chem. 2008, 56, 3876–3883. [Google Scholar] [CrossRef] [PubMed]
  95. Mahapatra, D.K.; Bharti, S.K.; Asati, V. Anti-cancer chalcones: Structural and molecular target perspectives. Eur. J. Med. Chem. 2015, 98, 69–114. [Google Scholar] [CrossRef] [PubMed]
  96. Pandey, K.B.; Rizvi, S.I. Plant polyphenols as dietary antioxidants in human health and disease. Oxid. Med. Cell. Longev. 2009, 2, 270–278. [Google Scholar] [CrossRef] [Green Version]
  97. Díaz, J.G.; Carmona, A.J.; Torres, F.; Quintana, J.; Estévez, F.; Herz, W. Cytotoxic activities of flavonoid glycoside acetates from Consolida oliveriana. Planta Med. 2008, 74, 171–174. [Google Scholar] [CrossRef] [Green Version]
  98. Li, Y.L.; Gan, G.P.; Zhang, H.Z.; Wu, H.Z.; Li, C.L.; Huang, Y.P.; Liu, Y.W.; Liu, J.W. A flavonoid glycoside isolated from Smilax china L. rhizome in vitro anticancer effects on human cancer cell lines. J. Ethnopharmacol. 2007, 113, 115–124. [Google Scholar] [CrossRef]
  99. Middleton, E., Jr.; Kandaswami, C.; Theoharides, T.C. The effects of plant flavonoids on mammalian cells: Implications for inflammation, heart disease, and cancer. Pharmacol. Rev. 2000, 52, 673–751. [Google Scholar]
  100. Ye, Q.; Liu, K.; Shen, Q.; Li, Q.; Hao, J.; Han, F.; Jiang, R.W. Reversal of multidrug resistance in cancer by multi-functional flavonoids. Front. Oncol. 2019, 9, 487. [Google Scholar] [CrossRef] [Green Version]
  101. Cui, Q.; Wen, S.; Huang, P. Targeting cancer cell mitochondria as a therapeutic approach: Recent updates. Future Med. Chem. 2017, 9, 929–949. [Google Scholar] [CrossRef]
  102. Selvaraj, S.; Krishnaswamy, S.; Devashya, V.; Sethuraman, S.; Krishnan, U.M. Influence of membrane lipid composition on flavonoid-membrane interactions: Implications on their biological activity. Prog. Lipid Res. 2015, 58, 1–13. [Google Scholar] [CrossRef]
  103. Kuete, V.; Sandjo, L.P.; Djeussi, D.E.; Zeino, M.; Kwamou, G.M.; Ngadjui, B.; Efferth, T. Cytotoxic flavonoids and isoflavonoids from Erythrina sigmoidea towards multi-factorial drug resistant cancer cells. Investig. New Drugs 2014, 32, 1053–1062. [Google Scholar] [CrossRef]
  104. Margina, D.; Ilie, M.; Manda, G.; Neagoe, I.; Mocanu, M.; Ionescu, D.; Gradinaru, D.; Ganea, C. Quercetin and epigallocatechin gallate effects on the cell membranes biophysical properties correlate with their antioxidant potential. Gen. Physiol. Biophys. 2012, 31, 47–55. [Google Scholar] [CrossRef] [Green Version]
  105. Ingólfsson, H.I.; Thakur, P.; Herold, K.F.; Hobart, E.A.; Ramsey, N.B.; Periole, X.; de Jong, D.H.; Zwama, M.; Yilmaz, D.; Hall, K.; et al. Phytochemicals perturb membranes and promiscuously alter protein function. ACS Chem. Biol. 2014, 9, 1788–1798. [Google Scholar] [CrossRef]
  106. Cui, Q.; Wang, J.Q.; Assaraf, Y.G.; Ren, L.; Gupta, P.; Wei, L.; Ashby, C.R., Jr.; Yang, D.H.; Chen, Z.S. Modulating ROS to overcome multidrug resistance in cancer. Drug Resist. Updates 2018, 41, 1–25. [Google Scholar] [CrossRef] [PubMed]
  107. Cui, Q.; Yang, D.H.; Chen, Z.S. Special Issue: Natural products: Anticancer and beyond. Molecules 2018, 23, 1246. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Ali, F.; Naz, F.; Jyoti, S.; Siddique, Y.H. Health functionality of apigenin: A review. Int. J. Food Prop. 2017, 20, 1197–1238. [Google Scholar] [CrossRef]
  109. Salehi, B.; Venditti, A.; Sharifi-Rad, M.; Kręgiel, D.; Sharifi-Rad, J.; Durazzo, A.; Lucarini, M.; Santini, A.; Souto, E.B.; Novellino, E.; et al. The therapeutic potential of apigenin. Int. J. Mol. Sci. 2019, 20, 1305. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Shukla, S.; Gupta, S. Apigenin: A promising molecule for cancer prevention. Pharm. Res. 2010, 27, 962–978. [Google Scholar] [CrossRef]
  111. Patel, D.; Shukla, S.; Gupta, S. Apigenin and cancer chemoprevention: Progress, potential and promise (review). Int. J. Oncol. 2007, 30, 233–245. [Google Scholar] [CrossRef] [Green Version]
  112. Birt, D.F.; Walker, B.; Tibbels, M.G.; Bresnick, E. Anti-mutagenesis and anti-promotion by apigenin, robinetin and indole-3-carbinol. Carcinogenesis 1986, 7, 959–963. [Google Scholar] [CrossRef]
  113. Ali, F.; Naz, F.; Jyoti, S.; Siddique, Y.H. Protective effect of apigenin against N-nitrosodiethylamine (NDEA)-induced hepatotoxicity in albino rats. Mutat. Res. Genet. Toxicol. Environ. Mutagen. 2014, 767, 13–20. [Google Scholar] [CrossRef]
  114. Singh, P.; Mishra, S.K.; Noel, S.; Sharma, S.; Rath, S.K. Acute exposure of apigenin induces hepatotoxicity in Swiss mice. PLoS ONE 2012, 7, e31964. [Google Scholar] [CrossRef] [Green Version]
  115. Ghițu, A.; Schwiebs, A.; Radeke, H.H.; Avram, S.; Zupko, I.; Bor, A.; Pavel, I.Z.; Dehelean, C.A.; Oprean, C.; Bojin, F.; et al. A comprehensive assessment of apigenin as an antiproliferative, proapoptotic, antiangiogenic and immunomodulatory phytocompound. Nutrients 2019, 11, 858. [Google Scholar] [CrossRef] [Green Version]
  116. Chen, V.; Staub, R.E.; Baggett, S.; Chimmani, R.; Tagliaferri, M.; Cohen, I.; Shtivelman, E. Identification and analysis of the active phytochemicals from the anti-cancer botanical extract Bezielle. PLoS ONE 2012, 7, e30107. [Google Scholar] [CrossRef] [Green Version]
  117. Lefort, E.C.; Blay, J. The dietary flavonoid apigenin enhances the activities of the anti-metastatic protein CD26 on human colon carcinoma cells. Clin. Exp. Metastasis 2011, 28, 337–349. [Google Scholar] [CrossRef]
  118. Lee, S.W.; Lee, J.T.; Lee, M.G.; Lee, H.W.; Ahn, S.J.; Lee, Y.J.; Lee, Y.L.; Yoo, J.; Ahn, B.C.; Ha, J.H. In vitro antiproliferative characteristics of flavonoids and diazepam on SNU-C4 colorectal adenocarcinoma cells. J. Nat. Med. 2009, 63, 124–129. [Google Scholar] [CrossRef]
  119. Gupta, S.; Afaq, F.; Mukhtar, H. Selective growth-inhibitory, cell-cycle deregulatory and apoptotic response of apigenin in normal versus human prostate carcinoma cells. Biochem. Biophys. Res. Commun. 2001, 287, 914–920. [Google Scholar] [CrossRef]
  120. Wu, Y.X.; Fang, X. Apigenin, chrysin, and luteolin selectively inhibit chymotrypsin-like and trypsin-like proteasome catalytic activities in tumor cells. Planta Med. 2010, 76, 128–132. [Google Scholar] [CrossRef]
  121. Adham, A.N.; Abdelfatah, S.; Naqishbandi, A.M.; Mahmoud, N.; Efferth, T. Cytotoxicity of apigenin toward multiple myeloma cell lines and suppression of iNOS and COX-2 expression in STAT1-transfected HEK293 cells. Phytomedicine 2021, 80, 153371. [Google Scholar] [CrossRef]
  122. Zhao, M.; Ma, J.; Zhu, H.Y.; Zhang, X.H.; Du, Z.Y.; Xu, Y.J.; Yu, X.D. Apigenin inhibits proliferation and induces apoptosis in human multiple myeloma cells through targeting the trinity of CK2, Cdc37 and Hsp90. Mol. Cancer 2011, 10, 104. [Google Scholar] [CrossRef] [Green Version]
  123. 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] [Green Version]
  124. Li-Weber, M. New therapeutic aspects of flavones: The anticancer properties of Scutellaria and its main active constituents Wogonin, Baicalein and Baicalin. Cancer Treat. Rev. 2009, 35, 57–68. [Google Scholar] [CrossRef]
  125. Yeh, C.H.; Ma, K.H.; Liu, P.S.; Kuo, J.K.; Chueh, S.H. Baicalein decreases hydrogen peroxide-induced damage to NG108-15 cells via upregulation of Nrf2. J. Cell. Physiol. 2015, 230, 1840–1851. [Google Scholar] [CrossRef]
  126. He, X.; Wei, Z.; Zhou, E.; Chen, L.; Kou, J.; Wang, J.; Yang, Z. Baicalein attenuates inflammatory responses by suppressing TLR4 mediated NF-κB and MAPK signaling pathways in LPS-induced mastitis in mice. Int. Immunopharmacol. 2015, 28, 470–476. [Google Scholar] [CrossRef]
  127. Michaelis, M.; Sithisarn, P.; Cinatl, J., Jr. Effects of flavonoid-induced oxidative stress on anti-H5N1 influenza a virus activity exerted by baicalein and biochanin A. BMC Res. Notes 2014, 7, 384. [Google Scholar] [CrossRef] [Green Version]
  128. Wu, J.Y.; Tsai, K.W.; Li, Y.Z.; Chang, Y.S.; Lai, Y.C.; Laio, Y.H.; Wu, J.D.; Liu, Y.W. Anti-bladder-tumor effect of baicalein from Scutellaria baicalensis Georgi and its application in vivo. Evid. Based Complement. Altern. Med. 2013, 2013, 579751. [Google Scholar] [CrossRef] [Green Version]
  129. Guo, Z.; Hu, X.; Xing, Z.; Xing, R.; Lv, R.; Cheng, X.; Su, J.; Zhou, Z.; Xu, Z.; Nilsson, S.; et al. Baicalein inhibits prostate cancer cell growth and metastasis via the caveolin-1/AKT/mTOR pathway. Mol. Cell Biochem. 2015, 406, 111–119. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. 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] [Green Version]
  131. Chandrashekar, N.; Selvamani, A.; Subramanian, R.; Pandi, A.; Thiruvengadam, D. Baicalein inhibits pulmonary carcinogenesis-associated inflammation and interferes with COX-2, MMP-2 and MMP-9 expressions in vivo. Toxicol. Appl. Pharmacol. 2012, 261, 10–21. [Google Scholar] [CrossRef]
  132. Liu, X.P.; He, L.; Zhang, Q.P.; Zeng, X.T.; Liu, S.Q. Baicalein inhibits proliferation of myeloma U266 cells by downregulating IKZF1 and IKZF3. Med. Sci. Monit. 2018, 24, 2809–2817. [Google Scholar] [CrossRef] [Green Version]
  133. Gu, Y.Y.; Liu, L.P.; Qin, J.; Zhang, M.; Chen, Y.; Wang, D.; Li, Z.; Tang, J.Z.; Mo, S.L. Baicalein decreases side population proportion via inhibition of ABCG2 in multiple myeloma cell line RPMI 8226 in vitro. Fitoterapia 2014, 94, 21–28. [Google Scholar] [CrossRef]
  134. Stacy, A.E.; Jansson, P.J.; Richardson, D.R. Molecular pharmacology of ABCG2 and its role in chemoresistance. Mol. Pharmacol. 2013, 84, 655–669. [Google Scholar] [CrossRef] [Green Version]
  135. Lin, M.G.; Liu, L.P.; Li, C.Y.; Zhang, M.; Chen, Y.; Qin, J.; Gu, Y.Y.; Li, Z.; Wu, X.L.; Mo, S.L. Scutellaria extract decreases the proportion of side population cells in a myeloma cell line by down-regulating the expression of ABCG2 protein. Asian Pac. J. Cancer Prev. 2013, 14, 7179–7186. [Google Scholar] [CrossRef]
  136. Zhang, R.B.; He, L.; Huang, Z.; Ma, Z.; Liu, S.Q. Synergistic effect and mechanism of baicalein in combination with lenalidomide-induced apoptosis of myeloma cells. Zhonghua Xue Ye Xue Za Zhi 2013, 34, 546–547. (In Chinese) [Google Scholar] [CrossRef] [PubMed]
  137. Liu, S.; Ma, Z.; Cai, H.; Li, Q.; Rong, W.; Kawano, M. Inhibitory effect of baicalein on IL-6-mediated signaling cascades in human myeloma cells. Eur. J. Haematol. 2010, 84, 137–144. [Google Scholar] [CrossRef] [PubMed]
  138. Otsuyama, K.I.; Ma, Z.; Abroun, S.; Amin, J.; Shamsasenjan, K.; Asaoku, H.; Kawano, M.M. PPAR beta-mediated growth suppression of baicalein and dexamethasone in human myeloma cells. Leukemia 2007, 21, 187–190. [Google Scholar] [CrossRef] [PubMed]
  139. Ma, Z.; Otsuyama, K.; Liu, S.; Abroun, S.; Ishikawa, H.; Tsuyama, N.; Obata, M.; Li, F.J.; Zheng, X.; Maki, Y.; et al. Baicalein, a component of Scutellaria radix from Huang-Lian-Jie-Du-Tang (HLJDT), leads to suppression of proliferation and induction of apoptosis in human myeloma cells. Blood 2005, 105, 3312–3318. [Google Scholar] [CrossRef] [Green Version]
  140. Lin, L.Z.; Lu, S.; Harnly, J.M. Detection and quantification of glycosylated flavonoid malonates in celery, Chinese celery, and celery seed by LC-DAD-ESI/MS. J. Agric. Food Chem. 2007, 55, 1321–1326. [Google Scholar] [CrossRef] [Green Version]
  141. Snijman, P.W.; Swanevelder, S.; Joubert, E.; Green, I.R.; Gelderblom, W.C. The antimutagenic activity of the major flavonoids of rooibos (Aspalathus linearis): Some dose-response effects on mutagen activation-flavonoid interactions. Mutat. Res. 2007, 631, 111–123. [Google Scholar] [CrossRef]
  142. Khan, A.U.; Gilani, A.H. Selective bronchodilatory effect of Rooibos tea (Aspalathus linearis) and its flavonoid, chrysoeriol. Eur. J. Nutr. 2006, 45, 463–469. [Google Scholar] [CrossRef]
  143. Choi, D.Y.; Lee, J.Y.; Kim, M.R.; Woo, E.R.; Kim, Y.G.; Kang, K.W. Chrysoeriol potently inhibits the induction of nitric oxide synthase by blocking AP-1 activation. J. Biomed. Sci. 2005, 12, 949–959. [Google Scholar] [CrossRef]
  144. Duke, J.A.; Bogenschutz, M.J.; du Cellier, J.; Duke, P.A.K. Handbook of Medicinal Herbs, 2nd ed.; CRC Press: Boca Raton, FL, USA, 2002; pp. 612–613. [Google Scholar]
  145. Wu, J.Y.; Chen, Y.J.; Bai, L.; Liu, Y.X.; Fu, X.Q.; Zhu, P.L.; Li, J.K.; Chou, J.Y.; Yin, C.L.; Wang, Y.P.; et al. Chrysoeriol ameliorates TPA-induced acute skin inflammation in mice and inhibits NF-κB and STAT3 pathways. Phytomedicine 2020, 68, 153173. [Google Scholar] [CrossRef]
  146. Kim, J.H.; Cho, Y.H.; Park, S.M.; Lee, K.E.; Lee, J.J.; Lee, B.C.; Pyo, H.B.; Song, K.S.; Park, H.D.; Yun, Y.P. Antioxidants and inhibitor of matrix metalloproteinase-1 expression from leaves of Zostera marina L. Arch. Pharm. Res. 2004, 27, 177–183. [Google Scholar] [CrossRef]
  147. Han, L.K.; Sumiyoshi, M.; Zheng, Y.N.; Okuda, H.; Kimura, Y. Anti-obesity action of Salix matsudana leaves (Part 2). Isolation of anti-obesity effectors from polyphenol fractions of Salix matsudana. Phytother. Res. 2003, 17, 1195–1198. [Google Scholar] [CrossRef]
  148. Schinella, G.R.; Giner, R.M.; Recio, M.C.; Mordujovich de Buschiazzo, P.; Ríos, J.L.; Máñez, S. Anti-inflammatory effects of South American Tanacetum vulgare. J. Pharm. Pharmacol. 1998, 50, 1069–1074. [Google Scholar] [CrossRef]
  149. Wei, W.; He, J.; Ruan, H.; Wang, Y. In vitro and in vivo cytotoxic effects of chrysoeriol in human lung carcinoma are facilitated through activation of autophagy, sub-G1/G0 cell cycle arrest, cell migration and invasion inhibition and modulation of MAPK/ERK signalling pathway. J. BUON 2019, 24, 936–942. [Google Scholar]
  150. Takemura, H.; Uchiyama, H.; Ohura, T.; Sakakibara, H.; Kuruto, R.; Amagai, T.; Shimoi, K. A methoxyflavonoid, chrysoeriol, selectively inhibits the formation of a carcinogenic estrogen metabolite in MCF-7 breast cancer cells. J. Steroid Biochem. Mol. Biol. 2010, 118, 70–76. [Google Scholar] [CrossRef]
  151. Yang, Y.; Zhou, X.; Xiao, M.; Hong, Z.; Gong, Q.; Jiang, L.; Zhou, J. Discovery of chrysoeriol, a PI3K-AKT-mTOR pathway inhibitor with potent antitumor activity against human multiple myeloma cells in vitro. J. Huazhong Univ. Sci. Technol. Med. Sci. 2010, 30, 734–740. [Google Scholar] [CrossRef]
  152. Aziz, N.; Kim, M.Y.; Cho, J.Y. Anti-inflammatory effects of luteolin: A review of in vitro, in vivo, and in silico studies. J. Ethnopharmacol. 2018, 225, 342–358. [Google Scholar] [CrossRef]
  153. Birt, D.F.; Hendrich, S.; Wang, W. Dietary agents in cancer prevention: Flavonoids and isoflavonoids. Pharmacol. Ther. 2001, 90, 157–177. [Google Scholar] [CrossRef]
  154. Sui, J.Q.; Xie, K.P.; Xie, M.J. Inhibitory effect of luteolin on the proliferation of human breast cancer cell lines induced by epidermal growth factor. Sheng Li Xue Bao 2016, 68, 27–34. [Google Scholar]
  155. Cook, M.T.; Liang, Y.; Besch-Williford, C.; Goyette, S.; Mafuvadze, B.; Hyder, S.M. Luteolin inhibits progestin-dependent angiogenesis, stem cell-like characteristics, and growth of human breast cancer xenografts. Springerplus 2015, 4, 444. [Google Scholar] [CrossRef] [Green Version]
  156. Sun, D.W.; Zhang, H.D.; Mao, L.; Mao, C.F.; Chen, W.; Cui, M.; Ma, R.; Cao, H.X.; Jing, C.W.; Wang, Z.; et al. Luteolin inhibits breast cancer development and progression in vitro and in vivo by suppressing Notch signaling and regulating miRNAs. Cell. Physiol. Biochem. 2015, 37, 1693–1711. [Google Scholar] [CrossRef] [Green Version]
  157. Park, S.H.; Ham, S.; Kwon, T.H.; Kim, M.S.; Lee, D.H.; Kang, J.W.; Oh, S.R.; Yoon, D.Y. Luteolin induces cell cycle arrest and apoptosis through extrinsic and intrinsic signaling pathways in MCF-7 breast cancer cells. J. Environ. Pathol. Toxicol. Oncol. 2014, 33, 219–231. [Google Scholar] [CrossRef]
  158. Pandurangan, A.K.; Esa, N.M. Luteolin, a bioflavonoid inhibits colorectal cancer through modulation of multiple signaling pathways: A review. Asian Pac. J. Cancer Prev. 2014, 15, 5501–5508. [Google Scholar] [CrossRef]
  159. Han, K.; Meng, W.; Zhang, J.J.; Zhou, Y.; Wang, Y.L.; Su, Y.; Lin, S.C.; Gan, Z.H.; Sun, Y.N.; Min, D.L. Luteolin inhibited proliferation and induced apoptosis of prostate cancer cells through miR-301. OncoTargets Ther. 2016, 9, 3085–3094. [Google Scholar] [CrossRef] [Green Version]
  160. Pratheeshkumar, P.; Son, Y.O.; Budhraja, A.; Wang, X.; Ding, S.; Wang, L.; Hitron, A.; Lee, J.C.; Kim, D.; Divya, S.P.; et al. Luteolin inhibits human prostate tumor growth by suppressing vascular endothelial growth factor receptor 2-mediated angiogenesis. PLoS ONE 2012, 7, e52279. [Google Scholar] [CrossRef]
  161. Horinaka, M.; Yoshida, T.; Shiraishi, T.; Nakata, S.; Wakada, M.; Nakanishi, R.; Nishino, H.; Sakai, T. The combination of TRAIL and luteolin enhances apoptosis in human cervical cancer HeLa cells. Biochem. Biophys. Res. Commun. 2005, 333, 833–838. [Google Scholar] [CrossRef]
  162. Cheng, W.Y.; Chiao, M.T.; Liang, Y.J.; Yang, Y.C.; Shen, C.C.; Yang, C.Y. Luteolin inhibits migration of human glioblastoma U-87 MG and T98G cells through downregulation of Cdc42 expression and PI3K/AKT activity. Mol. Biol. Rep. 2013, 40, 5315–5326. [Google Scholar] [CrossRef] [Green Version]
  163. Tjioe, K.C.; Tostes Oliveira, D.; Gavard, J. Luteolin impacts on the DNA damage pathway in oral squamous cell carcinoma. Nutr. Cancer 2016, 68, 838–847. [Google Scholar] [CrossRef]
  164. Yang, S.F.; Yang, W.E.; Chang, H.R.; Chu, S.C.; Hsieh, Y.S. Luteolin induces apoptosis in oral squamous cancer cells. J. Dent. Res. 2008, 87, 401–406. [Google Scholar] [CrossRef]
  165. Meng, G.; Chai, K.; Li, X.; Zhu, Y.; Huang, W. Luteolin exerts pro-apoptotic effect and anti-migration effects on A549 lung adenocarcinoma cells through the activation of MEK/ERK signaling pathway. Chem. Biol. Interact. 2016, 257, 26–34. [Google Scholar] [CrossRef]
  166. Ou, Y.C.; Li, J.R.; Kuan, Y.H.; Raung, S.L.; Wang, C.C.; Hung, Y.Y.; Pan, P.H.; Lu, H.C.; Chen, C.J. Luteolin sensitizes human 786-O renal cell carcinoma cells to TRAIL-induced apoptosis. Life Sci. 2014, 100, 110–117. [Google Scholar] [CrossRef]
  167. Wu, H.; Huang, M.; Liu, Y.; Shu, Y.; Liu, P. Luteolin induces apoptosis by up-regulating miR-34a in human gastric cancer cells. Technol. Cancer Res. Treat. 2015, 14, 747–755. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Lu, X.Y.; Li, Y.H.; Xiao, X.W.; Li, X.B. Inhibitory effects of luteolin on human gastric carcinoma xenografts in nude mice and its mechanism. Zhonghua Yi Xue Za Zhi 2013, 93, 142–146. (In Chinese) [Google Scholar] [PubMed]
  169. Zhang, Q.; Wan, L.; Guo, Y.; Cheng, N.; Cheng, W.; Sun, Q.; Zhu, J. Radiosensitization effect of luteolin on human gastric cancer SGC-7901 cells. J. Biol. Regul. Homeost. Agents 2009, 23, 71–78. [Google Scholar] [PubMed]
  170. Xu, H.; Yang, T.; Liu, X.; Tian, Y.; Chen, X.; Yuan, R.; Su, S.; Lin, X.; Du, G. Luteolin synergizes the antitumor effects of 5-fluorouracil against human hepatocellular carcinoma cells through apoptosis induction and metabolism. Life Sci. 2016, 144, 138–147. [Google Scholar] [CrossRef]
  171. Lee, L.T.; Huang, Y.T.; Hwang, J.J.; Lee, P.P.; Ke, F.C.; Nair, M.P.; Kanadaswam, C.; Lee, M.T. Blockade of the epidermal growth factor receptor tyrosine kinase activity by quercetin and luteolin leads to growth inhibition and apoptosis of pancreatic tumor cells. Anticancer Res. 2002, 22, 1615–1627. [Google Scholar]
  172. 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]
  173. Chen, T.; Li, X.F.; Wang, J.F.; Zhou, S.; Fang, F. Effects of luteolin on proliferation and programmed cell death of human multiple myeloma cell RPMI-8226. Zhongguo Shi Yan Xue Ye Xue Za Zhi 2018, 26, 1425–1429. (In Chinese) [Google Scholar] [CrossRef]
  174. Sun, C.Y.; Zhu, Y.; Li, X.F.; Wang, X.Q.; Tang, L.P.; Su, Z.Q.; Li, C.Y.; Zheng, G.J.; Feng, B. Scutellarin increases cisplatin-induced apoptosis and autophagy to overcome cisplatin resistance in non-small cell lung cancer via ERK/p53 and c-met/AKT signaling pathways. Front. Pharmacol. 2018, 9, 92. [Google Scholar] [CrossRef] [Green Version]
  175. Cao, P.; Liu, B.; Du, F.; Li, D.; Wang, Y.; Yan, X.; Li, X.; Li, Y. Scutellarin suppresses proliferation and promotes apoptosis in A549 lung adenocarcinoma cells via AKT/mTOR/4EBP1 and STAT3 pathways. Thorac. Cancer 2019, 10, 492–500. [Google Scholar] [CrossRef]
  176. Wang, Z.; Yu, J.; Wu, J.; Qi, F.; Wang, H.; Wang, Z.; Xu, Z. Scutellarin protects cardiomyocyte ischemia-reperfusion injury by reducing apoptosis and oxidative stress. Life Sci. 2016, 157, 200–207. [Google Scholar] [CrossRef]
  177. Yuan, Y.; Zha, H.; Rangarajan, P.; Ling, E.A.; Wu, C. Anti-inflammatory effects of Edaravone and Scutellarin in activated microglia in experimentally induced ischemia injury in rats and in BV-2 microglia. BMC Neurosci. 2014, 15, 125. [Google Scholar] [CrossRef] [Green Version]
  178. Wang, W.; Ma, X.; Han, J.; Zhou, M.; Ren, H.; Pan, Q.; Zheng, C.; Zheng, Q. Neuroprotective effect of Scutellarin on ischemic cerebral injury by down-regulating the expression of angiotensin-converting enzyme and AT1 receptor. PLoS ONE 2016, 11, e0146197. [Google Scholar] [CrossRef]
  179. Ke, Y.; Bao, T.; Wu, X.; Tang, H.; Wang, Y.; Ge, J.; Fu, B.; Meng, X.; Chen, L.; Zhang, C.; et al. Scutellarin suppresses migration and invasion of human hepatocellular carcinoma by inhibiting the STAT3/Girdin/Akt activity. Biochem. Biophys. Res. Commun. 2017, 483, 509–515. [Google Scholar] [CrossRef]
  180. Zhu, P.T.; Mao, M.; Liu, Z.G.; Tao, L.; Yan, B.C. Scutellarin suppresses human colorectal cancer metastasis and angiogenesis by targeting ephrinb2. Am. J. Transl. Res. 2017, 9, 5094–5104. [Google Scholar]
  181. Li, H.; Huang, D.; Gao, Z.; Chen, Y.; Zhang, L.; Zheng, J. Scutellarin inhibits the growth and invasion of human tongue squamous carcinoma through the inhibition of matrix metalloproteinase-2 and -9 and αvβ6 integrin. Int. J. Oncol. 2013, 42, 1674–1681. [Google Scholar] [CrossRef] [Green Version]
  182. Misso, G.; Zarone, M.R.; Lombardi, A.; Grimaldi, A.; Cossu, A.M.; Ferri, C.; Russo, M.; Vuoso, D.C.; Luce, A.; Kawasaki, H.; et al. miR-125b upregulates miR-34a and sequentially activates stress adaption and cell death mechanisms in multiple myeloma. Mol. Ther. Nucleic Acids 2019, 16, 391–406. [Google Scholar] [CrossRef] [Green Version]
  183. Baumann, S.; Fas, S.C.; Giaisi, M.; Müller, W.W.; Merling, A.; Gülow, K.; Edler, L.; Krammer, P.H.; Li-Weber, M. Wogonin preferentially kills malignant lymphocytes and suppresses T-cell tumor growth by inducing PLCgamma1- and Ca2+-dependent apoptosis. Blood 2008, 111, 2354–2363. [Google Scholar] [CrossRef] [Green Version]
  184. Piao, H.Z.; Jin, S.A.; Chun, H.S.; Lee, J.C.; Kim, W.K. Neuroprotective effect of wogonin: Potential roles of inflammatory cytokines. Arch. Pharm. Res. 2004, 27, 930–936. [Google Scholar] [CrossRef]
  185. Lee, H.; Kim, Y.O.; Kim, H.; Kim, S.Y.; Noh, H.S.; Kang, S.S.; Cho, G.J.; Choi, W.S.; Suk, K. Flavonoid wogonin from medicinal herb is neuroprotective by inhibiting inflammatory activation of microglia. FASEB J. 2003, 17, 1943–1944. [Google Scholar] [CrossRef] [Green Version]
  186. 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] [Green Version]
  187. Himeji, M.; Ohtsuki, T.; Fukazawa, H.; Tanaka, M.; Yazaki, S.; Ui, S.; Nishio, K.; Yamamoto, H.; Tasaka, K.; Mimura, A. Difference of growth-inhibitory effect of Scutellaria baicalensis-producing flavonoid wogonin among human cancer cells and normal diploid cell. Cancer Lett. 2007, 245, 269–274. [Google Scholar] [CrossRef]
  188. Chi, Y.S.; Lim, H.; Park, H.; Kim, H.P. Effects of wogonin, a plant flavone from Scutellaria radix, on skin inflammation: In vivo regulation of inflammation-associated gene expression. Biochem. Pharmacol. 2003, 66, 1271–1278. [Google Scholar] [CrossRef]
  189. Chang, Y.L.; Shen, J.J.; Wung, B.S.; Cheng, J.J.; Wang, D.L. Chinese herbal remedy wogonin inhibits monocyte chemotactic protein-1 gene expression in human endothelial cells. Mol. Pharmacol. 2001, 60, 507–513. [Google Scholar]
  190. 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] [CrossRef]
  191. Kim, H.; Kim, Y.S.; Kim, S.Y.; Suk, K. The plant flavonoid wogonin suppresses death of activated C6 rat glial cells by inhibiting nitric oxide production. Neurosci. Lett. 2001, 309, 67–71. [Google Scholar] [CrossRef]
  192. Wakabayashi, I.; Yasui, K. Wogonin inhibits inducible prostaglandin E(2) production in macrophages. Eur. J. Pharmacol. 2000, 406, 477–481. [Google Scholar] [CrossRef]
  193. Huynh, D.L.; Kwon, T.; Zhang, J.J.; Sharma, N.; Gera, M.; Ghosh, M.; Kim, N.; Cho, S.; Lee, D.S.; Park, Y.H.; et al. Wogonin suppresses stem cell-like traits of CD133 positive osteosarcoma cell via inhibiting matrix metallopeptidase-9 expression. BMC Complement. Altern. Med. 2017, 17, 304. [Google Scholar] [CrossRef] [Green Version]
  194. Huynh, D.L.; Sharma, N.; Kumar Singh, A.; Singh Sodhi, S.; Zhang, J.J.; Mongre, R.K.; Ghosh, M.; Kim, N.; Ho, P.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]
  195. Enomoto, R.; Koshiba, C.; Suzuki, C.; Lee, E. Wogonin potentiates the antitumor action of etoposide and ameliorates its adverse effects. Cancer Chemother. Pharmacol. 2011, 67, 1063–1072. [Google Scholar] [CrossRef]
  196. Bhaskar, A.; Gupta, R.; Vishnubhatla, S.; Kumar, L.; Sharma, A.; Sharma, M.C.; Das, P.; Thakur, S.C. Angiopoietins as biomarker of disease activity and response to therapy in multiple myeloma. Leuk. Lymphoma 2013, 54, 1473–1478. [Google Scholar] [CrossRef]
  197. Fan, G.C. Hypoxic exosomes promote angiogenesis. Blood 2014, 124, 3669–3670. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  198. Fu, R.; Chen, Y.; Wang, X.P.; An, T.; Tao, L.; Zhou, Y.X.; Huang, Y.J.; Chen, B.A.; Li, Z.Y.; You, Q.D.; et al. Wogonin inhibits multiple myeloma-stimulated angiogenesis via c-Myc/VHL/HIF-1α signaling axis. Oncotarget 2016, 7, 5715–5727. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  199. Song, X.; Yao, J.; Wang, F.; Zhou, M.; Zhou, Y.; Wang, H.; Wei, L.; Zhao, L.; Li, Z.; Lu, N.; et al. Wogonin inhibits tumor angiogenesis via degradation of HIF-1α protein. Toxicol. Appl. Pharmacol. 2013, 271, 144–155. [Google Scholar] [CrossRef] [PubMed]
  200. Zhang, J.; Sattler, M.; Tonon, G.; Grabher, C.; Lababidi, S.; Zimmerhackl, A.; Raab, M.S.; Vallet, S.; Zhou, Y.; Cartron, M.A.; et al. Targeting angiogenesis via a c-Myc/hypoxia-inducible factor-1alpha-dependent pathway in multiple myeloma. Cancer Res. 2009, 69, 5082–5090. [Google Scholar] [CrossRef] [Green Version]
  201. Zhang, M.; Liu, L.P.; Chen, Y.; Tian, X.Y.; Qin, J.; Wang, D.; Li, Z.; Mo, S.L. Wogonin induces apoptosis in RPMI 8226, a human myeloma cell line, by downregulating phospho-Akt and overexpressing Bax. Life Sci. 2013, 92, 55–62. [Google Scholar] [CrossRef]
  202. Brazil, D.P.; Park, J.; Hemmings, B.A. PKB binding proteins. Getting in on the Akt. Cell 2002, 111, 293–303. [Google Scholar] [CrossRef] [Green Version]
  203. Grynkiewicz, G.; Demchuk, O.M. New perspectives for fisetin. Front. Chem. 2019, 7, 697. [Google Scholar] [CrossRef]
  204. Jash, S.K.; Mondal, S. Bioactive flavonoid fisetin—A molecule of pharmacological interest. Signpost Open Access J. Org. Biomol. Chem. 2014, 2, 89–128. [Google Scholar]
  205. Harborne, J.B. Progress in the chemistry of organic natural products. Am. Chem. Soc. 1975. [Google Scholar] [CrossRef]
  206. Kimira, M.; Arai, Y.; Shimoi, K.; Watanabe, S. Japanese intake of flavonoids and isoflavonoids from foods. J. Epidemiol. 1998, 8, 168–175. [Google Scholar] [CrossRef] [Green Version]
  207. Jang, K.Y.; Jeong, S.J.; Kim, S.H.; Jung, J.H.; Kim, J.H.; Koh, W.; Chen, C.Y.; Kim, S.H. Activation of reactive oxygen species/AMP activated protein kinase signaling mediates fisetin-induced apoptosis in multiple myeloma U266 cells. Cancer Lett. 2012, 319, 197–202. [Google Scholar] [CrossRef]
  208. Sung, B.; Pandey, M.K.; Aggarwal, B.B. Fisetin, an inhibitor of cyclin-dependent kinase 6, down-regulates nuclear factor-kappaB-regulated cell proliferation, antiapoptotic and metastatic gene products through the suppression of TAK-1 and receptor-interacting protein-regulated IkappaBalpha kinase activation. Mol. Pharmacol. 2007, 71, 1703–1714. [Google Scholar] [CrossRef]
  209. Kashyap, D.; Garg, V.K.; Tuli, H.S.; Yerer, M.B.; Sak, K.; Sharma, A.K.; Kumar, M.; Aggarwal, V.; Sandhu, S.S. Fisetin and quercetin: Promising flavonoids with chemopreventive potential. Biomolecules 2019, 9, 174. [Google Scholar] [CrossRef] [Green Version]
  210. Kashyap, D.; Sharma, A.; Sak, K.; Tuli, H.S.; Buttar, H.S.; Bishayee, A. Fisetin: A bioactive phytochemical with potential for cancer prevention and pharmacotherapy. Life Sci. 2018, 194, 75–87. [Google Scholar] [CrossRef]
  211. Kadari, A.; Gudem, S.; Kulhari, H.; Bhandi, M.M.; Borkar, R.M.; Kolapalli, V.R.; Sistla, R. Enhanced oral bioavailability and anticancer efficacy of fisetin by encapsulating as inclusion complex with HPβCD in polymeric nanoparticles. Drug Deliv. 2017, 24, 224–232. [Google Scholar] [CrossRef] [Green Version]
  212. Mohtar, N.; Taylor, K.M.; Sheikh, K.; Somavarapu, S. Design and development of dry powder sulfobutylether-β-cyclodextrin complex for pulmonary delivery of fisetin. Eur. J. Pharm. Biopharm. 2017, 113, 1–10. [Google Scholar] [CrossRef] [Green Version]
  213. Sechi, M.; Syed, D.N.; Pala, N.; Mariani, A.; Marceddu, S.; Brunetti, A.; Mukhtar, H.; Sanna, V. Nanoencapsulation of dietary flavonoid fisetin: Formulation and in vitro antioxidant and α-glucosidase inhibition activities. Mater. Sci. Eng. C Mater. Biol. Appl. 2016, 68, 594–602. [Google Scholar] [CrossRef]
  214. Sowa, M.; Ślepokura, K.; Matczak-Jon, E. Improving solubility of fisetin by cocrystallization. CrystEngComm 2014, 16, 10592–10601. [Google Scholar] [CrossRef]
  215. Sowa, M.; Ślepokura, K.; Matczak-Jon, E. Cocrystals of fisetin, luteolin and genistein with pyridinecarboxamide coformers: Crystal structures, analysis of intermolecular interactions, spectral and thermal characterization. CrystEngComm 2013, 15, 7696. [Google Scholar] [CrossRef]
  216. Sengupta, B.; Banerjee, A.; Sengupta, P.K. Investigations on the binding and antioxidant properties of the plant flavonoid fisetin in model biomembranes. FEBS Lett. 2004, 570, 77–81. [Google Scholar] [CrossRef] [Green Version]
  217. Cos, P.; Ying, L.; Calomme, M.; Hu, J.P.; Cimanga, K.; Van Poel, B.; Pieters, L.; Vlietinck, A.J.; Berghe, V.D. Structure-activity relationship and classification of flavonoids as inhibitors of xanthine oxidase and superoxide scavengers. J. Nat. Prod. 1998, 61, 71–76. [Google Scholar] [CrossRef]
  218. De Whalley, C.V.; Rankin, S.M.; Hoult, J.R.; Jessup, W.; Leake, D.S. Flavonoids inhibit the oxidative modification of low density lipoproteins by macrophages. Biochem. Pharmacol. 1990, 39, 1743–1750. [Google Scholar] [CrossRef]
  219. Lee, J.D.; Huh, J.E.; Jeon, G.; Yang, H.R.; Woo, H.S.; Choi, D.Y.; Park, D.S. Flavonol-rich RVHxR from Rhus verniciflua Stokes and its major compound fisetin inhibits inflammation-related cytokines and angiogenic factor in rheumatoid arthritic fibroblast-like synovial cells and in vivo models. Int. Immunopharmacol. 2009, 9, 268–276. [Google Scholar] [CrossRef]
  220. Khan, N.; Afaq, F.; Khusro, F.H.; Mustafa Adhami, V.; Suh, Y.; Mukhtar, H. Dual inhibition of phosphatidylinositol 3-kinase/Akt and mammalian target of rapamycin signaling in human nonsmall cell lung cancer cells by a dietary flavonoid fisetin. Int. J. Cancer 2012, 130, 1695–1705. [Google Scholar] [CrossRef] [Green Version]
  221. Suh, Y.; Afaq, F.; Johnson, J.J.; Mukhtar, H. A plant flavonoid fisetin induces apoptosis in colon cancer cells by inhibition of COX2 and Wnt/EGFR/NF-kappaB-signaling pathways. Carcinogenesis 2009, 30, 300–307. [Google Scholar] [CrossRef] [Green Version]
  222. Maurya, B.K.; Trigun, S.K. Fisetin modulates antioxidant enzymes and inflammatory factors to inhibit aflatoxin-B1 induced hepatocellular carcinoma in rats. Oxid. Med. Cell. Longev. 2016, 2016, 1972793. [Google Scholar] [CrossRef] [Green Version]
  223. Ravichandran, N.; Suresh, G.; Ramesh, B.; Siva, G.V. Fisetin, a novel flavonol attenuates benzo(a)pyrene-induced lung carcinogenesis in Swiss albino mice. Food Chem. Toxicol. 2011, 49, 1141–1147. [Google Scholar] [CrossRef]
  224. Pal, H.C.; Athar, M.; Elmets, C.A.; Afaq, F. Fisetin inhibits UVB-induced cutaneous inflammation and activation of PI3K/AKT/NFκB signaling pathways in SKH-1 hairless mice. Photochem. Photobiol. 2015, 91, 225–234. [Google Scholar] [CrossRef]
  225. Wood, J.G.; Rogina, B.; Lavu, S.; Howitz, K.; Helfand, S.L.; Tatar, M.; Sinclair, D. Sirtuin activators mimic caloric restriction and delay ageing in metazoans. Nature 2004, 430, 686–689. [Google Scholar] [CrossRef] [PubMed]
  226. Kang, K.A.; Piao, M.J.; Hewage, S.R.K.M.; Ryu, Y.S.; Oh, M.C.; Kwon, T.K.; Chae, S.; Hyun, J.W. Fisetin induces apoptosis and endoplasmic reticulum stress in human non-small cell lung cancer through inhibition of the MAPK signaling pathway. Tumour Boil. 2016, 37, 9615–9624. [Google Scholar] [CrossRef] [PubMed]
  227. Mann, S.S.; Hammarback, J.A. Molecular characterization of light chain 3. A microtubule binding subunit of MAP1A and MAP1B. J. Biol. Chem. 1994, 269, 11492–11497. [Google Scholar] [CrossRef]
  228. Inkielewicz-Stepniak, I.; Radomski, M.W.; Wozniak, M. Fisetin prevents fluoride and dexamethasone-induced oxidative damage in osteoblast and hippocampal cells. Food Chem. Toxicol. 2012, 50, 583–589. [Google Scholar] [CrossRef]
  229. Liao, Y.C.; Shih, Y.W.; Chao, C.H.; Lee, X.Y.; Chiang, T.A. Involvement of the ERK signaling pathway in fisetin reduces invasion and migration in the human lung cancer cell line A549. J. Agric. Food Chem. 2009, 57, 8933–8941. [Google Scholar] [CrossRef]
  230. Park, H.H.; Lee, S.; Son, H.Y.; Park, S.B.; Kim, M.S.; Choi, E.J.; Singh, T.S.; Ha, J.H.; Lee, M.G.; Kim, J.E.; et al. Flavonoids inhibit histamine release and expression of proinflammatory cytokines in mast cells. Arch. Pharm. Res. 2008, 31, 1303–1311. [Google Scholar] [CrossRef]
  231. Ren, Q.; Guo, F.; Tao, S.; Huang, R.; Ma, L.; Fu, P. Flavonoid fisetin alleviates kidney inflammation and apoptosis via inhibiting Src-mediated NF-κB p65 and MAPK signaling pathways in septic AKI mice. Biomed. Pharmacother. 2020, 122, 109772. [Google Scholar] [CrossRef]
  232. Mukhtar, E.; Adhami, V.M.; Sechi, M.; Mukhtar, H. Dietary flavonoid fisetin binds to β-tubulin and disrupts microtubule dynamics in prostate cancer cells. Cancer Lett. 2015, 367, 173–183. [Google Scholar] [CrossRef] [Green Version]
  233. Hsu, Y.C.; Meng, X.; Ou, L.; Ip, M.M. Activation of the AMP-activated protein kinase-p38 MAP kinase pathway mediates apoptosis induced by conjugated linoleic acid in p53-mutant mouse mammary tumor cells. Cell. Signal. 2010, 22, 590–599. [Google Scholar] [CrossRef] [Green Version]
  234. Kim, H.J.; Kim, S.K.; Kim, B.S.; Lee, S.H.; Park, Y.S.; Park, B.K.; Kim, S.J.; Kim, J.; Choi, C.; Kim, J.S.; et al. Apoptotic effect of quercetin on HT-29 colon cancer cells via the AMPK signaling pathway. J. Agric. Food Chem. 2010, 58, 8643–8650. [Google Scholar] [CrossRef]
  235. Zhang, W.B.; Wang, Z.; Shu, F.; Jin, Y.H.; Liu, H.Y.; Wang, Q.J.; Yang, Y. Activation of AMP-activated protein kinase by temozolomide contributes to apoptosis in glioblastoma cells via p53 activation and mTORC1 inhibition. J. Biol. Chem. 2010, 285, 40461–40471. [Google Scholar] [CrossRef] [Green Version]
  236. Chen, G.; Zhang, X.; Zhao, M.; Wang, Y.; Cheng, X.; Wang, D.; Xu, Y.; Du, Z.; Yu, X. Celastrol targets mitochondrial respiratory chain complex I to induce reactive oxygen species-dependent cytotoxicity in tumor cells. BMC Cancer 2011, 11, 170. [Google Scholar] [CrossRef] [Green Version]
  237. Kim, J.H.; Jeong, S.J.; Kwon, T.R.; Yun, S.M.; Jung, J.H.; Kim, M.; Lee, H.J.; Lee, M.H.; Ko, S.G.; Chen, C.Y.; et al. Cryptotanshinone enhances TNF-α-induced apoptosis in chronic myeloid leukemia KBM-5 cells. Apoptosis 2011, 16, 696–707. [Google Scholar] [CrossRef]
  238. Afroze, N.; Pramodh, S.; Hussain, A.; Waleed, M.; Vakharia, K. A review on myricetin as a potential therapeutic candidate for cancer prevention. 3 Biotech 2020, 10, 211. [Google Scholar] [CrossRef]
  239. Xu, Y.; Xie, Q.; Wu, S.; Yi, D.; Yu, Y.; Liu, S.; Li, S.; Li, Z. Myricetin induces apoptosis via endoplasmic reticulum stress and DNA double-strand breaks in human ovarian cancer cells. Mol. Med. Rep. 2016, 13, 2094–2100. [Google Scholar] [CrossRef] [Green Version]
  240. Zang, W.; Wang, T.; Wang, Y.; Li, M.; Xuan, X.; Ma, Y.; Du, Y.; Liu, K.; Dong, Z.; Zhao, G. Myricetin exerts anti-proliferative, anti-invasive, and pro-apoptotic effects on esophageal carcinoma EC9706 and KYSE30 cells via RSK2. Tumour Biol. 2014, 35, 12583–12592. [Google Scholar] [CrossRef]
  241. Zhang, X.H.; Chen, S.Y.; Tang, L.; Shen, Y.Z.; Luo, L.; Xu, C.W.; Liu, Q.; Li, D. Myricetin induces apoptosis in HepG2 cells through Akt/p70S6K/bad signaling and mitochondrial apoptotic pathway. Anticancer Agents Med. Chem. 2013, 13, 1575–1581. [Google Scholar] [CrossRef]
  242. Phillips, P.A.; Sangwan, V.; Borja-Cacho, D.; Dudeja, V.; Vickers, S.M.; Saluja, A.K. Myricetin induces pancreatic cancer cell death via the induction of apoptosis and inhibition of the phosphatidylinositol 3-kinase (PI3K) signaling pathway. Cancer Lett. 2011, 308, 181–188. [Google Scholar] [CrossRef] [Green Version]
  243. Akhtar, S.; Najafzadeh, M.; Isreb, M.; Newton, L.; Gopalan, R.C.; Anderson, D. Anticancer potential of myricetin bulk and nano forms in vitro in lymphocytes from myeloma patients. Arch. Toxicol. 2021, 95, 337–343. [Google Scholar] [CrossRef]
  244. Anand David, A.V.; Arulmoli, R.; Parasuraman, S. Overviews of biological importance of quercetin: A bioactive flavonoid. Pharmacogn. Rev. 2016, 10, 84–89. [Google Scholar] [CrossRef] [Green Version]
  245. Batiha, G.E.; Beshbishy, A.M.; Ikram, M.; Mulla, Z.S.; El-Hack, M.E.A.; Taha, A.E.; Algammal, A.M.; Elewa, Y.H.A. The pharmacological activity, biochemical properties, and pharmacokinetics of the major natural polyphenolic flavonoid: Quercetin. Foods 2020, 9, 374. [Google Scholar] [CrossRef] [Green Version]
  246. Boots, A.W.; Haenen, G.R.; Bast, A. Health effects of quercetin: From antioxidant to nutraceutical. Eur. J. Pharmacol. 2008, 585, 325–337. [Google Scholar] [CrossRef]
  247. Lakhanpal, P.; Rai, D.K. Quercetin: A versatile flavonoid. Int. J. Med. 2007, 2, 22–37. [Google Scholar] [CrossRef] [Green Version]
  248. Harwood, M.; Danielewska-Nikiel, B.; Borzelleca, J.F.; Flamm, G.W.; Williams, G.M.; Lines, T.C. A critical review of the data related to the safety of quercetin and lack of evidence of in vivo toxicity, including lack of genotoxic/carcinogenic properties. Food Chem. Toxicol. 2007, 45, 2179–2205. [Google Scholar] [CrossRef] [PubMed]
  249. Hirpara, K.V.; Aggarwal, P.; Mukherjee, A.J.; Joshi, N.; Burman, A.C. Quercetin and its derivatives: Synthesis, pharmacological uses with special emphasis on anti-tumor properties and prodrug with enhanced bio-availability. Anticancer Agents Med. Chem. 2009, 9, 138–161. [Google Scholar] [CrossRef] [PubMed]
  250. Ren, K.W.; Li, Y.H.; Wu, G.; Ren, J.Z.; Lu, H.B.; Li, Z.M.; Han, X.W. Quercetin nanoparticles display antitumor activity via proliferation inhibition and apoptosis induction in liver cancer cells. Int. J. Oncol. 2017, 50, 1299–1311. [Google Scholar] [CrossRef]
  251. Ren, M.X.; Deng, X.H.; Ai, F.; Yuan, G.Y.; Song, H.Y. Effect of quercetin on the proliferation of the human ovarian cancer cell line SKOV-3 in vitro. Exp. Ther. Med. 2015, 10, 579–583. [Google Scholar] [CrossRef] [Green Version]
  252. Deng, X.H.; Song, H.Y.; Zhou, Y.F.; Yuan, G.Y.; Zheng, F.J. Effects of quercetin on the proliferation of breast cancer cells and expression of survivin in vitro. Exp. Ther. Med. 2013, 6, 1155–1158. [Google Scholar] [CrossRef] [Green Version]
  253. Lamson, D.W.; Brignall, M.S. Antioxidants and cancer, part 3: Quercetin. Altern. Med. Rev. 2000, 5, 196–208. [Google Scholar]
  254. Li, X.; Zhou, N.; Wang, J.; Liu, Z.; Wang, X.; Zhang, Q.; Liu, Q.; Gao, L.; Wang, R. Quercetin suppresses breast cancer stem cells (CD44+/CD24-) by inhibiting the PI3K/Akt/mTOR-signaling pathway. Life Sci. 2018, 196, 56–62. [Google Scholar] [CrossRef]
  255. Baby, B.; Antony, P.; Vijayan, R. Interactions of quercetin with receptor tyrosine kinases associated with human lung carcinoma. Nat. Prod. Res. 2018, 32, 2928–2931. [Google Scholar] [CrossRef]
  256. Lee, J.; Han, S.I.; Yun, J.H.; Kim, J.H. Quercetin 3-O-glucoside suppresses epidermal growth factor-induced migration by inhibiting EGFR signaling in pancreatic cancer cells. Tumour Biol. 2015, 36, 9385–9393. [Google Scholar] [CrossRef]
  257. Yang, Y.; Wang, T.; Chen, D.; Ma, Q.; Zheng, Y.; Liao, S.; Wang, Y.; Zhang, J. Quercetin preferentially induces apoptosis in KRAS-mutant colorectal cancer cells via JNK signaling pathways. Cell Biol. Int. 2019, 43, 117–124. [Google Scholar] [CrossRef]
  258. Sharmila, G.; Bhat, F.A.; Arunkumar, R.; Elumalai, P.; Raja Singh, P.; Senthilkumar, K.; Arunakaran, J. Chemopreventive effect of quercetin, a natural dietary flavonoid on prostate cancer in in vivo model. Clin. Nutr. 2014, 33, 718–726. [Google Scholar] [CrossRef]
  259. Liu, Y.; Gong, W.; Yang, Z.Y.; Zhou, X.S.; Gong, C.; Zhang, T.R.; Wei, X.; Ma, D.; Ye, F.; Gao, Q.L. Quercetin induces protective autophagy and apoptosis through ER stress via the p-STAT3/Bcl-2 axis in ovarian cancer. Apoptosis 2017, 22, 544–557. [Google Scholar] [CrossRef]
  260. Huang, D.Y.; Dai, Z.R.; Li, W.M.; Wang, R.G.; Yang, S.M. Inhibition of EGF expression and NF-κB activity by treatment with quercetin leads to suppression of angiogenesis in nasopharyngeal carcinoma. Saudi J. Biol. Sci. 2018, 25, 826–831. [Google Scholar] [CrossRef]
  261. Polukonova, N.V.; Navolokin, N.A.; Bucharskaya, A.B.; Mudrak, D.A.; Baryshnikova, M.A.; Stepanova, E.V.; Solomko, E.S.; Polukonova, A.V.; Maslyakova, G.N. The apoptotic activity of flavonoid-containing Gratiola officinalis extract in cell cultures of human kidney cancer. Russ. Open Med. J. 2018. [Google Scholar] [CrossRef] [Green Version]
  262. Li, S.Z.; Qiao, S.F.; Zhang, J.H.; Li, K. Quercetin increase the chemosensitivity of breast cancer cells to doxorubicin via PTEN/Akt pathway. Anticancer Agents Med. Chem. 2015, 15, 1185–1189. [Google Scholar] [CrossRef]
  263. Yuan, Z.; Wang, H.; Hu, Z.; Huang, Y.; Yao, F.; Sun, S.; Wu, B. Quercetin inhibits proliferation and drug resistance in KB/VCR oral cancer cells and enhances its sensitivity to vincristine. Nutr. Cancer 2015, 67, 126–136. [Google Scholar] [CrossRef]
  264. Xavier, C.P.; Lima, C.F.; Rohde, M.; Pereira-Wilson, C. Quercetin enhances 5-fluorouracil-induced apoptosis in MSI colorectal cancer cells through p53 modulation. Cancer Chemother. Pharmacol. 2011, 68, 1449–1457. [Google Scholar] [CrossRef]
  265. Xu, Y.W.; Zou, L.F.; Li, F. Effect of quercetin on proliferation and apoptosis of multiple myeloma cells and its related mechanism. Zhongguo Shi Yan Xue Ye Xue Za Zhi 2020, 28, 1234–1239. (In Chinese) [Google Scholar] [CrossRef]
  266. He, D.; Guo, X.; Zhang, E.; Zi, F.; Chen, J.; Chen, Q.; Lin, X.; Yang, L.; Li, Y.; Wu, W.; et al. Quercetin induces cell apoptosis of myeloma and displays a synergistic effect with dexamethasone in vitro and in vivo xenograft models. Oncotarget 2016, 7, 45489–45499. [Google Scholar] [CrossRef] [Green Version]
  267. Ma, Y.; Jin, Z.; Huang, J.; Zhou, S.; Ye, H.; Jiang, S.; Yu, K. Quercetin suppresses the proliferation of multiple myeloma cells by down-regulating IQ motif-containing GTPase activating protein 1 expression and extracellular signal-regulated kinase activation. Leuk. Lymphoma 2014, 55, 2597–2604. [Google Scholar] [CrossRef] [PubMed]
  268. Chu, C.; Deng, J.; Man, Y.; Qu, Y. Green tea extracts epigallocatechin-3-gallate for different treatments. Biomed. Res. Int. 2017, 2017, 5615647. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  269. Khan, N.; Afaq, F.; Saleem, M.; Ahmad, N.; Mukhtar, H. Targeting multiple signaling pathways by green tea polyphenol (-)-epigallocatechin-3-gallate. Cancer Res. 2006, 66, 2500–2505. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  270. Kondo, T.; Ohta, T.; Igura, K.; Hara, Y.; Kaji, K. Tea catechins inhibit angiogenesis in vitro, measured by human endothelial cell growth, migration and tube formation, through inhibition of VEGF receptor binding. Cancer Lett. 2002, 180, 139–144. [Google Scholar] [CrossRef]
  271. Gan, R.Y.; Li, H.B.; Sui, Z.Q.; Corke, H. Absorption, metabolism, anti-cancer effect and molecular targets of epigallocatechin gallate (EGCG): An updated review. Crit. Rev. Food Sci. Nutr. 2018, 58, 924–941. [Google Scholar] [CrossRef]
  272. Elbling, L.; Herbacek, I.; Weiss, R.M.; Jantschitsch, C.; Micksche, M.; Gerner, C.; Pangratz, H.; Grusch, M.; Knasmüller, S.; Berger, W. Hydrogen peroxide mediates EGCG-induced antioxidant protection in human keratinocytes. Free Radic. Biol. Med. 2010, 49, 1444–1452. [Google Scholar] [CrossRef]
  273. Xu, Y.; Ho, C.T.; Amin, S.G.; Han, C.; Chung, F.L. Inhibition of tobacco-specific nitrosamine-induced lung tumorigenesis in A/J mice by green tea and its major polyphenol as antioxidants. Cancer Res. 1992, 52, 3875–3879. [Google Scholar]
  274. Braicu, C.; Gherman, C.D.; Irimie, A.; Berindan-Neagoe, I. Epigallocatechin-3-Gallate (EGCG) inhibits cell proliferation and migratory behaviour of triple negative breast cancer cells. J. Nanosci. Nanotechnol. 2013, 13, 632–637. [Google Scholar] [CrossRef]
  275. Shankar, S.; Marsh, L.; Srivastava, R.K. EGCG inhibits growth of human pancreatic tumors orthotopically implanted in Balb C nude mice through modulation of FKHRL1/FOXO3a and neuropilin. Mol. Cell Biochem. 2013, 372, 83–94. [Google Scholar] [CrossRef]
  276. Mantena, S.K.; Roy, A.M.; Katiyar, S.K. Epigallocatechin-3-gallate inhibits photocarcinogenesis through inhibition of angiogenic factors and activation of CD8+ T cells in tumors. Photochem. Photobiol. 2005, 81, 1174–1179. [Google Scholar] [CrossRef]
  277. Jung, Y.D.; Kim, M.S.; Shin, B.A.; Chay, K.O.; Ahn, B.W.; Liu, W.; Bucana, C.D.; Gallick, G.E.; Ellis, L.M. EGCG, a major component of green tea, inhibits tumour growth by inhibiting VEGF induction in human colon carcinoma cells. Br. J. Cancer 2001, 84, 844–850. [Google Scholar] [CrossRef] [Green Version]
  278. Koh, Y.W.; Choi, E.C.; Kang, S.U.; Hwang, H.S.; Lee, M.H.; Pyun, J.; Park, R.; Lee, Y.; Kim, C.H. Green tea (-)-epigallocatechin-3-gallate inhibits HGF-induced progression in oral cavity cancer through suppression of HGF/c-Met. J. Nutr. Biochem. 2011, 22, 1074–1083. [Google Scholar] [CrossRef]
  279. Kwak, I.H.; Shin, Y.H.; Kim, M.; Cha, H.Y.; Nam, H.J.; Lee, B.S.; Chaudhary, S.C.; Pai, K.S.; Lee, J.H. Epigallocatechin-3-gallate inhibits paracrine and autocrine hepatocyte growth factor/scatter factor-induced tumor cell migration and invasion. Exp. Mol. Med. 2011, 43, 111–120. [Google Scholar] [CrossRef]
  280. Kushima, Y.; Iida, K.; Nagaoka, Y.; Kawaratani, Y.; Shirahama, T.; Sakaguchi, M.; Baba, K.; Hara, Y.; Uesato, S. Inhibitory effect of (-)-epigallocatechin and (-)-epigallocatechin gallate against heregulin beta1-induced migration/invasion of the MCF-7 breast carcinoma cell line. Biol. Pharm. Bull. 2009, 32, 899–904. [Google Scholar] [CrossRef] [Green Version]
  281. Lim, Y.C.; Park, H.Y.; Hwang, H.S.; Kang, S.U.; Pyun, J.H.; Lee, M.H.; Choi, E.C.; Kim, C.H. (-)-Epigallocatechin-3-gallate (EGCG) inhibits HGF-induced invasion and metastasis in hypopharyngeal carcinoma cells. Cancer Lett. 2008, 271, 140–152. [Google Scholar] [CrossRef]
  282. Zubair, H.; Azim, S.; Ahmad, A.; Khan, M.A.; Patel, G.K.; Singh, S.; Singh, A.P. Cancer chemoprevention by phytochemicals: Nature’s healing touch. Molecules 2017, 22, 395. [Google Scholar] [CrossRef] [Green Version]
  283. Jin, G.; Yang, Y.; Liu, K.; Zhao, J.; Chen, X.; Liu, H.; Bai, R.; Li, X.; Jiang, Y.; Zhang, X.; et al. Combination curcumin and (-)-epigallocatechin-3-gallate inhibits colorectal carcinoma microenvironment-induced angiogenesis by JAK/STAT3/IL-8 pathway. Oncogenesis 2017, 6, e384. [Google Scholar] [CrossRef]
  284. Chan, M.M.; Chen, R.; Fong, D. Targeting cancer stem cells with dietary phytochemical—Repositioned drug combinations. Cancer Lett. 2018, 433, 53–64. [Google Scholar] [CrossRef]
  285. Wang, W.; Chen, D.; Zhu, K. SOX2OT variant 7 contributes to the synergistic interaction between EGCG and Doxorubicin to kill osteosarcoma via autophagy and stemness inhibition. J. Exp. Clin. Cancer Res. 2018, 37, 37. [Google Scholar] [CrossRef] [Green Version]
  286. Zhou, Y.; Tang, J.; Du, Y.; Ding, J.; Liu, J.Y. The green tea polyphenol EGCG potentiates the antiproliferative activity of sunitinib in human cancer cells. Tumour Biol. 2016, 37, 8555–8566. [Google Scholar] [CrossRef]
  287. Mayr, C.; Wagner, A.; Neureiter, D.; Pichler, M.; Jakab, M.; Illig, R.; Berr, F.; Kiesslich, T. The green tea catechin epigallocatechin gallate induces cell cycle arrest and shows potential synergism with cisplatin in biliary tract cancer cells. BMC Complement. Altern. Med. 2015, 15, 194. [Google Scholar] [CrossRef] [Green Version]
  288. Hajipour, H.; Hamishehkar, H.; Nazari Soltan Ahmad, S.; Barghi, S.; Maroufi, N.F.; Taheri, R.A. Improved anticancer effects of epigallocatechin gallate using rgd-containing nanostructured lipid carriers. Artif. Cells Nanomed. Biotechnol. 2018. [Google Scholar] [CrossRef] [Green Version]
  289. Yuan, X.; He, Y.; Zhou, G.; Li, X.; Feng, A.; Zheng, W. Target challenging-cancer drug delivery to gastric cancer tissues with a fucose graft epigallocatechin-3-gallate-gold particles nanocomposite approach. J. Photochem. Photobiol. 2018, 183, 147–153. [Google Scholar] [CrossRef]
  290. Sanna, V.; Singh, C.K.; Jashari, R.; Adhami, V.M.; Chamcheu, J.C.; Rady, I.; Sechi, M.; Mukhtar, H.; Siddiqui, I.A. Targeted nanoparticles encapsulating (-)-epigallocatechin-3-gallate for prostate cancer prevention and therapy. Sci. Rep. 2017, 7, 41573. [Google Scholar] [CrossRef] [Green Version]
  291. Krupkova, O.; Ferguson, S.J.; Wuertz-Kozak, K. Stability of (-)-epigallocatechin gallate and its activity in liquid formulations and delivery systems. J. Nutr. Biochem. 2016, 37, 1–12. [Google Scholar] [CrossRef]
  292. Luo, K.W.; Lung, W.Y.; Chun, X.; Luo, X.L.; Huang, W.R. EGCG inhibited bladder cancer T24 and 5637 cell proliferation and migration via PI3K/AKT pathway. Oncotarget 2018, 9, 12261–12272. [Google Scholar] [CrossRef] [Green Version]
  293. Yang, C.S.; Wang, H. Cancer preventive activities of tea catechins. Molecules 2016, 21, 1679. [Google Scholar] [CrossRef]
  294. Gu, J.J.; Qiao, K.S.; Sun, P.; Chen, P.; Li, Q. Study of EGCG induced apoptosis in lung cancer cells by inhibiting PI3K/Akt signaling pathway. Eur. Rev. Med. Pharmacol. Sci. 2018, 22, 4557–4563. [Google Scholar] [CrossRef]
  295. Olotu, F.A.; Agoni, C.; Adeniji, E.; Abdullahi, M.; Soliman, M.E. Probing gallate-mediated selectivity and high-affinity binding of epigallocatechin gallate: A way-forward in the design of selective inhibitors for anti-apoptotic Bcl-2 proteins. Appl. Biochem. Biotechnol. 2019, 187, 1061–1080. [Google Scholar] [CrossRef]
  296. Velavan, B.; Divya, T.; Sureshkumar, A.; Sudhandiran, G. Nano-chemotherapeutic efficacy of (-)-epigallocatechin 3-gallate mediating apoptosis in A549 cells: Involvement of reactive oxygen species mediated Nrf2/Keap1signaling. Biochem. Biophys. Res. Commun. 2018, 503, 1723–1731. [Google Scholar] [CrossRef]
  297. Shammas, M.A.; Neri, P.; Koley, H.; Batchu, R.B.; Bertheau, R.C.; Munshi, V.; Prabhala, R.; Fulciniti, M.; Tai, Y.T.; Treon, S.P.; et al. Specific killing of multiple myeloma cells by (-)-epigallocatechin-3-gallate extracted from green tea: Biologic activity and therapeutic implications. Blood 2006, 108, 2804–2810. [Google Scholar] [CrossRef] [PubMed]
  298. Nakazato, T.; Ito, K.; Ikeda, Y.; Kizaki, M. Green tea component, catechin, induces apoptosis of human malignant B cells via production of reactive oxygen species. Clin. Cancer Res. 2005, 11, 6040–6049. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  299. Kumazoe, M.; Fujimura, Y.; Hidaka, S.; Kim, Y.; Murayama, K.; Takai, M.; Huang, Y.; Yamashita, S.; Murata, M.; Miura, D.; et al. Metabolic profiling-based data-mining for an effective chemical combination to induce apoptosis of cancer cells. Sci. Rep. 2015, 5, 9474. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  300. Tsukamoto, S.; Hirotsu, K.; Kumazoe, M.; Goto, Y.; Sugihara, K.; Suda, T.; Tsurudome, Y.; Suzuki, T.; Yamashita, S.; Kim, Y.; et al. Green tea polyphenol EGCG induces lipid-raft clustering and apoptotic cell death by activating protein kinase Cδ and acid sphingomyelinase through a 67 kDa laminin receptor in multiple myeloma cells. Biochem. J. 2012, 443, 525–534. [Google Scholar] [CrossRef]
  301. Zhou, C.G.; Hui, L.M.; Luo, J.M. Epigallocatechin gallate inhibits the proliferation and induces apoptosis of multiple myeloma cells via inactivating EZH2. Eur. Rev. Med. Pharmacol. Sci. 2018, 22, 2093–2098. [Google Scholar] [CrossRef]
  302. Wang, Q.; Li, J.; Gu, J.; Huang, B.; Zhao, Y.; Zheng, D.; Ding, Y.; Zeng, L. Potentiation of (-)-epigallocatechin-3-gallate-induced apoptosis by bortezomib in multiple myeloma cells. Acta Biochim. Biophys. Sin. 2009, 41, 1018–1026. [Google Scholar] [CrossRef] [Green Version]
  303. Tsukamoto, S.; Huang, Y.; Kumazoe, M.; Lesnick, C.; Yamada, S.; Ueda, N.; Suzuki, T.; Yamashita, S.; Kim, Y.H.; Fujimura, Y.; et al. Sphingosine kinase-1 protects multiple myeloma from apoptosis driven by cancer-specific inhibition of RTKs. Mol. Cancer Ther. 2015, 14, 2303–2312. [Google Scholar] [CrossRef] [Green Version]
  304. Bae, J.; Kumazoe, M.; Yamashita, S.; Tachibana, H. Hydrogen sulphide donors selectively potentiate a green tea polyphenol EGCG-induced apoptosis of multiple myeloma cells. Sci. Rep. 2017, 7, 6665. [Google Scholar] [CrossRef]
  305. James, K.D.; Kennett, M.J.; Lambert, J.D. Potential role of the mitochondria as a target for the hepatotoxic effects of (-)-epigallocatechin-3-gallate in mice. Food Chem. Toxicol. 2018, 111, 302–309. [Google Scholar] [CrossRef]
  306. Church, R.J.; Gatti, D.M.; Urban, T.J.; Long, N.; Yang, X.; Shi, Q.; Eaddy, J.S.; Mosedale, M.; Ballard, S.; Churchill, G.A.; et al. Sensitivity to hepatotoxicity due to epigallocatechin gallate is affected by genetic background in diversity outbred mice. Food Chem. Toxicol. 2015, 76, 19–26. [Google Scholar] [CrossRef]
  307. Lambert, J.D.; Kennett, M.J.; Sang, S.; Reuhl, K.R.; Ju, J.; Yang, C.S. Hepatotoxicity of high oral dose (-)-epigallocatechin-3-gallate in mice. Food Chem. Toxicol. 2010, 48, 409–416. [Google Scholar] [CrossRef] [Green Version]
  308. Mazzanti, G.; Menniti-Ippolito, F.; Moro, P.A.; Cassetti, F.; Raschetti, R.; Santuccio, C.; Mastrangelo, S. Hepatotoxicity from green tea: A review of the literature and two unpublished cases. Eur. J. Clin. Pharmacol. 2009, 65, 331–341. [Google Scholar] [CrossRef]
  309. Bae, J.; Kumazoe, M.; Takeuchi, C.; Hidaka, S.; Fujimura, Y.; Tachibana, H. Epigallocatechin-3-O-gallate induces acid sphingomyelinase activation through activation of phospholipase C. Biochem. Biophys. Res. Commun. 2019, 520, 186–191. [Google Scholar] [CrossRef]
  310. Xiao, F.; Cui, H.; Zhong, X. Beneficial effect of daidzin in dry eye rat model through the suppression of inflammation and oxidative stress in the cornea. Saudi J. Biol. Sci. 2018, 25, 832–837. [Google Scholar] [CrossRef]
  311. Isoda, H.; Talorete, T.P.; Kimura, M.; Maekawa, T.; Inamori, Y.; Nakajima, N.; Seki, H. Phytoestrogens genistein and daidzin enhance the acetylcholinesterase activity of the rat pheochromocytoma cell line PC12 by binding to the estrogen receptor. Cytotechnology 2002, 40, 117–123. [Google Scholar] [CrossRef]
  312. Hua, F.; Li, C.H.; Chen, X.G.; Liu, X.P. Daidzein exerts anticancer activity towards SKOV3 human ovarian cancer cells by inducing apoptosis and cell cycle arrest, and inhibiting the Raf/MEK/ERK cascade. Int. J. Mol. Med. 2018, 41, 3485–3492. [Google Scholar] [CrossRef]
  313. Kato, K.; Takahashi, S.; Cui, L.; Toda, T.; Suzuki, S.; Futakuchi, M.; Sugiura, S.; Shirai, T. Suppressive effects of dietary genistin and daidzin on rat prostate carcinogenesis. Jpn. J. Cancer Res. 2000, 91, 786–791. [Google Scholar] [CrossRef]
  314. Pyo, Y.H.; Seong, K.S. Hypolipidemic effects of Monascus-fermented soybean extracts in rats fed a high-fat and -cholesterol diet. J. Agric. Food Chem. 2009, 57, 8617–8622. [Google Scholar] [CrossRef]
  315. Yusakul, G.; Sakamoto, S.; Juengwatanatrakul, T.; Putalun, W.; Tanaka, H.; Morimoto, S. Preparation and application of a monoclonal antibody against the isoflavone glycoside daidzin using a mannich reaction-derived hapten conjugate. Phytochem. Anal. 2016, 27, 81–88. [Google Scholar] [CrossRef]
  316. Choi, H.; Tostes, R.C.; Webb, R.C. Mitochondrial aldehyde dehydrogenase prevents ROS-induced vascular contraction in angiotensin-II hypertensive mice. J. Am. Soc. Hypertens. 2011, 5, 154–160. [Google Scholar] [CrossRef] [Green Version]
  317. Li, X.H.; Zhang, J.C.; Sui, S.F.; Yang, M.S. Effect of daidzin, genistin, and glycitin on osteogenic and adipogenic differentiation of bone marrow stromal cells and adipocytic transdifferentiation of osteoblasts. Acta Pharmacol. Sin. 2005, 26, 1081–1086. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  318. Zhu, Y.; Yao, Y.; Shi, Z.; Everaert, N.; Ren, G. Synergistic effect of bioactive anticarcinogens from soybean on anti-proliferative activity in MDA-MB-231 and MCF-7 human breast cancer cells in vitro. Molecules 2018, 23, 1557. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  319. Johnson, K.A.; Vemuri, S.; Alsahafi, S.; Castillo, R.; Cheriyath, V. Glycone-rich soy isoflavone extracts promote estrogen receptor positive breast cancer cell growth. Nutr. Cancer 2016, 68, 622–633. [Google Scholar] [CrossRef] [PubMed]
  320. Yang, M.H.; Jung, S.H.; Chinnathambi, A.; Alahmadi, T.A.; Alharbi, S.A.; Sethi, G.; Ahn, K.S. Attenuation of STAT3 signaling cascade by daidzin can enhance the apoptotic potential of bortezomib against multiple myeloma. Biomolecules 2019, 10, 23. [Google Scholar] [CrossRef] [Green Version]
  321. Kim, C.; Lee, J.H.; Ko, J.H.; Chinnathambi, A.; Alharbi, S.A.; Shair, O.H.M.; Sethi, G.; Ahn, K.S. Formononetin regulates multiple oncogenic signaling cascades and enhances sensitivity to bortezomib in a multiple myeloma mouse model. Biomolecules 2019, 9, 262. [Google Scholar] [CrossRef] [Green Version]
  322. Sun, D.; Lu, Y.; Zhang, S.J.; Wang, K.G.; Sun, Z. Research on the effect of formononetin on photodynamic therapy in K562 cells. Gen. Physiol. Biophys. 2017, 36, 423–430. [Google Scholar] [CrossRef]
  323. Wu, Y.; Zhang, X.; Li, Z.; Yan, H.; Qin, J.; Li, T. Formononetin inhibits human bladder cancer cell proliferation and invasiveness via regulation of miR-21 and PTEN. Food Funct. 2017, 8, 1061–1066. [Google Scholar] [CrossRef]
  324. Li, T.; Zhao, X.; Mo, Z.; Huang, W.; Yan, H.; Ling, Z.; Ye, Y. Formononetin promotes cell cycle arrest via downregulation of Akt/Cyclin D1/CDK4 in human prostate cancer cells. Cell. Physiol. Biochem. 2014, 34, 1351–1358. [Google Scholar] [CrossRef]
  325. Zhou, R.; Xu, L.; Ye, M.; Liao, M.; Du, H.; Chen, H. Formononetin inhibits migration and invasion of MDA-MB-231 and 4T1 breast cancer cells by suppressing MMP-2 and MMP-9 through PI3K/AKT signaling pathways. Horm. Metab. Res. 2014, 46, 753–760. [Google Scholar] [CrossRef] [Green Version]
  326. Yang, Y.; Zhao, Y.; Ai, X.; Cheng, B.; Lu, S. Formononetin suppresses the proliferation of human non-small cell lung cancer through induction of cell cycle arrest and apoptosis. Int. J. Clin. Exp. Pathol. 2014, 7, 8453–8461. [Google Scholar]
  327. Wang, H.; Zhang, D.; Ge, M.; Li, Z.; Jiang, J.; Li, Y. Formononetin inhibits enterovirus 71 replication by regulating COX-2/PGE2 expression. Virol. J. 2015, 12, 35. [Google Scholar] [CrossRef] [Green Version]
  328. Ma, Z.; Ji, W.; Fu, Q.; Ma, S. Formononetin inhibited the inflammation of LPS-induced acute lung injury in mice associated with induction of PPAR gamma expression. Inflammation 2013, 36, 1560–1566. [Google Scholar] [CrossRef]
  329. Tian, Z.; Liu, S.B.; Wang, Y.C.; Li, X.Q.; Zheng, L.H.; Zhao, M.G. Neuroprotective effects of formononetin against NMDA-induced apoptosis in cortical neurons. Phytother. Res. 2013, 27, 1770–1775. [Google Scholar] [CrossRef]
  330. Huh, J.E.; Nam, D.W.; Baek, Y.H.; Kang, J.W.; Park, D.S.; Choi, D.Y.; Lee, J.D. Formononetin accelerates wound repair by the regulation of early growth response factor-1 transcription factor through the phosphorylation of the ERK and p38 MAPK pathways. Int. Immunopharmacol. 2011, 11, 46–54. [Google Scholar] [CrossRef]
  331. Kim, C.; Lee, S.G.; Yang, W.M.; Arfuso, F.; Um, J.Y.; Kumar, A.P.; Bian, J.; Sethi, G.; Ahn, K.S. Formononetin-induced oxidative stress abrogates the activation of STAT3/5 signaling axis and suppresses the tumor growth in multiple myeloma preclinical model. Cancer Lett. 2018, 431, 123–141. [Google Scholar] [CrossRef]
  332. Fan, F.; Bashari, M.H.; Morelli, E.; Tonon, G.; Malvestiti, S.; Vallet, S.; Jarahian, M.; Seckinger, A.; Hose, D.; Bakiri, L.; et al. The AP-1 transcription factor JunB is essential for multiple myeloma cell proliferation and drug resistance in the bone marrow microenvironment. Leukemia 2017, 31, 1570–1581. [Google Scholar] [CrossRef]
  333. Takeda, T.; Tsubaki, M.; Kino, T.; Kawamura, A.; Isoyama, S.; Itoh, T.; Imano, M.; Tanabe, G.; Muraoka, O.; Matsuda, H.; et al. Mangiferin enhances the sensitivity of human multiple myeloma cells to anticancer drugs through suppression of the nuclear factor κB pathway. Int. J. Oncol. 2016, 48, 2704–2712. [Google Scholar] [CrossRef] [Green Version]
  334. Ososki, A.L.; Kennelly, E.J. Phytoestrogens: A review of the present state of research. Phytother. Res. 2003, 17, 845–869. [Google Scholar] [CrossRef]
  335. Jagadeesh, S.; Kyo, S.; Banerjee, P.P. Genistein represses telomerase activity via both transcriptional and posttranslational mechanisms in human prostate cancer cells. Cancer Res. 2006, 66, 2107–2115. [Google Scholar] [CrossRef] [Green Version]
  336. Davis, J.N.; Kucuk, O.; Sarkar, F.H. Genistein inhibits NF-kappa B activation in prostate cancer cells. Nutr. Cancer 1999, 35, 167–174. [Google Scholar] [CrossRef]
  337. Lee, J.Y.; Kim, H.S.; Song, Y.S. Genistein as a potential anticancer agent against ovarian cancer. J. Tradit. Complement. Med. 2012, 2, 96–104. [Google Scholar] [CrossRef] [Green Version]
  338. Ouyang, G.; Yao, L.; Ruan, K.; Song, G.; Mao, Y.; Bao, S. Genistein induces G2/M cell cycle arrest and apoptosis of human ovarian cancer cells via activation of DNA damage checkpoint pathways. Cell Biol. Int. 2009, 33, 1237–1244. [Google Scholar] [CrossRef] [PubMed]
  339. Choi, E.J.; Kim, T.; Lee, M.S. Pro-apoptotic effect and cytotoxicity of genistein and genistin in human ovarian cancer SK-OV-3 cells. Life Sci. 2007, 80, 1403–1408. [Google Scholar] [CrossRef] [PubMed]
  340. Huang, W.; Wan, C.; Luo, Q.; Huang, Z.; Luo, Q. Genistein-inhibited cancer stem cell-like properties and reduced chemoresistance of gastric cancer. Int. J. Mol. Sci. 2014, 15, 3432–3443. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  341. 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] [Green Version]
  342. Dai, W.; Wang, F.; He, L.; Lin, C.; Wu, S.; Chen, P.; Zhang, Y.; Shen, M.; Wu, D.; Wang, C.; et al. Genistein inhibits hepatocellular carcinoma cell migration by reversing the epithelial-mesenchymal transition: Partial mediation by the transcription factor NFAT1. Mol. Carcinog. 2015, 54, 301–311. [Google Scholar] [CrossRef]
  343. George, J.; Banik, N.L.; Ray, S.K. Genistein induces receptor and mitochondrial pathways and increases apoptosis during BCL-2 knockdown in human malignant neuroblastoma SK-N-DZ cells. J. Neurosci. Res. 2010, 88, 877–886. [Google Scholar] [CrossRef] [Green Version]
  344. Xie, J.; Wang, J.; Zhu, B. Genistein inhibits the proliferation of human multiple myeloma cells through suppression of nuclear factor-κB and upregulation of microRNA-29b. Mol. Med. Rep. 2016, 13, 1627–1632. [Google Scholar] [CrossRef] [Green Version]
  345. He, H.; Chen, L.; Zhai, M.; Chen, J.Z. Genistein down-regulates the constitutive activation of nuclear factor-kappaB in human multiple myeloma cells, leading to suppression of proliferation and induction of apoptosis. Phytother. Res. 2009, 23, 868–873. [Google Scholar] [CrossRef]
  346. Kłósek, M.; Mertas, A.; Król, W.; Jaworska, D.; Szymszal, J.; Szliszka, E. Tumor necrosis factor-related apoptosis-inducing ligand-induced apoptosis in prostate cancer cells after treatment with xanthohumol-a natural compound present in Humulus lupulus L. Int. J. Mol. Sci. 2016, 17, 837. [Google Scholar] [CrossRef] [Green Version]
  347. Kuete, V.; Sandjo, L.P. Isobavachalcone: An overview. Chin. J. Integr. Med. 2012, 18, 543–547. [Google Scholar] [CrossRef]
  348. Sahu, N.K.; Balbhadra, S.S.; Choudhary, J.; Kohli, D.V. Exploring pharmacological significance of chalcone scaffold: A review. Curr. Med. Chem. 2012, 19, 209–225. [Google Scholar] [CrossRef]
  349. Ashour, H.F.; Abou-Zeid, L.A.; El-Sayed, M.A.; Selim, K.B. 1,2,3-Triazole-Chalcone hybrids: Synthesis, in vitro cytotoxic activity and mechanistic investigation of apoptosis induction in multiple myeloma RPMI-8226. Eur. J. Med. Chem. 2020, 189, 112062. [Google Scholar] [CrossRef]
  350. Nowakowska, Z. A review of anti-infective and anti-inflammatory chalcones. Eur. J. Med. Chem. 2007, 42, 125–137. [Google Scholar] [CrossRef]
  351. Rozmer, Z.; Perjési, P. Naturally occurring chalcones and their biological activities. Phytochem. Rev. 2014, 15, 87–120. [Google Scholar] [CrossRef]
  352. Wang, G.; Qiu, J.; Xiao, X.; Cao, A.; Zhou, F. Synthesis, biological evaluation and molecular docking studies of a new series of chalcones containing naphthalene moiety as anticancer agents. Bioorg. Chem. 2018, 76, 249–257. [Google Scholar] [CrossRef]
  353. Ušjak, D.; Ivković, B.; Božić, D.D.; Bošković, L.; Milenković, M. Antimicrobial activity of novel chalcones and modulation of virulence factors in hospital strains of Acinetobacter baumannii and Pseudomonas aeruginosa. Microb. Pathog. 2019, 131, 186–196. [Google Scholar] [CrossRef]
  354. Castaño, L.F.; Cuartas, V.; Bernal, A.; Insuasty, A.; Guzman, J.; Vidal, O.; Rubio, V.; Puerto, G.; Lukáč, P.; Vimberg, V.; et al. New chalcone-sulfonamide hybrids exhibiting anticancer and antituberculosis activity. Eur. J. Med. Chem. 2019, 176, 50–60. [Google Scholar] [CrossRef]
  355. Rashid, H.U.; Xu, Y.; Ahmad, N.; Muhammad, Y.; Wang, L. Promising anti-inflammatory effects of chalcones via inhibition of cyclooxygenase, prostaglandin E2, inducible NO synthase and nuclear factor κb activities. Bioorg. Chem. 2019, 87, 335–365. [Google Scholar] [CrossRef]
  356. Polo, E.; Ibarra-Arellano, N.; Prent-Peñaloza, L.; Morales-Bayuelo, A.; Henao, J.; Galdámez, A.; Gutiérrez, M. Ultrasound-assisted synthesis of novel chalcone, heterochalcone and bis-chalcone derivatives and the evaluation of their antioxidant properties and as acetylcholinesterase inhibitors. Bioorg. Chem. 2019, 90, 103034. [Google Scholar] [CrossRef]
  357. Mathew, B.; Adeniyi, A.; Joy, M.; Mathew, G.E.; Singh-Pillay, A.; Sudarsanakumar, C.; Soliman, M.; Suresh, J. Anti-oxidant behavior of functionalized chalcone-a combined quantum chemical and crystallographic structural investigation. J. Mol. Struct. 2017, 1146, 301–308. [Google Scholar] [CrossRef]
  358. Shin, J.; Jang, M.G.; Park, J.C.; Koo, Y.D.; Lee, J.Y.; Park, K.S.; Chung, S.S.; Park, K. Antidiabetic effects of trihydroxychalcone derivatives via activation of AMP-activated protein kinase. J. Ind. Eng. Chem. 2018, 60, 177–184. [Google Scholar] [CrossRef]
  359. Kumar, H.; Devaraji, V.; Joshi, R.; Jadhao, M.; Ahirkar, P.; Prasath, R.; Bhavana, P.; Ghosh, S.K. Antihypertensive activity of a quinoline appended chalcone derivative and its site specific binding interaction with a relevant target carrier protein. RSC Adv. 2015, 5, 65496–65513. [Google Scholar] [CrossRef]
  360. Orlikova, B.; Tasdemir, D.; Golais, F.; Dicato, M.; Diederich, M. Dietary chalcones with chemopreventive and chemotherapeutic potential. Genes Nutr. 2011, 6, 125–147. [Google Scholar] [CrossRef] [Green Version]
  361. Das, M.; Manna, K. Chalcone scaffold in anticancer armamentarium: A molecular insight. J. Toxicol. 2016, 2016, 7651047. [Google Scholar] [CrossRef] [Green Version]
  362. Padmavathi, G.; Rathnakaram, S.R.; Monisha, J.; Bordoloi, D.; Roy, N.K.; Kunnumakkara, A.B. Potential of butein, a tetrahydroxychalcone to obliterate cancer. Phytomedicine 2015, 22, 1163–1171. [Google Scholar] [CrossRef]
  363. Kang, H.M.; Kim, J.H.; Lee, M.Y.; Son, K.H.; Yang, D.C.; Baek, N.I.; Kwon, B.M. Relationship between flavonoid structure and inhibition of farnesyl protein transferase. Nat. Prod. Res. 2004, 18, 349–356. [Google Scholar] [CrossRef]
  364. Liu, S.C.; Chen, C.; Chung, C.H.; Wang, P.C.; Wu, N.L.; Cheng, J.K.; Lai, Y.W.; Sun, H.L.; Peng, C.Y.; Tang, C.H.; et al. Inhibitory effects of butein on cancer metastasis and bioenergetic modulation. J. Agric. Food Chem. 2014, 62, 9109–9117. [Google Scholar] [CrossRef]
  365. Pandey, M.K.; Sung, B.; Ahn, K.S.; Aggarwal, B.B. Butein suppresses constitutive and inducible signal transducer and activator of transcription (STAT) 3 activation and STAT3-regulated gene products through the induction of a protein tyrosine phosphatase SHP-1. Mol. Pharmacol. 2009, 75, 525–533. [Google Scholar] [CrossRef] [Green Version]
  366. Bowman, T.; Garcia, R.; Turkson, J.; Jove, R. STATs in oncogenesis. Oncogene 2000, 19, 2474–2488. [Google Scholar] [CrossRef] [Green Version]
  367. Cavo, M. Proteasome inhibitor bortezomib for the treatment of multiple myeloma. Leukemia 2006, 20, 1341–1352. [Google Scholar] [CrossRef] [PubMed]
  368. Nawaz, J.; Rasul, A.; Shah, M.A.; Hussain, G.; Riaz, A.; Sarfraz, I.; Zafar, S.; Adnan, M.; Khan, A.H.; Selamoglu, Z. Cardamonin: A new player to fight cancer via multiple cancer signaling pathways. Life Sci. 2020, 250, 117591. [Google Scholar] [CrossRef] [PubMed]
  369. Gonçalves, L.M.; Valente, I.M.; Rodrigues, J.A. An overview on cardamonin. J. Med. Food 2014, 17, 633–640. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  370. Ohtsuki, T.; Kikuchi, H.; Koyano, T.; Kowithayakorn, T.; Sakai, T.; Ishibashi, M. Death receptor 5 promoter-enhancing compounds isolated from Catimbium speciosum and their enhancement effect on TRAIL-induced apoptosis. Bioorg. Med. Chem. 2009, 17, 6748–6754. [Google Scholar] [CrossRef]
  371. Simirgiotis, M.J.; Adachi, S.; To, S.; Yang, H.; Reynertson, K.A.; Basile, M.J.; Gil, R.R.; Weinstein, I.B.; Kennelly, E.J. Cytotoxic chalcones and antioxidants from the fruits of a Syzygium samarangense (Wax Jambu). Food Chem. 2008, 107, 813–819. [Google Scholar] [CrossRef] [Green Version]
  372. Bajgai, S.P.; Prachyawarakorn, V.; Mahidol, C.; Ruchirawat, S.; Kittakoop, P. Hybrid flavan-chalcones, aromatase and lipoxygenase inhibitors, from Desmos cochinchinensis. Phytochemistry 2011, 72, 2062–2067. [Google Scholar] [CrossRef]
  373. Lee, M.Y.; Seo, C.S.; Lee, J.A.; Shin, I.S.; Kim, S.J.; Ha, H.; Shin, H.K. Alpinia katsumadai H(AYATA) seed extract inhibit LPS-induced inflammation by induction of heme oxygenase-1 in RAW264.7 cells. Inflammation 2012, 35, 746–757. [Google Scholar] [CrossRef]
  374. Lee, J.H.; Jung, H.S.; Giang, P.M.; Jin, X.; Lee, S.; Son, P.T.; Lee, D.; Hong, Y.S.; Lee, K.; Lee, J.J. Blockade of nuclear factor-kappaB signaling pathway and anti-inflammatory activity of cardamomin, a chalcone analog from Alpinia conchigera. J. Pharmacol. Exp. Ther. 2006, 316, 271–278. [Google Scholar] [CrossRef]
  375. Qin, Y.; Sun, C.Y.; Lu, F.R.; Shu, X.R.; Yang, D.; Chen, L.; She, X.M.; Gregg, N.M.; Guo, T.; Hu, Y. Cardamonin exerts potent activity against multiple myeloma through blockade of NF-κB pathway in vitro. Leuk. Res. 2012, 36, 514–520. [Google Scholar] [CrossRef]
  376. Yamamoto, N.; Kawabata, K.; Sawada, K.; Ueda, M.; Fukuda, I.; Kawasaki, K.; Murakami, A.; Ashida, H. Cardamonin stimulates glucose uptake through translocation of glucose transporter-4 in L6 myotubes. Phytother. Res. 2011, 25, 1218–1224. [Google Scholar] [CrossRef]
  377. Liao, Q.; Shi, D.H.; Zheng, W.; Xu, X.J.; Yu, Y.H. Antiproliferation of cardamonin is involved in mTOR on aortic smooth muscle cells in high fructose-induced insulin resistance rats. Eur. J. Pharmacol. 2010, 641, 179–186. [Google Scholar] [CrossRef]
  378. Hou, G.; Yuan, X.; Li, Y.; Hou, G.; Liu, X. Cardamonin, a natural chalcone, reduces 5-fluorouracil resistance of gastric cancer cells through targeting Wnt/β-catenin signal pathway. Investig. New Drugs 2020, 38, 329–339. [Google Scholar] [CrossRef]
  379. Wang, Z.; Tang, X.; Wu, X.; Yang, M.; Wang, W.; Wang, L.; Tang, D.; Wang, D. Cardamonin exerts anti-gastric cancer activity via inhibiting LncRNA-PVT1-STAT3 axis. Biosci. Rep. 2019, 39. [Google Scholar] [CrossRef] [Green Version]
  380. Li, Y.; Qin, Y.; Yang, C.; Zhang, H.; Li, Y.; Wu, B.; Huang, J.; Zhou, X.; Huang, B.; Yang, K.; et al. Retraction Note: Cardamonin induces ROS-mediated G2/M phase arrest and apoptosis through inhibition of NF-κB pathway in nasopharyngeal carcinoma. Cell Death Dis. 2019, 10, 289. [Google Scholar] [CrossRef] [Green Version]
  381. Chen, H.; Shi, D.; Niu, P.; Zhu, Y.; Zhou, J. Anti-inflammatory effects of cardamonin in ovarian cancer cells are mediated via mTOR suppression. Planta Med. 2018, 84, 1183–1190. [Google Scholar] [CrossRef]
  382. Shrivastava, S.; Jeengar, M.K.; Thummuri, D.; Koval, A.; Katanaev, V.L.; Marepally, S.; Naidu, V.G.M. Cardamonin, a chalcone, inhibits human triple negative breast cancer cell invasiveness by downregulation of Wnt/β-catenin signaling cascades and reversal of epithelial-mesenchymal transition. Biofactors 2017, 43, 152–169. [Google Scholar] [CrossRef] [Green Version]
  383. Zhou, X.; Zhou, R.; Li, Q.; Jie, X.; Hong, J.; Zong, Y.; Dong, X.; Zhang, S.; Li, Z.; Wu, G. Cardamonin inhibits the proliferation and metastasis of non-small-cell lung cancer cells by suppressing the PI3K/Akt/mTOR pathway. Anticancer Drugs 2019, 30, 241–250. [Google Scholar] [CrossRef]
  384. Park, S.; Gwak, J.; Han, S.J.; Oh, S. Cardamonin suppresses the proliferation of colon cancer cells by promoting β-catenin degradation. Biol. Pharm. Bull. 2013, 36, 1040–1044. [Google Scholar] [CrossRef] [Green Version]
  385. Zhang, J.; Sikka, S.; Siveen, K.S.; Lee, J.H.; Um, J.Y.; Kumar, A.P.; Chinnathambi, A.; Alharbi, S.A.; Rangappa, K.S.; Sethi, G.; et al. Cardamonin represses proliferation, invasion, and causes apoptosis through the modulation of signal transducer and activator of transcription 3 pathway in prostate cancer. Apoptosis 2017, 22, 158–168. [Google Scholar] [CrossRef]
  386. James, S.; Aparna, J.S.; Paul, A.M.; Lankadasari, M.B.; Mohammed, S.; Binu, V.S.; Santhoshkumar, T.R.; Reshmi, G.; Harikumar, K.B. Cardamonin inhibits colonic neoplasia through modulation of MicroRNA expression. Sci. Rep. 2017, 7, 13945. [Google Scholar] [CrossRef]
  387. Kim, Y.J.; Kang, K.S.; Choi, K.C.; Ko, H. Cardamonin induces autophagy and an antiproliferative effect through JNK activation in human colorectal carcinoma HCT116 cells. Bioorg. Med. Chem. Lett. 2015, 25, 2559–2564. [Google Scholar] [CrossRef] [PubMed]
  388. Liao, N.C.; Shih, Y.L.; Chou, J.S.; Chen, K.W.; Chen, Y.L.; Lee, M.H.; Peng, S.F.; Leu, S.J.; Chung, J.G. Cardamonin induces cell cycle arrest, apoptosis and alters apoptosis associated gene expression in WEHI-3 mouse leukemia cells. Am. J. Chin. Med. 2019, 47, 635–656. [Google Scholar] [CrossRef] [PubMed]
  389. Bhalla, V.X.; Nayak, U.R.; Dev, S. Some new flavonoids from Psoralea Corylifolia. Tetrahedron Lett. 1968, 20, 2401–2406. [Google Scholar] [CrossRef]
  390. Yan, C.; Wu, Y.; Weng, Z.; Gao, Q.; Yang, G.; Chen, Z.; Cai, B.; Li, W. Development of an HPLC method for absolute quantification and QAMS of flavonoids components in Psoralea corylifolia L. J. Anal. Methods Chem. 2015, 2015, 792637. [Google Scholar] [CrossRef] [Green Version]
  391. ElSohly, H.N.; Joshi, A.S.; Nimrod, A.C.; Walker, L.A.; Clark, A.M. Antifungal chalcones from Maclura tinctoria. Planta Med. 2001, 67, 87–89. [Google Scholar] [CrossRef]
  392. Jantan, I.; Mohd Yasin, Y.H.; Jamil, S.; Sirat, H.; Basar, N. Effect of prenylated flavonoids and chalcones isolated from Artocarpus species on platelet aggregation in human whole blood. J. Nat. Med. 2010, 64, 365–369. [Google Scholar] [CrossRef]
  393. Tsai, W.J.; Hsin, W.C.; Chen, C.C. Antiplatelet flavonoids from seeds of Psoralea corylifolia. J. Nat. Prod. 1996, 59, 671–672. [Google Scholar] [CrossRef]
  394. Jing, H.; Zhou, X.; Dong, X.; Cao, J.; Zhu, H.; Lou, J.; Hu, Y.; He, Q.; Yang, B. Abrogation of Akt signaling by Isobavachalcone contributes to its anti-proliferative effects towards human cancer cells. Cancer Lett. 2010, 294, 167–177. [Google Scholar] [CrossRef]
  395. Szliszka, E.; Czuba, Z.P.; Mazur, B.; Sedek, L.; Paradysz, A.; Krol, W. Chalcones enhance TRAIL-induced apoptosis in prostate cancer cells. Int. J. Mol. Sci. 2009, 11, 1–13. [Google Scholar] [CrossRef] [Green Version]
  396. Ohno, O.; Watabe, T.; Nakamura, K.; Kawagoshi, M.; Uotsu, N.; Chiba, T.; Yamada, M.; Yamaguchi, K.; Yamada, K.; Miyamoto, K.; et al. Inhibitory effects of bakuchiol, bavachin, and isobavachalcone isolated from Piper longum on melanin production in B16 mouse melanoma cells. Biosci. Biotechnol. Biochem. 2010, 74, 1504–1506. [Google Scholar] [CrossRef] [Green Version]
  397. Zhao, S.; Ma, C.M.; Liu, C.X.; Wei, W.; Sun, Y.; Yan, H.; Wu, Y.L. Autophagy inhibition enhances isobavachalcone-induced cell death in multiple myeloma cells. Int. J. Mol. Med. 2012, 30, 939–944. [Google Scholar] [CrossRef] [Green Version]
  398. Nishimura, R.; Tabata, K.; Arakawa, M.; Ito, Y.; Kimura, Y.; Akihisa, T.; Nagai, H.; Sakuma, A.; Kohno, H.; Suzuki, T. Isobavachalcone, a chalcone constituent of Angelica keiskei, induces apoptosis in neuroblastoma. Biol. Pharm. Bull. 2007, 30, 1878–1883. [Google Scholar] [CrossRef] [Green Version]
  399. Wang, K.L.; Yu, Y.C.; Hsia, S.M. Perspectives on the role of isoliquiritigenin in cancer. Cancers 2021, 13, 115. [Google Scholar] [CrossRef]
  400. Wang, K.L.; Hsia, S.M.; Chan, C.J.; Chang, F.Y.; Huang, C.Y.; Bau, D.T.; Wang, P.S. Inhibitory effects of isoliquiritigenin on the migration and invasion of human breast cancer cells. Expert Opin. Ther. Targets 2013, 17, 337–349. [Google Scholar] [CrossRef]
  401. Jin, H.; Lee, S.H.; Lee, S.H. Isoliquiritigenin-mediated p62/SQSTM1 induction regulates apoptotic potential through attenuation of caspase-8 activation in colorectal cancer cells. Eur. J. Pharmacol. 2018, 841, 90–97. [Google Scholar] [CrossRef]
  402. Chen, H.Y.; Huang, T.C.; Shieh, T.M.; Wu, C.H.; Lin, L.C.; Hsia, S.M. Isoliquiritigenin induces autophagy and inhibits ovarian cancer cell growth. Int. J. Mol. Sci. 2017, 18, 2025. [Google Scholar] [CrossRef] [Green Version]
  403. Cao, Z.X.; Wen, Y.; He, J.L.; Huang, S.Z.; Gao, F.; Guo, C.J.; Liu, Q.Q.; Zheng, S.W.; Gong, D.Y.; Li, Y.Z.; et al. Isoliquiritigenin, an orally available natural FLT3 inhibitor from licorice, exhibits selective anti-acute myeloid leukemia efficacy in vitro and in vivo. Mol. Pharmacol. 2019, 96, 589–599. [Google Scholar] [CrossRef]
  404. Xiang, S.; Chen, H.; Luo, X.; An, B.; Wu, W.; Cao, S.; Ruan, S.; Wang, Z.; Weng, L.; Zhu, H.; et al. Isoliquiritigenin suppresses human melanoma growth by targeting miR-301b/LRIG1 signaling. J. Exp. Clin. Cancer Res. 2018, 37, 184. [Google Scholar] [CrossRef]
  405. Chen, X.; Wu, Y.; Jiang, Y.; Zhou, Y.; Wang, Y.; Yao, Y.; Yi, C.; Gou, L.; Yang, J. Isoliquiritigenin inhibits the growth of multiple myeloma via blocking IL-6 signaling. J. Mol. Med. 2012, 90, 1311–1319. [Google Scholar] [CrossRef]
  406. Botta, B.; Vitali, A.; Menendez, P.; Misiti, D.; Delle Monache, G. Prenylated flavonoids: Pharmacology and biotechnology. Curr. Med. Chem. 2005, 12, 717–739. [Google Scholar] [CrossRef]
  407. Stevens, J.F.; Page, J.E. Xanthohumol and related prenylflavonoids from hops and beer: To your good health! Phytochemistry 2004, 65, 1317–1330. [Google Scholar] [CrossRef]
  408. Jiang, C.H.; Sun, T.L.; Xiang, D.X.; Wei, S.S.; Li, W.Q. Anticancer activity and mechanism of xanthohumol: A prenylated flavonoid from hops (Humulus lupulus L.). Front. Pharmacol. 2018, 9, 530. [Google Scholar] [CrossRef]
  409. Roehrer, S.; Stork, V.; Ludwig, C.; Minceva, M.; Behr, J. Analyzing bioactive effects of the minor hop compound xanthohumol C on human breast cancer cells using quantitative proteomics. PLoS ONE 2019, 14, e0213469. [Google Scholar] [CrossRef]
  410. Logan, I.E.; Miranda, C.L.; Lowry, M.B.; Maier, C.S.; Stevens, J.F.; Gombart, A.F. Antiproliferative and cytotoxic activity of xanthohumol and its non-estrogenic derivatives in colon and hepatocellular carcinoma cell lines. Int. J. Mol. Sci. 2019, 20, 1203. [Google Scholar] [CrossRef] [Green Version]
  411. Yong, W.K.; Malek, S.N.A. Xanthohumol induces growth inhibition and apoptosis in ca ski human cervical cancer cells. Evid. Based Complement. Alternat. Med. 2015, 2015, 921306. [Google Scholar] [CrossRef]
  412. Sławińska-Brych, A.; Zdzisińska, B.; Dmoszyńska-Graniczka, M.; Jeleniewicz, W.; Kurzepa, J.; Gagoś, M.; Stepulak, A. Xanthohumol inhibits the extracellular signal regulated kinase (ERK) signalling pathway and suppresses cell growth of lung adenocarcinoma cells. Toxicology 2016, 357–358, 65–73. [Google Scholar] [CrossRef] [PubMed]
  413. Mi, X.; Wang, C.; Sun, C.; Chen, X.; Huo, X.; Zhang, Y.; Li, G.; Xu, B.; Zhang, J.; Xie, J.; et al. Xanthohumol induces paraptosis of leukemia cells through p38 mitogen activated protein kinase signaling pathway. Oncotarget 2017, 8, 31297–31304. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  414. Sławińska-Brych, A.; Zdzisińska, B.; Czerwonka, A.; Mizerska-Kowalska, M.; Dmoszyńska-Graniczka, M.; Stepulak, A.; Gagoś, M. Xanthohumol exhibits anti-myeloma activity in vitro through inhibition of cell proliferation, induction of apoptosis via the ERK and JNK-dependent mechanism, and suppression of sIL-6R and VEGF production. Biochim. Biophys. Acta Gen. Subj. 2019, 1863, 129408. [Google Scholar] [CrossRef] [PubMed]
  415. Choi, J.H.; Rho, M.C.; Lee, S.W.; Choi, J.N.; Kim, K.; Song, G.Y.; Kim, Y.K. Bavachin and isobavachalcone, acyl-coenzyme A: Cholesterol acyltransferase inhibitors from Psoralea corylifolia. Arch. Pharm. Res. 2008, 31, 1419–1423. [Google Scholar] [CrossRef] [PubMed]
  416. Hung, Y.L.; Wang, S.C.; Suzuki, K.; Fang, S.H.; Chen, C.S.; Cheng, W.C.; Su, C.C.; Yeh, H.C.; Tu, H.P.; Liu, P.L.; et al. Bavachin attenuates LPS-induced inflammatory response and inhibits the activation of NLRP3 inflammasome in macrophages. Phytomedicine 2019, 59, 152785. [Google Scholar] [CrossRef]
  417. Takeda, T.; Tsubaki, M.; Tomonari, Y.; Kawashima, K.; Itoh, T.; Imano, M.; Satou, T.; Nishida, S. Bavachin induces the apoptosis of multiple myeloma cell lines by inhibiting the activation of nuclear factor kappa B and signal transducer and activator of transcription 3. Biomed. Pharmacother. 2018, 100, 486–494. [Google Scholar] [CrossRef]
  418. Lee, H.; Li, H.; Noh, M.; Ryu, J.H. Bavachin from Psoralea corylifolia improves insulin-dependent glucose uptake through insulin signaling and AMPK activation in 3T3-L1 adipocytes. Int. J. Mol. Sci. 2016, 17, 527. [Google Scholar] [CrossRef]
  419. Tao, Z.; Liu, J.; Jiang, Y.; Gong, L.; Yang, B. Synthesis of prenylated flavonols and their potents as estrogen receptor modulator. Sci. Rep. 2017, 7, 12445. [Google Scholar] [CrossRef]
  420. Wang, Y.; Dong, H.; Zhu, M.; Ou, Y.; Zhang, J.; Luo, H.; Luo, R.; Wu, J.; Mao, M.; Liu, X.; et al. Icariin exterts negative effects on human gastric cancer cell invasion and migration by vasodilator-stimulated phosphoprotein via Rac1 pathway. Eur. J. Pharmacol. 2010, 635, 40–48. [Google Scholar] [CrossRef]
  421. Li, W.; Wang, M.; Wang, L.; Ji, S.; Zhang, J.; Zhang, C. Icariin synergizes with arsenic trioxide to suppress human hepatocellular carcinoma. Cell Biochem. Biophys. 2014, 68, 427–436. [Google Scholar] [CrossRef]
  422. Li, S.; Dong, P.; Wang, J.; Zhang, J.; Gu, J.; Wu, X.; Wu, W.; Fei, X.; Zhang, Z.; Wang, Y.; et al. Icariin, a natural flavonol glycoside, induces apoptosis in human hepatoma SMMC-7721 cells via a ROS/JNK-dependent mitochondrial pathway. Cancer Lett. 2010, 298, 222–230. [Google Scholar] [CrossRef]
  423. Zhang, D.C.; Liu, J.L.; Ding, Y.B.; Xia, J.G.; Chen, G.Y. Icariin potentiates the antitumor activity of gemcitabine in gallbladder cancer by suppressing NF-κB. Acta Pharmacol. Sin. 2013, 34, 301–308. [Google Scholar] [CrossRef]
  424. Ma, H.R.; Wang, J.; Chen, Y.F.; Chen, H.; Wang, W.S.; Aisa, H.A. Icariin and icaritin stimulate the proliferation of SKBr3 cells through the GPER1-mediated modulation of the EGFR-MAPK signaling pathway. Int. J. Mol. Med. 2014, 33, 1627–1634. [Google Scholar] [CrossRef] [Green Version]
  425. Li, J.; Jiang, K.; Zhao, F. Icariin regulates the proliferation and apoptosis of human ovarian cancer cells through microRNA-21 by targeting PTEN, RECK and Bcl-2. Oncol. Rep. 2015, 33, 2829–2836. [Google Scholar] [CrossRef] [Green Version]
  426. Gu, Z.F.; Zhang, Z.T.; Wang, J.Y.; Xu, B.B. Icariin exerts inhibitory effects on the growth and metastasis of KYSE70 human esophageal carcinoma cells via PI3K/AKT and STAT3 pathways. Environ. Toxicol. Pharmacol. 2017, 54, 7–13. [Google Scholar] [CrossRef]
  427. Fan, C.; Yang, Y.; Liu, Y.; Jiang, S.; Di, S.; Hu, W.; Ma, Z.; Li, T.; Zhu, Y.; Xin, Z.; et al. Icariin displays anticancer activity against human esophageal cancer cells via regulating endoplasmic reticulum stress-mediated apoptotic signaling. Sci. Rep. 2016, 6, 21145. [Google Scholar] [CrossRef] [Green Version]
  428. Jung, Y.Y.; Lee, J.H.; Nam, D.; Narula, A.S.; Namjoshi, O.A.; Blough, B.E.; Um, J.Y.; Sethi, G.; Ahn, K.S. Anti-myeloma effects of icariin are mediated through the attenuation of JAK/STAT3-dependent signaling cascade. Front. Pharmacol. 2018, 9, 531. [Google Scholar] [CrossRef]
  429. Yang, X.J.; Xi, Y.M.; Li, Z.J. Icaritin: A novel natural candidate for hematological malignancies therapy. Biomed. Res. Int. 2019, 2019, 4860268. [Google Scholar] [CrossRef]
  430. Yin, L.; Qi, X.W.; Liu, X.Z.; Yang, Z.Y.; Cai, R.L.; Cui, H.J.; Chen, L.; Yu, S.C. Icaritin enhances the efficacy of cetuximab against triple-negative breast cancer cells. Oncol. Lett. 2020, 19, 3950–3958. [Google Scholar] [CrossRef]
  431. Wang, X.; Zheng, N.; Dong, J.; Wang, X.; Liu, L.; Huang, J. Estrogen receptor-α36 is involved in icaritin induced growth inhibition of triple-negative breast cancer cells. J. Steroid Biochem. Mol. Biol. 2017, 171, 318–327. [Google Scholar] [CrossRef]
  432. Han, H.; Xu, B.; Hou, P.; Jiang, C.; Liu, L.; Tang, M.; Yang, X.; Zhang, Y.; Liu, Y. Icaritin sensitizes human glioblastoma cells to TRAIL-induced apoptosis. Cell Biochem. Biophys. 2015, 72, 533–542. [Google Scholar] [CrossRef]
  433. Tiong, C.T.; Chen, C.; Zhang, S.J.; Li, J.; Soshilov, A.; Denison, M.S.; Lee, L.S.; Tam, V.H.; Wong, S.P.; Xu, H.E.; et al. A novel prenylflavone restricts breast cancer cell growth through AhR-mediated destabilization of ERα protein. Carcinogenesis 2012, 33, 1089–1097. [Google Scholar] [CrossRef] [Green Version]
  434. Guo, Y.; Zhang, X.; Meng, J.; Wang, Z.Y. An anticancer agent icaritin induces sustained activation of the extracellular signal-regulated kinase (ERK) pathway and inhibits growth of breast cancer cells. Eur. J. Pharmacol. 2011, 658, 114–122. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  435. Gupta, P.B.; Onder, T.T.; Jiang, G.; Tao, K.; Kuperwasser, C.; Weinberg, R.A.; Lander, E.S. Identification of selective inhibitors of cancer stem cells by high-throughput screening. Cell 2009, 138, 645–659. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  436. Lu, P.H.; Chen, M.B.; Liu, Y.Y.; Wu, M.H.; Li, W.T.; Wei, M.X.; Liu, C.Y.; Qin, S.K. Identification of sphingosine kinase 1 (SphK1) as a primary target of icaritin in hepatocellular carcinoma cells. Oncotarget 2017, 8, 22800–22810. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  437. Hu, J.; Zhu, W.; Wei, B.; Wen, H.; Mao, S.; Xu, H.; Hu, M.; Yang, T.; Jiang, H. Antitumoral action of icaritin in LNCaP prostate cancer cells by regulating PEA3/HER2/AR signaling. Anticancer Drugs 2016, 27, 944–952. [Google Scholar] [CrossRef]
  438. Sun, L.; Peng, Q.; Qu, L.; Gong, L.; Si, J. Anticancer agent icaritin induces apoptosis through caspase-dependent pathways in human hepatocellular carcinoma cells. Mol. Med. Rep. 2015, 11, 3094–3100. [Google Scholar] [CrossRef] [Green Version]
  439. Zheng, Q.; Liu, W.W.; Li, B.; Chen, H.J.; Zhu, W.S.; Yang, G.X.; Chen, M.J.; He, G.Y. Anticancer effect of icaritin on human lung cancer cells through inducing S phase cell cycle arrest and apoptosis. J. Huazhong Univ. Sci. Technol. Med. Sci. 2014, 34, 497–503. [Google Scholar] [CrossRef]
  440. Jin, L.; Miao, J.; Liu, Y.; Li, X.; Jie, Y.; Niu, Q.; Han, X. Icaritin induces mitochondrial apoptosis by up-regulating miR-124 in human oral squamous cell carcinoma cells. Biomed. Pharmacother. 2017, 85, 287–295. [Google Scholar] [CrossRef]
  441. Yang, J.G.; Lu, R.; Ye, X.J.; Zhang, J.; Tan, Y.Q.; Zhou, G. Icaritin reduces oral squamous cell carcinoma progression via the inhibition of STAT3 signaling. Int. J. Mol. Sci. 2017, 18, 132. [Google Scholar] [CrossRef] [Green Version]
  442. Han, S.; Gou, Y.; Jin, D.; Ma, J.; Chen, M.; Dong, X. Effects of Icaritin on the physiological activities of esophageal cancer stem cells. Biochem. Biophys. Res. Commun. 2018, 504, 792–796. [Google Scholar] [CrossRef]
  443. Gao, L.; Chen, M.; Ouyang, Y.; Li, R.; Zhang, X.; Gao, X.; Lin, S.; Wang, X. Icaritin induces ovarian cancer cell apoptosis through activation of p53 and inhibition of Akt/mTOR pathway. Life Sci. 2018, 202, 188–194. [Google Scholar] [CrossRef]
  444. Le, A.; Wang, Z.H.; Dai, X.Y.; Xiao, T.H.; Zhuo, R.; Zhang, B.; Xiao, Z.; Fan, X. Icaritin inhibits decidualization of endometrial stromal cells. Exp. Ther. Med. 2017, 14, 5949–5955. [Google Scholar] [CrossRef]
  445. Tong, J.S.; Zhang, Q.H.; Huang, X.; Fu, X.Q.; Qi, S.T.; Wang, Y.P.; Hou, Y.; Sheng, J.; Sun, Q.Y. Icaritin causes sustained ERK1/2 activation and induces apoptosis in human endometrial cancer cells. PLoS ONE 2011, 6, e16781. [Google Scholar] [CrossRef] [Green Version]
  446. Zhou, C.; Gu, J.; Zhang, G.; Dong, D.; Yang, Q.; Chen, M.B.; Xu, D. AMPK-autophagy inhibition sensitizes icaritin-induced anti-colorectal cancer cell activity. Oncotarget 2017, 8, 14736–14747. [Google Scholar] [CrossRef] [Green Version]
  447. Wang, X.; Liu, Y.; Liu, X.; Yang, J.; Teng, G.; Zhang, L.; Zhou, C. MiR-124 inhibits cell proliferation, migration and invasion by directly targeting SOX9 in lung adenocarcinoma. Oncol. Rep. 2016, 35, 3115–3121. [Google Scholar] [CrossRef] [Green Version]
  448. Liu, Y.; Shi, L.; Liu, Y.; Li, P.; Jiang, G.; Gao, X.; Zhang, Y.; Jiang, C.; Zhu, W.; Han, H.; et al. Activation of PPARγ mediates icaritin-induced cell cycle arrest and apoptosis in glioblastoma multiforme. Biomed. Pharmacother. 2018, 100, 358–366. [Google Scholar] [CrossRef]
  449. Wang, Z.D.; Wang, R.Z.; Xia, Y.Z.; Kong, L.Y.; Yang, L. Reversal of multidrug resistance by icaritin in doxorubicin-resistant human osteosarcoma cells. Chin. J. Nat. Med. 2018, 16, 20–28. [Google Scholar] [CrossRef]
  450. Xu, B.; Jiang, C.; Han, H.; Liu, H.; Tang, M.; Liu, L.; Ji, W.; Lu, X.; Yang, X.; Zhang, Y.; et al. Icaritin inhibits the invasion and epithelial-to-mesenchymal transition of glioblastoma cells by targeting EMMPRIN via PTEN/AKt/HIF-1α signalling. Clin. Exp. Pharmacol. Physiol. 2015, 42, 1296–1307. [Google Scholar] [CrossRef]
  451. Zhu, J.F.; Li, Z.J.; Zhang, G.S.; Meng, K.; Kuang, W.Y.; Li, J.; Zhou, X.F.; Li, R.J.; Peng, H.L.; Dai, C.W.; et al. Icaritin shows potent anti-leukemia activity on chronic myeloid leukemia in vitro and in vivo by regulating MAPK/ERK/JNK and JAK2/STAT3/AKT signalings. PLoS ONE 2011, 6, e23720. [Google Scholar] [CrossRef]
  452. Zhu, S.; Wang, Z.; Li, Z.; Peng, H.; Luo, Y.; Deng, M.; Li, R.; Dai, C.; Xu, Y.; Liu, S.; et al. Icaritin suppresses multiple myeloma, by inhibiting IL-6/JAK2/STAT3. Oncotarget 2015, 6, 10460–10472. [Google Scholar] [CrossRef] [Green Version]
  453. Wu, T.; Wang, S.; Wu, J.; Lin, Z.; Sui, X.; Xu, X.; Shimizu, N.; Chen, B.; Wang, X. Icaritin induces lytic cytotoxicity in extranodal NK/T-cell lymphoma. J. Exp. Clin. Cancer Res. 2015, 34, 17. [Google Scholar] [CrossRef] [Green Version]
  454. Li, Z.J.; Yao, C.; Liu, S.F.; Chen, L.; Xi, Y.M.; Zhang, W.; Zhang, G.S. Cytotoxic effect of icaritin and its mechanisms in inducing apoptosis in human burkitt lymphoma cell line. Biomed. Res. Int. 2014, 2014, 391512. [Google Scholar] [CrossRef]
  455. Li, Q.; Huai, L.; Zhang, C.; Wang, C.; Jia, Y.; Chen, Y.; Yu, P.; Wang, H.; Rao, Q.; Wang, M.; et al. Icaritin induces AML cell apoptosis via the MAPK/ERK and PI3K/AKT signal pathways. Int. J. Hematol. 2013, 97, 617–623. [Google Scholar] [CrossRef]
  456. Li, Z.Y.; Li, Z.J.; Chen, X.; Huang, X.R.; Fang, Z.Q.; Zhang, J.G.; Fang, J. Icaritin reverses multidrug resistance of multiple myeloma cell line KM3/BTZ. Zhongguo Shi Yan Xue Ye Xue Za Zhi 2017, 25, 1690–1695. (In Chinese) [Google Scholar] [CrossRef]
  457. Xia, Q.; Xu, D.; Huang, Z.; Liu, J.; Wang, X.; Wang, X.; Liu, S. Preparation of icariside II from icariin by enzymatic hydrolysis method. Fitoterapia 2010, 81, 437–442. [Google Scholar] [CrossRef] [PubMed]
  458. Lee, K.S.; Lee, H.J.; Ahn, K.S.; Kim, S.H.; Nam, D.; Kim, D.K.; Choi, D.Y.; Ahn, K.S.; Lu, J.; Kim, S.H. Cyclooxygenase-2/prostaglandin E2 pathway mediates icariside II induced apoptosis in human PC-3 prostate cancer cells. Cancer Lett. 2009, 280, 93–100. [Google Scholar] [CrossRef] [PubMed]
  459. Choi, H.J.; Eun, J.S.; Kim, D.K.; Li, R.H.; Shin, T.Y.; Park, H.; Cho, N.P.; Soh, Y. Icariside II from Epimedium koreanum inhibits hypoxia-inducible factor-1alpha in human osteosarcoma cells. Eur. J. Pharmacol. 2008, 579, 58–65. [Google Scholar] [CrossRef] [PubMed]
  460. Kim, S.H.; Ahn, K.S.; Jeong, S.J.; Kwon, T.R.; Jung, J.H.; Yun, S.M.; Han, I.; Lee, S.G.; Kim, D.K.; Kang, M.; et al. Janus activated kinase 2/signal transducer and activator of transcription 3 pathway mediates icariside II-induced apoptosis in U266 multiple myeloma cells. Eur. J. Pharmacol. 2011, 654, 10–16. [Google Scholar] [CrossRef]
  461. Lamorte, D.; Faraone, I.; Laurenzana, I.; Milella, L.; Trino, S.; De Luca, L.; Del Vecchio, L.; Armentano, M.F.; Sinisgalli, C.; Chiummiento, L.; et al. Future in the past: Azorella glabra Wedd. as a source of new natural compounds with antiproliferative and cytotoxic activity on multiple myeloma cells. Int. J. Mol. Sci. 2018, 19, 3348. [Google Scholar] [CrossRef] [Green Version]
  462. Kiraz, Y.; Neergheen-Bhujun, V.S.; Rummun, N.; Baran, Y. Apoptotic effects of non-edible parts of Punica granatum on human multiple myeloma cells. Tumour Biol. 2016, 37, 1803–1815. [Google Scholar] [CrossRef] [Green Version]
  463. Kumar, G.; Mittal, S.; Sak, K.; Tuli, H.S. Molecular mechanisms underlying chemopreventive potential of curcumin: Current challenges and future perspectives. Life Sci. 2016, 148, 313–328. [Google Scholar] [CrossRef]
  464. Cho, J.; Rho, O.; Junco, J.; Carbajal, S.; Siegel, D.; Slaga, T.J.; DiGiovanni, J. Effect of combined treatment with ursolic acid and resveratrol on skin tumor promotion by 12-O-tetradecanoylphorbol-13-acetate. Cancer Prev. Res. 2015, 8, 817–825. [Google Scholar] [CrossRef] [Green Version]
  465. Chen, J.; Li, L.; Su, J.; Li, B.; Chen, T.; Wong, Y.S. Synergistic apoptosis-inducing effects on A375 human melanoma cells of natural borneol and curcumin. PLoS ONE 2014, 9, e101277. [Google Scholar] [CrossRef] [Green Version]
  466. Cheng, Y.T.; Yang, C.C.; Shyur, L.F. Phytomedicine-modulating oxidative stress and the tumor microenvironment for cancer therapy. Pharmacol. Res. 2016, 114, 128–143. [Google Scholar] [CrossRef]
  467. Efferth, T. Cancer combination therapies with artemisinin-type drugs. Biochem. Pharmacol. 2017, 139, 56–70. [Google Scholar] [CrossRef]
  468. Siveen, K.S.; Uddin, S.; Mohammad, R.M. Targeting acute myeloid leukemia stem cell signaling by natural products. Mol. Cancer 2017, 16, 13. [Google Scholar] [CrossRef] [Green Version]
  469. Cragg, G.M.; Newman, D.J. Plants as a source of anti-cancer agents. J. Ethnopharmacol. 2005, 100, 72–79. [Google Scholar] [CrossRef] [Green Version]
  470. Frost, B.M.; Lönnerholm, G.; Koopmans, P.; Abrahamsson, J.; Behrendtz, M.; Castor, A.; Forestier, E.; Uges, D.R.A.; De Graaf, S.S.N. Vincristine in childhood leukaemia: No pharmacokinetic rationale for dose reduction in adolescents. Acta Paediatr. 2003, 92, 551–557. [Google Scholar] [CrossRef]
  471. Okouneva, T.; Hill, B.T.; Wilson, L.; Jordan, M.A. The effects of vinflunine, vinorelbine, and vinblastine on centromere dynamics. Mol. Cancer Ther. 2003, 2, 427–436. [Google Scholar]
  472. Mann, J. Natural products in cancer chemotherapy: Past, present and future. Nat. Rev. Cancer. 2002, 2, 143–148. [Google Scholar] [CrossRef]
  473. İşeri, Ö.D.; Yurtcu, E.; Sahin, F.I.; Haberal, M. Corchorus olitorius (jute) extract induced cytotoxicity and genotoxicity on human multiple myeloma cells (ARH-77). Pharm. Biol. 2013, 51, 766–770. [Google Scholar] [CrossRef]
  474. Gao, R.; Miao, X.; Sun, C.; Su, S.; Zhu, Y.; Qian, D.; Ouyang, Z.; Duan, J. Frankincense and myrrh and their bioactive compounds ameliorate the multiple myeloma through regulation of metabolome profiling and JAK/STAT signaling pathway based on U266 cells. BMC Complement. Med. Ther. 2020, 20, 96. [Google Scholar] [CrossRef]
  475. Adham, A.N.; Naqishbandi, A.M.; Efferth, T. Cytotoxicity and apoptosis induction by Fumaria officinalis extracts in leukemia and multiple myeloma cell lines. J. Ethnopharmacol. 2021, 266, 113458. [Google Scholar] [CrossRef]
  476. Malacrida, A.; Maggioni, D.; Cassetti, A.; Nicolini, G.; Cavaletti, G.; Miloso, M. Antitumoral effect of Hibiscus sabdariffa on human squamous cell carcinoma and multiple myeloma cells. Nutr. Cancer 2016, 68, 1161–1170. [Google Scholar] [CrossRef]
  477. Peng, J.; Chen, Y.; Lin, J.; Zhuang, Q.; Xu, W.; Hong, Z.; Sferra, T.J. Patrinia scabiosaefolia extract suppresses proliferation and promotes apoptosis by inhibiting the STAT3 pathway in human multiple myeloma cells. Mol. Med. Rep. 2011, 4, 313–318. [Google Scholar] [CrossRef] [PubMed]
  478. Kim, C.; Song, H.S.; Park, H.; Kim, B. Activation of ER stress-dependent miR-216b has a critical role in Salviamiltiorrhiza ethanol-extract-induced apoptosis in U266 and U937 cells. Int. J. Mol. Sci. 2018, 19, 1240. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  479. Che, Y.; Hou, S.; Kang, Z.; Lin, Q. Serenoa repens induces growth arrest and apoptosis of human multiple myeloma cells via inactivation of STAT 3 signaling. Oncol. Rep. 2009, 22, 377–383. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  480. Rao, P.S.; Prasad, M.N. Strychnos nux-vomica root extract induces apoptosis in the human multiple myeloma cell line-U266B1. Cell Biochem. Biophys. 2013, 66, 443–450. [Google Scholar] [CrossRef] [PubMed]
  481. Rao, P.S.; Ramanadham, M.; Prasad, M.N. Anti-proliferative and cytotoxic effects of Strychnos nux-vomica root extract on human multiple myeloma cell line—RPMI 8226. Food Chem. Toxicol. 2009, 47, 283–288. [Google Scholar] [CrossRef]
  482. Adham, A.N.; Hegazy, M.E.F.; Naqishbandi, A.M.; Efferth, T. Induction of apoptosis, autophagy and ferroptosis by Thymus vulgaris and Arctium lappa extract in leukemia and multiple myeloma cell lines. Molecules 2020, 25, 5016. [Google Scholar] [CrossRef]
  483. Kovacs, E. Investigation of the proliferation, apoptosis/necrosis, and cell cycle phases in several human multiple myeloma cell lines. Comparison of Viscum album QuFrF extract with vincristine in an in vitro model. Sci. World J. 2010, 10, 311–320. [Google Scholar] [CrossRef]
  484. Hou, J.; Qian, J.; Li, Z.; Gong, A.; Zhong, S.; Qiao, L.; Qian, S.; Zhang, Y.; Dou, R.; Li, R.; et al. Bioactive compounds from Abelmoschus manihot L. alleviate the progression of multiple myeloma in mouse model and improve bone marrow microenvironment. OncoTargets Ther. 2020, 13, 959–973. [Google Scholar] [CrossRef] [Green Version]
Figure 1. The chemical structure of apigenin.
Figure 1. The chemical structure of apigenin.
Applsci 11 04451 g001
Figure 2. The chemical structure of baicalein.
Figure 2. The chemical structure of baicalein.
Applsci 11 04451 g002
Figure 3. The chemical structure of chrysoeriol.
Figure 3. The chemical structure of chrysoeriol.
Applsci 11 04451 g003
Figure 4. The chemical structure of luteolin.
Figure 4. The chemical structure of luteolin.
Applsci 11 04451 g004
Figure 5. The chemical structure of scutellarin.
Figure 5. The chemical structure of scutellarin.
Applsci 11 04451 g005
Figure 6. The chemical structure of wogonin.
Figure 6. The chemical structure of wogonin.
Applsci 11 04451 g006
Figure 7. The chemical structure of fisetin.
Figure 7. The chemical structure of fisetin.
Applsci 11 04451 g007
Figure 8. The chemical structure of myricetin.
Figure 8. The chemical structure of myricetin.
Applsci 11 04451 g008
Figure 9. The chemical structure of quercetin.
Figure 9. The chemical structure of quercetin.
Applsci 11 04451 g009
Figure 10. The chemical structure of epigallocatechin-3-gallate.
Figure 10. The chemical structure of epigallocatechin-3-gallate.
Applsci 11 04451 g010
Figure 11. The chemical structure of daidzin.
Figure 11. The chemical structure of daidzin.
Applsci 11 04451 g011
Figure 12. The chemical structure of formononetin.
Figure 12. The chemical structure of formononetin.
Applsci 11 04451 g012
Figure 13. The chemical structure of genistein.
Figure 13. The chemical structure of genistein.
Applsci 11 04451 g013
Figure 14. The chemical structure of butein.
Figure 14. The chemical structure of butein.
Applsci 11 04451 g014
Figure 15. The chemical structure of cardamonin.
Figure 15. The chemical structure of cardamonin.
Applsci 11 04451 g015
Figure 16. The chemical structure of isobavachalcone.
Figure 16. The chemical structure of isobavachalcone.
Applsci 11 04451 g016
Figure 17. The chemical structure of isoliquiritigenin.
Figure 17. The chemical structure of isoliquiritigenin.
Applsci 11 04451 g017
Figure 18. The chemical structure of xanthohumol.
Figure 18. The chemical structure of xanthohumol.
Applsci 11 04451 g018
Figure 19. The chemical structure of bavachin.
Figure 19. The chemical structure of bavachin.
Applsci 11 04451 g019
Figure 20. The chemical structure of icariin.
Figure 20. The chemical structure of icariin.
Applsci 11 04451 g020
Figure 21. The chemical structure of icaritin.
Figure 21. The chemical structure of icaritin.
Applsci 11 04451 g021
Figure 22. The chemical structure of icariside II.
Figure 22. The chemical structure of icariside II.
Applsci 11 04451 g022
Figure 23. Molecular functions of flavonoids against multiple myeloma (MM) from in vitro and in vivo experiments and clinical studies. ↑ increased/upregulated; ↓ decreased/downregulated; PI3K, phosphatidylinositol 3-kinase/protein kinase B; Akt, protein kinase B; mTOR, mechanistic target of rapamycin; STAT3/5, signal transducer and activator of transcription 3/5; CDK2, cyclin-dependent kinase 2; CDK4, cyclin-dependent kinase 4; NF-κB, nuclear factor-κB; p-JAK2, phosphorylated janus kinase 2; STAT3, signal transducer and activator of transcription 3; p-STAT3, phosphorylated signal transducer and activator of transcription 3; EGCG, epigallocatechin-gallate; CDKc, cyclin-dependent kinases; VEGF, vascular endothelial growth factor; PDGF, platelet-derived growth factor; bFGF, basic fibroblast growth factor; HIF-1-α, hypoxia-inducible factor 1-alpha; Bcl-2, B-cell lymphoma-2; Mcl-1L, myeloid cell leukemia-1 long; Bax, Bcl-2-associated X protein; Bad, Bcl-2-associated death promoter; MMP-2, matrix metalloproteinase 2; MMP-9, matrix metalloproteinase 9; ER stress, endoplasmic reticulum stress; MAPK, mitogen-activated kinase; ROS, reactive oxygen species; PARP, poly(ADP-ribose) polymerase; AMPK, AMP, activated protein kinase.
Figure 23. Molecular functions of flavonoids against multiple myeloma (MM) from in vitro and in vivo experiments and clinical studies. ↑ increased/upregulated; ↓ decreased/downregulated; PI3K, phosphatidylinositol 3-kinase/protein kinase B; Akt, protein kinase B; mTOR, mechanistic target of rapamycin; STAT3/5, signal transducer and activator of transcription 3/5; CDK2, cyclin-dependent kinase 2; CDK4, cyclin-dependent kinase 4; NF-κB, nuclear factor-κB; p-JAK2, phosphorylated janus kinase 2; STAT3, signal transducer and activator of transcription 3; p-STAT3, phosphorylated signal transducer and activator of transcription 3; EGCG, epigallocatechin-gallate; CDKc, cyclin-dependent kinases; VEGF, vascular endothelial growth factor; PDGF, platelet-derived growth factor; bFGF, basic fibroblast growth factor; HIF-1-α, hypoxia-inducible factor 1-alpha; Bcl-2, B-cell lymphoma-2; Mcl-1L, myeloid cell leukemia-1 long; Bax, Bcl-2-associated X protein; Bad, Bcl-2-associated death promoter; MMP-2, matrix metalloproteinase 2; MMP-9, matrix metalloproteinase 9; ER stress, endoplasmic reticulum stress; MAPK, mitogen-activated kinase; ROS, reactive oxygen species; PARP, poly(ADP-ribose) polymerase; AMPK, AMP, activated protein kinase.
Applsci 11 04451 g023
Table 1. Flavonoids involved in the management of multiple myeloma.
Table 1. Flavonoids involved in the management of multiple myeloma.
Class of FlavonoidsBioactive Compounds
FlavonesApigenin, Baicalein, Crysoeriol, Luteolin, Scutellarin, Wogonin
FlavonolsFisetin, Myricetin, Quercetin
FlavanolsEpigallocatehin-3-gallate
IsoflavononesDaidzin, Genistein
ChalconesButein, Cardamonin, Isobavachalcone, Isoliquiritigenin, Xanthohumol
Prenylated flavonoidsBavachin, Icariin, Icaritin, Icariside II
Table 2. Pharmacological effects of different plant extracts in multiple myeloma.
Table 2. Pharmacological effects of different plant extracts in multiple myeloma.
Herbal ExtractsIn Vitro/In VivoCancer Cell Line and Animal ModelBioactive EffectReferences
Azorella cglabraIn vitroSKMM1, RPMI-8226, and MM1S cells
PBMCs
↑ antioxidant activity
↓ cell viability, induced apoptosis, and cell cycle arrest
[461]
Corchorus olitorius leaf extracts (LE) and seed extracts (SE)In vitroARH-77 cells- cytotoxic effects on cells (LE and SE)
-SE had higher cytotoxicity on cells
- genotoxicity effects in dose-dependent manner (LE and SE)
[473]
Frankincense and myrrh extractsIn vitroU266 cells- inhibited cell proliferation
- ameliorated the secretion of cytokines
↓ expression of JAK/STAT signaling pathway
-3-O-acetyl-α-boswellic acid, 11-keto-boswellic acid and 3-acetyl-11 keto-boswellic acid had the most significant anti-multiple myeloma activities
[474]
Fumaria officinalis extractsIn vitroMOLP-8, NCI-H929, KMS-12BM, RPMI-8226, KMS-11, AMO-I, L-363, OPM-2, JJN-3 cells- CF induced vialbility of NCI-H929 cell line
- EF induced cytotoxicity on OPM-2 cells
- CF and EF induced apoptosis in NCI-H929 cells by loss of MMP, generation of ROS
- EF induced autophagic cell death, while CF stimulated iron-dependent cell death
[475]
Hibiscus sabdariffa extracts
In vitroRPMI-8226 cells- antiproliferative effects (Hib-ester and Hib-carbaldehyde compounds)
↓ cell migration and invasion, events involved in the process of metastasis
[6]
Hibiscus sabdariffa extract (HSE)In vitroRPMI-8226 cells↓ cell growth, motility and invasiveness by ERK1/2 modulation
↑ cytostatic effects dependent on p38 activation
[476]
Patrinia scabiosaefolia ethanol extract (EEEEPS)In vitroU266 cells- inhibited the activation of STAT3
- inhibited the proliferation
- induced apoptosis
↓ the expression of Bcl-2 and cyclin D1
[477]
Punica granatum leaves, flowers and stem extractsIn vitroU266 cells- inhibited proliferation
- apoptotic effect
- caused cell cycle arrest in G2/M and S phases
[462]
Salvia miltiorrhiza (SM) extractIn vitroU266 and U937 cells- suppressed the growth of U266 and U937 cells
↑ ROS generation and cytotoxic effect was dependent on ROS
- induced ER stress mediated apoptosis
↑ expression of miR-216b and ↓ its target, c-Jun, in U266 and U937 Cells
[478]
Scutellaria baicalensis extractIn vitroRPMI-8226 cells- Scutellaria extract riched in baicalin, wogonoside, baicalein and wogonin
- inhibited the proliferation
↓ the expression level of ABCG2 protein
[135]
Serenoa repensIn vitroU266, RPMI 8226 multiple myeloma cells- induced growth arrest
- induced apoptosis
↑ the expression of cleaved-PARP or p27 protein
↓ basal level of phosphorylated form of STAT 3
↓ IL-6 induced level of phosphorylated form of STAT 3 and ERK
- inhibition of STAT 3 signaling
[479]
Strychnos nux-vomica root extractIn vitroU266B1 cells- anti-proliferative effect
- accumulation of the sub-G0/G1 cell population with consequent decline in other phases of cell cycle
↓ IL-6 and CD-138
[480]
Strychnos nux-vomica root extractIn vitroRPMI 8226- anti-proliferative activity
-the SN-treated cells exhibited significant features associated with apoptosis: cell shrinkage, condensed chromatin, nuclear fragmentation and membrane blebbing
↑ the accumulation of cells at sub-G0/G1 phase
- induced disruption of mitochondrial membrane potential and subsequent leakage of mitochondrial cytochrome c
[481]
Thymus vulgaris and Arctium lappa extractsIn vitroMOLP-8, KMS-11, NCI-H929, RPMI-8226, KMS-12BM, JJN-3,
L-363, AMO-I, and OPM-2 cells
- cytotoxicity in CCRF-CEM and CEM/ADR 5000 cell lines
- TCF induced apoptosis in NCI-H929 cells with a higher ratio, compared to ACF
- ACF demonstrated more potent autophagy activity than TCF
- TCF and ACF induced cell cycle arrest and ferroptosis
[482]
Viscum album QuFrF (VAQuFrF) extractIn vitroMolp-8, LP-1, OPM-2, Colo-677, RPMI-8226, and KMS-12-BM cells- VAQuFrF + vincristine inhibited cell proliferation, arrested the cell cycle phases, and increased the number of apoptotic/necrotic cells.
- VAQuFrF inhibited the proliferation of the cells more effectively than vincristine in Molp-8, LP-1, and RPMI-8226 cells at a dose of 10 μg/105 cells
- VAQuFrF affected the tumour cells mainly via cytostatic effect
[483]
Bioactive compounds from Abelmoschus manihot L.In vitro
In vivo
ARP1 and H929 human MM cell lines and murine MM cell line, 5TMM3VT
C57BL/KaLwRij mice
- HKC improved survival rate of MM-prone mice
- promoted osteoblastogenesis and suppressed osteoclastogenesis in murine cell lines
- HK-11 inhibited MM cells proliferation and ↓ β-catenin signaling
[484]
Legend:↑ increased/upregulated; ↓ decreased/downregulated; ACF—Arctium lappa chloroform fractions; CDKs—cyclin-dependent kinases; CF—chloroform fractions; CRBN—cereblon; EF—ethyl acetate fractions; ERK—extracellular signal-related kinase; HK-11—3-O-kaempferol-3-O-acetyl-6-O-(p-coumaroyl)-β-D-glucopyranoside; HKC—Abelmoschus manihot (L.) Medik. derived Huangkui capsules; HSE—Hibiscus sabdariffa extract; IL-6—interleukin-6; KEGG—Kyoto Encyclopedia of Genes and Genomes; LE—Corchorus olitorius leaf extracts; MYC—myrrh ethanol extracts; PARP—poly (ADP-ribose) polymerase; PBMCs—peripheral blood mononuclear cells; RXC—frankincense ethanol extracts; SE—Corchorus olitorius seed extracts; SM—ethanol extract of Salvia miltiorrhiza; SN—Strychnos nux-vomica L.; SP—side population; STAT 3—signal transducer and activator of transcription 3; TCF—Thymus vulgaris chloroform fractions; VAQuFrF—Viscum album QuFrF; YDC—frankincense-myrrh ethanol extracts; YDS—frankincense-myrrh water extracts.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Cotoraci, C.; Ciceu, A.; Sasu, A.; Miutescu, E.; Hermenean, A. Bioactive Compounds from Herbal Medicine Targeting Multiple Myeloma. Appl. Sci. 2021, 11, 4451. https://doi.org/10.3390/app11104451

AMA Style

Cotoraci C, Ciceu A, Sasu A, Miutescu E, Hermenean A. Bioactive Compounds from Herbal Medicine Targeting Multiple Myeloma. Applied Sciences. 2021; 11(10):4451. https://doi.org/10.3390/app11104451

Chicago/Turabian Style

Cotoraci, Coralia, Alina Ciceu, Alciona Sasu, Eftimie Miutescu, and Anca Hermenean. 2021. "Bioactive Compounds from Herbal Medicine Targeting Multiple Myeloma" Applied Sciences 11, no. 10: 4451. https://doi.org/10.3390/app11104451

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