Next Article in Journal
The Prediction of Antioxidant Q-Markers for Angelica dahurica Based on the Dynamics Change in Chemical Compositions and Network Pharmacology
Next Article in Special Issue
Coumarins from Jinhua Finger Citron: Separation by Liquid–Liquid Chromatography and Potential Antitumor Activity
Previous Article in Journal
An In Vivo Assessment of the Effect of Hexane Extract from Endlicheria paniculata Branches and Its Main Compound, Methyldehydrodieugenol B, on Murine Sponge-Induced Inflammation
Previous Article in Special Issue
Terpinen-4-ol Induces Ferroptosis of Glioma Cells via Downregulating JUN Proto-Oncogene
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Epigallocatechin-3-Gallate Therapeutic Potential in Cancer: Mechanism of Action and Clinical Implications

1
Department of Molecular Biotechnology and Genetics, University of Lodz, Banacha Street 12/16, 90-237 Lodz, Poland
2
Doctoral School of Exact and Natural Sciences, University of Lodz, Banacha Street 12/16, 90-237 Lodz, Poland
3
Centre for Interdisciplinary Research in Basic Sciences, Jamia Millia Islamia, New Delhi 110025, India
4
Department of Pharmacology and Toxicology, College of Pharmacy, King Saud University, Riyadh 11451, Saudi Arabia
5
Department of Pharmacology & Toxicology, College of Pharmacy, Prince Sattam Bin Abdulaziz University, P.O. Box 173, Al-Kharj 11942, Saudi Arabia
6
Department of Medical Biochemistry, Medical University of Lodz, Mazowiecka 6/8, 90-001 Lodz, Poland
7
Doctoral School of Medical University of Lodz, Hallera 1 Square, 90-700 Lodz, Poland
8
Department of Biochemistry, Centre for Drug Discovery, Karpagam Academy of Higher Education, Coimbatore 641021, India
9
Department of Pharmaceutical Chemistry, Faculty of Pharmacy, Erciyes University, Kayseri 38280, Turkey
10
Bioprocess Technology Division, School of Industrial Technology, Universiti Sains Malaysia, Penang 11800, Malaysia
11
Computational Chemistry and Drug Discovery Division, Quanta Calculus, Greater Noida 201310, India
12
Department of Pharmacology, Saveetha Institute of Medical and Technical Sciences, Saveetha Dental College and Hospital, Chennai 600077, India
13
LMCE Laboratory, Group of Computational and Pharmaceutical Chemistry, University of Biskra, Biskra 07000, Algeria
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2023, 28(13), 5246; https://doi.org/10.3390/molecules28135246
Submission received: 9 May 2023 / Revised: 30 June 2023 / Accepted: 3 July 2023 / Published: 6 July 2023
(This article belongs to the Special Issue Anti-tumor Effects of Natural Products)

Abstract

:
Cellular signaling pathways involved in the maintenance of the equilibrium between cell proliferation and apoptosis have emerged as rational targets that can be exploited in the prevention and treatment of cancer. Epigallocatechin-3-gallate (EGCG) is the most abundant phenolic compound found in green tea. It has been shown to regulate multiple crucial cellular signaling pathways, including those mediated by EGFR, JAK-STAT, MAPKs, NF-κB, PI3K-AKT-mTOR, and others. Deregulation of the abovementioned pathways is involved in the pathophysiology of cancer. It has been demonstrated that EGCG may exert anti-proliferative, anti-inflammatory, and apoptosis-inducing effects or induce epigenetic changes. Furthermore, preclinical and clinical studies suggest that EGCG may be used in the treatment of numerous disorders, including cancer. This review aims to summarize the existing knowledge regarding the biological properties of EGCG, especially in the context of cancer treatment and prophylaxis.

1. Introduction

Green tea (Camellia sinensis) consumption is known for its health benefits [1,2]. The majority of green tea’s positive effects on human health can be attributed to the high polyphenol and flavonoid content of the beverage. Catechins, which are the primary flavonoids found in green tea, account for about 30–40% of the solid components of this plant. Epicatechin (EC), epigallocatechin (EGC), epicatechin-3-gallate (ECG), and epigallocatechin-3-gallate (EGCG) are the primary catechins found in tea (Figure 1) [3].
EGCG has gained particular scientific interest in the past decade due to its numerous health benefits resulting from its antioxidant, anti-inflammatory, anti-fibrotic, and anti-cancer properties [4]. EGCG is consumed mainly via brew drinking. One cup of green tea contains approximately 177 mg of EGCG [5,6]. Therapeutic advantages of green tea consumption have been observed in inflammatory diseases and several types of cancer [7,8,9]. The anti-tumor activity of EGCG has been linked to the inhibition of the proteins involved in cellular signaling pathways that are frequently disrupted in cancer. These include the protein networks mediated by epidermal-growth-factor receptor (EGFR) [10,11,12,13], tyrosine-protein kinase JAK (JAK), signal transducer and activator of transcription (STAT) (JAK/STAT pathway) [14,15,16], phosphoinositide-3-kinase (PI3K), RAC-alpha serine/threonine-protein kinase (AKT), serine/threonine-protein kinase mTOR (mTOR) (PI3K/AKT/mTOR pathway) [10,17,18], and mitogen-activated protein kinases (MAPK/ERK pathways) [19,20]. Moreover, EGCG exhibits antioxidant, anti-inflammatory, and anti-angiogenic effects [4] and may promote both caspase-dependent [21,22,23] and caspase-independent cell death [24]. The PubMed database (https://pubmed.ncbi.nlm.nih.gov/; accessed on 29 March 2023) was searched using the combination of the following keywords: “epigallocatechin 3-gallate”, ”EGCG”, “bioavailability”, “epigenetics”, and “signaling pathways”. Additionally, the “clinical trial” option of the PubMed search was chosen to include all the reported clinical trials on EGCG. The rationale behind this study was to gather information regarding the molecular mode of the anticancer activity of EGCG and to discuss the critical concerns associated with the use of EGCG in clinical settings including its bioavailability and toxicity. We aimed to answer the following questions: (a) What are the major pathways associated with the anticancer activity of EGCG? (b) Does EGCG exert its activity through epigenetic influence? (c) What are the emerging targets of EGCG? (d) What evidence has been provided by epidemiological studies and clinical studies of EGCG? (e) What is and what influences the bioavailability of EGCG? (f) Is EGCG use safe and what doses are tolerable in humans?

2. Mechanism of Anticancer Activity of EGCG

Medicinal plants have been used for centuries in the prevention and management of various diseases. Natural products (NPs) constitute sources of bioactive molecules that can be employed in the treatment of human conditions or can be modified to generate new, more efficient, and less toxic drugs with improved bioavailability [25,26,27]. Their therapeutic properties are notably valued in oncology. It is believed that approximately 60% of all oncopharmaceuticals currently available on the market have been designed based on plant-derived compounds [28,29,30]. In the meantime, the multidimensional action mechanisms of these compounds are being discovered along with their abilities to concurrently impact multiple oncogenic signaling pathways. This translates into the capability of cells to either proliferate, migrate, metastasize, or bypass apoptotic cell death. Because of the variety of their properties, natural compounds can be utilized as either direct cytotoxic agents or chemotherapy adjuvants. This allows them to support anticancer activity, overcome the therapy resistance of cancer cells, and enhance the repair processes of normal cells. Furthermore, compared to synthetically obtained compounds, the costly and time-consuming selection of leading structures is avoided, and efforts can be directed at synthesizing and testing new derivatives to obtain better pharmacological responses [31,32].
EGCG’s multidirectional anti-tumor effects have been confirmed in multiple in vitro and in vivo studies investigating the effects of this polyphenol in cancer cells originating from different tissues, including bladder cancer [33,34,35,36,37,38,39,40,41,42,43], breast cancer [21,44,45,46,47,48,49,50,51,52], cervical cancer [53,54,55,56,57,58,59], colorectal cancer [60,61,62,63,64,65], gastric [66,67,68,69], liver [70,71,72,73], lung cancer [11,22,74,75,76], and head and neck cancers [77,78,79].

2.1. Antioxidant Properties

In aerobic organisms, reactive oxygen species (ROS) are produced in the electron transport chain (ETC) or via the activity of catabolic oxidases, anabolic metabolism, and peroxisomal metabolism [80,81,82]. At low levels, ROS operate as cellular messengers in redox-signaling events [83]. However, an excessive production of ROS can lead to DNA damage in the form of roughly 100 distinct oxidative base lesions and 2-deoxyribose alterations [84,85,86]. ROS generation is normally controlled in cells by three distinct mechanisms: the restriction of respiration in the mitochondrial compartment, the complexation of DNA with protective histones, and the elimination of excess ROS by antioxidant enzymes [87,88,89,90,91]. Superoxide radicals, hydrogen peroxide, and hydroxyl radicals are the most prominent ROS species generated in cells [92,93]. EGCG was found to have a free-radical-scavenging activity that allows it to trap hydroxyl radicals and superoxide anions and thus suppress and terminate the free radical chain reaction that occurs during lipid peroxidation. The oxidative-stress-relieving properties of EGCG are achieved via the inhibition of pro-oxidant enzymes such as NADPH oxidase, the activation of antioxidant systems such as superoxide dismutase, catalase, or glutathione, and by decreasing the generation of nitric oxide metabolites by inducible nitric oxide synthase (iNOS) [94].
It would appear that EGCG is the most active of all the catechins, having an antioxidant activity that is 25 to 100 times stronger than that of vitamins C and E [95]. The antioxidant properties of this flavonoid have been extensively described in the literature [3,96,97] and will not be discussed here in detail. Attempts have been undertaken to further increase the antioxidant capacities of EGCG and enhance its lipophilicity. It was found that acetylation of this compound may boost its antioxidant properties [98].

2.2. Suppression of Inflammation

Prolonged chronic inflammation is considered a hallmark of tumor development. It has been found to promote all stages of tumorigenesis. In the process of inflammation, a large number of immune cells, including neutrophils, monocytes, and macrophages, are recruited to the sites of action and secrete various molecules (cytokines and enzymes) that regulate the process. These include tumor necrosis factor α (TNF-α), interleukins (IL-1β, IL-4), matrix metalloproteinases (MMPs), and cyclooxygenases (COXs). An increased expression of many of these inflammatory mediators has been linked to worse prognosis of tumor-bearing patients [99,100,101,102]. The results of multiple studies have demonstrated that EGCG modulates the expression of inflammation-associated genes and proteins such as TNF-α, IL-1β, and MMPs [8]. Here, we focus on the targets of EGCG that exhibit a role in inflammation in the context involving cancer development.
TNF-α is a pleiotropic cytokine that plays an important part in apoptosis and cell survival in addition to its canonical roles in inflammation and immunity. TNFα earned its acronym because of its ability to fight tumors; however, recent research has linked it to numerous other disorders. TNF-α modulates versatile cellular functions as a result of its contact with TNF receptors 1 and 2 (TNFR-1 and TNFR-2). This association stimulates multiple signal transduction pathways including nuclear factor kappa B (NF-κB) and c-Jun N-terminal kinase (JNK), which in turn lead to multiple context-dependent outcomes including apoptosis, necrosis, angiogenesis, immune cell activation, differentiation, and cell migration. While the continuous activation of JNK is associated with cell death, NF-κB is a significant anti-apoptotic factor that promotes cell survival. It should therefore not come as a surprise that TNFα, in a manner that is dependent on the setting, can demonstrate both pro- and antitumoral actions. Hence, the regulation of the activity of the TNFR pathway opens up a wide variety of therapeutic avenues for the treatment of cancer [103,104,105,106,107]. Multiple studies have found that EGCG can suppress TNFα expression and inhibit the activation of TNFα-mediated pathways. However, the studies were not conducted in the context of cancer. For example, EGCG was able to significantly inhibit the expression of monocyte chemoattractant protein-1 (MCP-1) that was stimulated by TNF-α in human umbilical vein endothelial cells (HUVECs). EGCG appeared to inhibit the TNFα-induced activation of NF-κB. In addition, the suppression of 67-kD laminin receptor (67LR) played a significant role in reducing the synthesis of MCP-1 following EGCG treatment [108]. The generation of TNF-α, IL-6, and IL-8 was decreased in human mast cells (HMC-1 cell line) by EGCG. This was accomplished by decreasing the intracellular Ca2+ levels as well as via the repression of the ERK1/2- and NF-κB-mediated signaling pathways [109].
Cyclooxygenases (COXs) are enzymes that are important for the metabolic conversion of arachidonic acid to prostaglandins, particularly PGE2, which is a significant mediator of both inflammation and angiogenesis. COX-1 and COX-2 are two different isoforms of the enzyme. Both are found in cells as membrane-bound proteins, which constitute their integral part. Inside the cell, they are typically found on the luminal side of the endoplasmic reticulum and the nuclear envelope. COX-1 is referred to as a housekeeping enzyme since it is necessary for the upkeep of baseline levels of prostaglandins. In contrast, COX-2 is not often detectable in the healthy tissues and organs of the body but is highly inducible, and its expression can be rapidly increased in reaction to a wide range of proinflammatory agents, such as cytokines or mitogens [110]. COX-2 is commonly over-expressed in a wide variety of malignancies. This enzyme plays a pleiotropic and multidimensional role in the initiation or development of cancer as well as its resistance to chemo- and radiotherapy. Cancer-associated fibroblasts (CAFs) and tumor-associated macrophages (TAMs) that are present in the tumor microenvironment (TME), contribute to the production of COX-2, which in the latter step is released into the tumor setting [111]. It appears that the constitutive expression of COX-2 and the prolonged synthesis of PGE2 play predominant roles in the onset and development of cancer. PGE2 is capable of mediating these effects through a wide variety of signaling pathways, such as the activation of the vascular endothelial growth factor receptor (VEGFR) pathway, which can lead to an increased proliferation, metastatic potential, and angiogenesis of cells and an increased expression of protooncogenes such as BCL-2 [110]. This topic has been extensively discussed by previous authors [110,111,112,113,114,115].
It was demonstrated that EGCG had a strong inhibitory effect on the constitutive overexpression of COX-2 in colon cancer [116] and prostate cancer cells [117]. It was found that EGCG concurrently decreased the activity of the ERK1/2 and AKT pathways and reduced COX-2 promoter activity via its suppressive effect on NF-κB action [116]. Moreover, it was found that a combination of EGCG and COX-2 inhibitor (NS-398) exhibited a synergic anticancer activity, which was indicated by an elevated apoptotic response (increase in the expression of pro-apoptotic BAX protein, pro-caspase-6, and pro-caspase-9 and the cleavage of poly(ADP)ribose polymerase (PARP)) and the inhibition of NF-κB activity [118]. The selective inhibition of COX2 was also observed in EGCG-treated prostate cancer cells (LNCaP and PC-3) [117]. EGCG was found to inhibit early- but not late-stage prostate cancer in vivo, as evidenced by the suppression of cell proliferation and the induction of apoptosis. Furthermore, a decrease in the protein levels of androgen receptor (AR), insulin-like growth factor-1 (IGF-1), IGF-1 receptor (IGF-1R), phosphorylated ERK1 and ERK2, COX-2, and iNOS were observed [119].
Nitric oxide (NO) is a pleiotropic mediator that plays an essential role in a wide variety of biological processes. Some of these functions include vasodilatation or macrophage-mediated immune protection. NOS activity has been found in tumor cells with a variety of different histogenetic origins. This activity has been linked to the rate of tumor proliferation and the expression of significant signaling components that are associated with the development of cancer, such as the estrogen receptor. It would appear that a high production of NO (for example, that generated by activated macrophages) would be cytotoxic or cytostatic for tumor cells, but low levels of NO may have the reverse effect and encourage the growth of tumors. In addition, NO can affect the DNA repair mechanisms of tumors through the up-regulation of genes such as cellular tumor antigen p53 (TP53), PARP, and DNA-dependent protein kinase (DNA-PK) [120,121,122]. It was discovered that EGCG lowers NO generation. This effect may be caused by EGCG through one of two mechanisms: either a decrease in iNOS gene expression or the suppression of enzyme function [123,124].
NF-κB plays a role not only in the control of cellular immunity, inflammation, and stress but also cell differentiation, proliferation, and apoptosis. NF-κB activity and function are often altered in both solid and hematologic malignancies [125]. NF-κB can be activated by a variety of stimuli, including cytokines (such as TNF-α and IL-1β), growth factors (such as EGF), ultraviolet and ionizing radiation, ROS, and DNA damage. The common feature of these triggers is that they ultimately result in the activation of a large cytoplasmic protein complex known as the inhibitor of the κB (IκB) kinase (IKK) complex [126]. NF-κB was found to induce the expression of anti-apoptotic factors such as caspase-8 inhibitor cellular FLICE-like inhibitory protein (FLIP), the inhibitor of the apoptosis proteins c-IAP1/2 and E3 ubiquitin-protein ligase XIAP, or members of the BCL2 family of apoptosis regulators. Moreover, it controls the expression of many angiogenic factors and metastasis-associated proteins including VEGF or MMPs and glucose-metabolism-associated proteins, including glucose transporters (GLUTs) [126].
MMPs belong to the family of zinc-dependent extracellular matrix (ECM) remodeling enzymes that are capable of degrading practically every component of the ECM. The breakdown of the ECM is essential for cancer cells to spread from the primary tumor site to other sites of the body. The degradation of ECM by MMPs not only promotes tumor invasion but also alters the behavior of tumor cells and contributes to the progression of cancer, as previously reviewed [127,128,129]. Several studies have indicated that EGCG may inhibit the activity and expression of MMPs. In line with this evidence, a molecular docking study confirmed these inhibitory effects. In their in silico studies, Chowdhury et al. showed that EGCG and ECG exhibit superior inhibitory properties towards proMMP2 compared with MMP2 [130]. Furthermore, Sarkar et al. employed a molecular docking approach and revealed that EGCG may inhibit proMMP9 and MMP9. The enhanced interaction between EGCG/MMPs compared to EC and EGC may be attributed to the presence of a gallate group in the EGCG structure. In particular, EGCG’s galloyl moiety binds to Phe201, His401, Glu402, His411, and Pro421 residues in the MMP-9 structure. The inclusion of a galloyl residue in EGCG and ECG contributes to a binding affinity that is approximately 1.5-fold higher than that of EC and EGC [131]. These findings are consistent with earlier zymographic studies conducted by Maeda-Yamamoto et al. [132]. Furthermore, it was shown that EGCG inhibited the migration and invasion of T24 human bladder cancer cells through the inhibition of the PI3K/AKT pathway that further conferred the inactivation of NF-κB and the down-regulation of MMP-9 expression, ultimately limiting the metastatic potential of the cells [40,42]. A similar effect was observed for SW780 bladder cancer cells, where EGCG down-regulated the expression of NF-κB and MMP-9 and triggered the apoptosis of cancer cells [33]. EGCG was tested for synergic anticancer effects with a commonly used anticancer agent—doxorubicin (DOX). EGCG was found to boost DOX’s ability to induce apoptosis. In addition, EGCG enhanced DOX’s ability to prevent the migration of bladder cancer cells. According to the findings of mechanistic investigations, the combination of DOX and EGCG suppressed the expression of phosphorylated NF-κB and E3 ubiquitin-protein ligase Mdm2 (MDM2) while simultaneously elevating the expression of TP53 in tumor cells in vivo [34]. Similar observations were made in breast cancer cells, where EGCG suppressed the activation of hypoxia-inducible factor 1-alpha (HIF-1α) and NFκB as well as VEGF expression [47]. Cervical cancer cells (HeLa cell line) treated with EGCG and black tea polyphenol theaflavins displayed a significant inhibition of proliferation, an increase in the number of cells in the sub-G1 cell-cycle phase, and a reduction in the mitochondrial membrane potential (MMP) alongside an increase in the production of ROS, an overexpression of TP53, a change in the BAX/BCL-2 ratio, a release of cytochrome-c, procaspase-3, -9 activation, and PARP cleavage. In addition, both EGCG and theaflavins were able to suppress the activation of AKT and NF-κB, which resulted in the down-regulation of COX2 and cyclin D [55]. Furthermore, EGCG may down-regulate the expression of MMP-9 and therefore decrease the invasive potential of HeLa cells [56]. The anti-inflammatory potential of EGCG is presented in Figure 2.

2.3. Modulation of Epigenetic Targets

The recent decades of cancer research have revealed that epigenetic mechanisms play crucial roles in both the onset and development of tumors. Genome-wide DNA hypomethylation and site-specific DNA hypermethylation, histone modifications, micro RNAs, long non-coding RNAs, and nucleosome remodeling dysregulations are all common epigenetic alterations observed in cancer. An evident therapeutic potential for reversing these epigenetic changes has been demonstrated in a variety of lymphomas, and similar results are beginning to appear in the treatment of solid tumors [133,134]. Thereby, the use of selective agents (termed epi-drugs) to target the cancer epigenome or cancer epigenetic dysregulation is an emerging and exciting strategy for cancer prevention and treatment [135].
DNA methylation is one of the epigenetic regulatory mechanisms that was discovered first and has received the most attention in cancer research. DNA methyltransferases (DNMTs) and DNA demethylases are primarily responsible for the cooperative regulation that determines the level of methylation across the genome. Using S-adenosylmethionine (SAM) as a donor, DNMTs add a methyl group (CH3) to the C5 position of the cytosine ring in cytosine-phosphate-guanine (CpG) islands located in the promoters of genes [136]. DNMTs such as DNMT1, DNMT3A, and DNMT3B write and control patterns of DNA methylation in mammalian cells. DNMT1 is typically responsible for the proper methylation pattern on the daughter strand copies produced from parent DNA during the replication process. In contrast, de novo methyltransferases 3A and 3B are essential in the process of genomic imprinting during embryogenesis and germ cell development. The overexpression of DNMTs has been linked with carcinogenesis and poor disease prognosis [137,138,139,140]. Moreover, many tumors exhibit somatic mutations in DNMTs, which significantly contribute to malignant transformation [141].
A time-dependent inhibition of the enzymatic activity of DNMTs was observed in HeLa cells treated with EGCG. The expression of DNMT3B was also found to decrease over time in EGCG-treated HeLa cells. Thereby, it was suggested that EGCG may serve as a potential epigenetic modifier by altering the methylation status of genes encoding retinoic acid receptor beta (RARβ), cadherin 1 (CDH1), and death-associated protein kinase 1 (DAPK1) tumor suppressors to reactivate their expression [57]. RARβ is a nuclear receptor that is crucial in cell differentiation, apoptosis, and proliferation. Research has indicated that RARβ is commonly repressed or silenced in multiple cancer types, and its loss has been linked with resistance to therapy and tumor progression. RARβ can halt the cell cycle at the G1/S phase [142,143]. CDH1 (E-cadherin), is a molecule crucial for cell adhesion, polarity, and epithelial tissue integrity. Mutations or losses of CDH1 are frequently observed in cancer and contribute to the invasion and metastasis of tumors [144]. DAPK1 is a protein kinase that regulates tumor suppression, autophagy, and apoptosis. DAPK1 is frequently down-regulated, which promotes tumor formation, metastasis, and resistance to chemotherapy [145].
Human epidermoid carcinoma A431 cells treated with EGCG showed a dose-dependent reduction in global DNA methylation. DNMT activity, mRNA, and protein levels for DNMT1, DNMT3a, and DNMT3b were all reduced by EGCG with a concomitant decrease in 5-methylcytosine abundance. Additionally, EGCG lowered histone deacetylase activity (HDAC) and increased acetylated lysine levels in histone H3 (H3K9 and H3K14) and histone H4 (H4K5, H4K12, and H4K16). In most cases, these acetylation events contribute to the open chromatin state, which makes DNA more accessible to transcription factors and other proteins that regulate gene expression. Analogously, an observed decrease in methylated H3K9 leads to an open chromatin state. P16INK4a and P21/P27 tumor suppressor genes that had been silenced were re-expressed in response to EGCG treatment. The primary function of p16INK4a is to suppress the activities of CDK4/6, which controls the cell cycle by facilitating the transition from the G1 phase to the S phase. In contrast, P21 is a protein that serves as a regulator of the cell cycle as well as cellular responses to DNA damage. The primary function of P21 is to regulate the G1/S checkpoint, which is the moment of the cell cycle at which the cell decides whether to continue the cell division or enter a non-dividing state [146].
Histone and non-histone proteins are frequently modified post-translationally with reversible acetylation. Two classes of enzymes, histone acetyltransferases (HATs) and HDACs, work against each other to acetylate and deacetylate proteins, respectively. Both an abnormal expression of HDACs and aberrant protein acetylation, especially histones, have been linked to cancer. Therefore, HDACs have become interesting therapeutic targets for cancer [147]. EC, EGC, and EGCG were indicated as HAT inhibitors in AR-dependent prostate cancer cells. EGCG showed the strongest inhibitory potential among the catechins studied and suppressed AR acetylation and its translocation from the cytoplasm to the nucleus [148]. As nuclear factor NF-kappa-B p65 subunit (P65) acetylation by P300/CBP is required for NF-κB activation, decreasing P65 acetylation may constitute a useful therapeutic approach for the alleviation of persistent inflammation. EGCG was discovered to block HATs and, as a result, reduce the P300-dependent acetylation of P65 in vitro and in vivo, elevate cytosolic IκBα levels, and inhibit NF-κB activation in response to TNF-α. In addition, EGCG therapy suppressed P65 acetylation and NF-κB target gene expression in response to multiple stressors. The necessity of a proper balance between HATs and HDACs in the NF-κB-mediated inflammatory signaling pathway was further highlighted by the observation that EGCG inhibited p300 binding to the promoter region of the IL-6 gene and resulted in an enhanced recruitment of HDAC3. These findings suggest that EGCG can be helpful in the prevention of EBV-induced B lymphocyte transformation [149]. Furthermore, the study of Borutinskaitė et al. showed that EGCG was able to suppress the proliferation of NB4 and HL-60 cells, induce apoptosis, and stimulate the gene expression of proteins that are involved in the cell cycle arrest and the differentiation of cells (e.g., P27, histone acetyltransferase KAT2B (PCAF), and CCAAT/enhancer-binding proteins (C/EBPa and C/EBPE)). EGCG down-regulated the expression of DNMT1 and HDAC1/2 as well as histone methyltransferase G9a and polycomb-repressive complex 2 (PRC2), which are involved in chromatin remodeling. These findings suggested that EGCG could be an effective epigenetic therapeutic agent that could be used in the treatment of acute promyelocytic leukemia (APL) [150].
The abnormal epigenetic silencing of the tissue inhibitor of the matrix metalloproteinase-3 (TIMP-3)-encoding gene, which functions to negatively limit the activity of MMPs, has been linked to the progression and metastasis of breast and prostate cancers. EGCG was shown to mediate the epigenetic elevation of TIMP-3 levels and play a critical role in the regulation of the enzymatic activity of MMP-2 and MMP-9 in breast cancer cells. This was associated with a reduction in the levels of enhancers of zeste homolog 2 (EZH2) and class I (HDAC1). EZH2 works as a histone methyltransferase that catalyzes the addition of a methyl group to lysine 27 of histone H3 (H3K27) and functions as a part of the catalytic subunit of PRC2. It was shown that transcriptional activation of TIMP-3 was related to decreased EZH2 localization and H3K27 trimethylation enrichment at the TIMP-3 promoter, along with a simultaneous rise in histone H3K9/18 acetylation. H3K27 trimethylation introduced as a PRC2 works as a repressive mark that is associated with transcriptional silencing and the formation of heterochromatin. Additionally, it provides a binding site for proteins that contain a chromodomain, such as polycomb group (PcG) proteins. In contrast, H3K9/18 acetylation is associated with an open chromatin state [151,152].
A synergistic induction of apoptosis and an up-regulated expression of growth arrest and DNA-damage-inducible gene 153 (GADD153), death receptor 5 (DR5), and P21 were shown following the treatment of human lung cancer cells (PC-9 cell line) with a combination of EGCG and synthetic retinoid (Am80). Despite no change in acetylation levels in histones H3 or H4, the combination boosted acetylated TP53 and acetylated tubulin levels through a decrease in HDAC activity in the cytosol fraction. Additionally, the combination decreased the protein levels of HDAC4, −5, and −6 by 20% to 80%. By inhibiting HDAC4, −5, and −6, EGCG and Am80 altered acetylation levels in nonhistone proteins and induced the cell death of cancer cells [153]. The modulation of epigenetic targets by EGCG is shown in Figure 3.

2.4. Modulation of Cellular Signaling Molecules by EGCG

2.4.1. EGFR Pathway

The EGFR pathway plays a major role in epithelial cell proliferation, survival, and differentiation as well as autophagy and cell metabolism [154,155]. The disruption of its function often leads to neoplasia and therapy escape mechanisms [156,157]. In several human cancers, including glioblastoma, colorectal cancer, breast cancer, and lung cancer, the EGFR gene is often amplified and mutated and the EGFR protein is overexpressed, leading to chronic receptor activation, excessive signaling, and the promotion of neoplastic transformation [156,158,159].
Normally, EGFR is situated on the cell surface. While not activated by one of seven possible ligands, it remains a monomer with minimal kinase activity [160,161]. After ligand activation, EGFR can either create a homodimer or a heterodimer with other receptors from the ErbB/HER tyrosine kinase receptor family, with the second option usually generating a stronger signal transduction and more oncogenic potential [162,163]. These conformational changes lead to the autophosphorylation of tyrosine residues in the receptor tail located in the cytoplasm, triggering a signaling cascade [163]. The cellular response is dependent on the specific ligand that has activated EGFR. The reaction usually involves the activation of multiple signaling pathways, such as the RAS/MAPK, PI3K/AKT, and protein kinase C (PKC) signaling pathways [154,164].
The PKC pathway is involved in controlling numerous cellular functions such as cell growth, differentiation, survival, and apoptosis. The PKC family consists of at least ten isoforms that are categorized into three groups: classical (α, βI, βII, and γ), novel (δ, ε, η, and θ), and atypical (ζ and ι/λ). The deregulation of PKC signaling has been associated with the onset and advancement of various types of cancer, including breast, lung, colon, and pancreatic cancer. PKC isoforms have been found to play a role in cancer cell proliferation, invasion, metastasis, and resistance to chemotherapy and radiation therapy [165]. Multiple types of research have indicated that the PKC pathway plays a crucial role in cancer development, and different treatment methods aimed at PKC signaling are currently being researched. One of these approaches involves the application of PKC inhibitors, which have exhibited encouraging outcomes in both preclinical studies and clinical trials, implying that targeting the PKC pathway could be a viable cancer therapy method [166].
It was shown that EGCG may work as a tyrosine kinase inhibitor. For instance, it has been observed that advanced non-small-cell lung cancer (NSCLC) showed a good response to EGCG treatment. EGCG not only decreased the EGFR expression but also its downstream protein phosphorylation. However, those results were observed in a cell line with a wild-type EGFR gene [167]. It appears that mutations occurring in EGFR, which are common in NSCLC, influence EGCG treatment efficacy. In a study that compared the sensitivity of three different NSCLC cell lines to EGCG, resistance to treatment was observed for two cell lines harboring mutations in EGFR, indicating a strong association between the anticancer activity of EGCG and EGFR mutation status. Mutations in EGFR appeared to have an impact on its ability to bind ECGC, so the receptor still could undergo phosphorylation [11]. Nevertheless, the results remain inconclusive, as there are also studies reporting no significant difference in the mutated EGFR cell line’s response to ECGC [168].
As described in further investigations, EGCG was tested in various types of cancer cells, usually with the outcome indicating that EGCG blocks the EGFR pathway via binding with the EGFR receptor. In a study examining its impact on three human thyroid carcinoma cell lines—TT, TPC-1, and ARO—EGCG was shown to decrease the protein levels of phosphorylated EGFR in concentrations of 50–200 µM [169]. In hepatocellular carcinoma cells, a down-regulation of EGFR expression and one of its ligands—HER2—was observed following treatment with EGCG. However, no clear dose dependency was established [170]. Moreover, a short pretreatment (30 min) with 100 and 200 µM EGCG was proven to partially inhibit the EGFR phosphorylation induced by EGF in an A-431 epidermoid carcinoma cell line [171]. Furthermore, in a human salivary gland adenoid cystic carcinoma cell line SACC-38, EGCG (in concentrations of 5, 10, 20, 40, and 80 μM for 48 h) significantly decreased EGFR protein levels and inhibited EGFR phosphorylation [172].

2.4.2. JAK/STAT Pathway

The JAK/STAT signaling pathway constitutes one of the major cell signaling hubs in cellular homeostasis. This evolutionarily conserved pathway can be activated by numerous hormones, growth factors, interferons (IFNs), interleukins (ILs), and colony-stimulating factors and is involved in the control of a variety of processes, including apoptosis, survival, and inflammation [173,174]. In the traditional JAK/STAT pathway, the interaction between the ligand and its receptor triggers the dimerization and transphosphorylation of the second. This facilitates the formation of a docking site for STATs. JAK phosphorylates STATs at this docking location, and after that, STAT dissociates from the receptor and forms homodimers or heterodimers through interactions involving the SH2 domain and phosphotyrosine. These dimers then translocate to the promoters of the target genes, where they regulate their transcription. STATs either bind to their DNA target sites to activate transcription or associate with non-STAT DNA-binding factors to promote STAT-dependent transcription. Both STAT and non-STAT transcription factors can synergistically activate transcription by binding to clusters of independent DNA-binding sites. The deregulation of this pathway has been frequently linked to various malignancies including breast cancer, glioblastoma, hepatocellular carcinoma, Hodgkin’s lymphoma, myeloproliferative neoplasm, natural killer/T-cell lymphoma, non-small-cell lung cancer, prostate cancer, thyroid cancer, and others, as previously reviewed by multiple authors [175,176,177,178,179].
According to research carried out by Tang and colleagues, EGCG inhibited the expansion, invasion, and migration of pancreatic cancer cells while simultaneously inducing apoptosis via interfering with the STAT3 signaling pathway in vitro. In addition, EGCG contributed to an increase in the gemcitabine and CP690550 efficacy in AsPC-1 and PANC-1 pancreatic cancer cell lines [180]. It was also found that the treatment of breast cancer and head and neck cancer cell lines with EGCG decreased the phosphorylation of the ERK, STAT3, and EGFR proteins. The fact that EGCG induced a decrease in the level of BCL-2 and BCL-X(L) proteins and increased BAX protein expression and the activation of caspase 9 lends credence to the hypothesis that EGCG triggers apoptosis by the mitochondrial apoptosis pathway. The growth-inhibiting effects of 5-fluorouracil were significantly amplified by the addition of EGCG. These findings provide insights into the molecular mechanisms of growth suppression by EGCG and suggest that this molecule may be effective in the chemoprevention and/or therapy of breast cancer and head and neck squamous cell carcinoma (HNSCC) when used alone or in conjunction with other drugs [13,181]. It has been discovered that EGCG can limit the proliferation of chronic myeloid leukemia cells by inducing apoptosis via the suppression of the BCR/ABL oncoprotein and managing its downstream pathways, such as p38-MAPK/JNK and JAK2/STAT3/AKT [182]. The pretreatment of cholangiocarcinoma (CCA) cells with quercetin and EGCG inhibited the activation of the JAK/STAT pathway caused by the IL-6 and INF-γ treatment. This was demonstrated by a dose-dependent reduction in the levels of phosphorylated STAT1 and STAT3 proteins. The quercetin and EGCG pretreatments were successful in inhibiting the cytokine-mediated up-regulation of iNOS and intercellular adhesion molecule-1 (ICAM-1) via the JAK/STAT cascade. In addition, these flavonoids have the potential to suppress the proliferation of CCA cells as well as cytokine-induced cell migration [183]. Experiments conducted in in vitro settings have shown that treatment with EGCG suppressed INFγ-induced JAK-STAT signaling as well as the production of immune-checkpoint molecules, namely programmed death-ligand 1 and 2 (PD-L1 and PD-L2). It was demonstrated that this impact was caused by a decreased expression of STAT1 as well as its phosphorylation, which led to a reduction in the activity of the regulator interferon regulatory factor 1 (IRF1) for PD-L1/PD-L2 in human and mouse melanoma cells. The in vivo tumor-inhibitory action of EGCG was shown to be mediated by CD8+ T lymphocytes, and the effect was comparable to that of anti-PD-1 therapy in animals. However, their modes of action were rather distinct from one another. EGCG was able to decrease JAK/STAT signaling as well as PD-L1 expression in tumor cells, which resulted in the reactivation of T cells. This was in contrast to the anti-PD-1 treatment, which prevented the interaction between PD-1/PD-L1. In conclusion, it was shown that EGCG inhibits JAK-STAT signaling in melanoma, which in turn increases anti-tumor immune responses [15]. By using surface-plasmon-resonance (SPR)-binding assays and in silico docking studies, Wang et al. showed that EGCG disrupted STAT3 peptide binding at micromolar concentrations, and the docking experiments showed that EGCG interacts with Arg-609, which is one of the pivotal residues in the STAT3 SH2 domain responsible for STAT3 and phosphorylated peptide binding [184]. A significant inhibition of tyrosine and serine phosphorylation of STAT 1 and STAT3 was seen in human oral cancer cells [185,186] and cholangiocarcinoma cells [183] treated with EGCG.

2.4.3. MAPKs Pathways

MAPK pathways influence gene expression and play an essential role in plenty of cellular processes that decide a cell’s fate, such as proliferation, differentiation, apoptosis, survival, and other stress responses. MAPK pathways consist of a three-step kinase cascade of MAP3K, MAP2K, and MAPK kinases and are tightly regulated by other factors including phosphatases as well as their pseudoenzyme equivalents [187,188]. Traditionally, among the most recognized MAPK pathways, ERK is classified as a mitogen-responsive MAPK, while JNK and p38 are MAPKs engaged in a reaction to stressors. However, in reality, the response pathways often overlap. In total, 40% of all human cancers have RAS-RAF-MEK-ERK pathway malfunctions, which makes MAPK pathways an excellent target of anti-cancer therapies [189].
EGCG was found to inhibit the p38-MAPK/JNK pathway in chronic myeloid leukemia cells in a dose-dependent manner. The experiment was carried out on five cell lines—K562, K562R, KCL-22, BaF3/p210, and BaF3/p210T315I—including those unresponsive towards a currently used chemotherapeutic, imatinib. Not only was EGCG effective in imatinib-resistant cells, causing a decrease in the cells’ viability, but it was also more cytotoxic towards leukemic cells in comparison to bone marrow mononuclear cells isolated from healthy donors. Interestingly, it was observed that EGCG-induced apoptosis was triggered by the apoptosis-inducing factor (AIF) and was executed in a caspase-independent pathway, including autophagy. EGCG was shown to down-regulate both P38 and phosphorylated P38 levels in leukemic cells while up-regulating phosphorylated JNK, though did not influence JNK and ERK protein levels [182].
Though EGCG is usually believed to inhibit MAPK pathways, in an experiment on hepatocellular carcinoma cells, it caused the phosphorylation of ERK. It was shown that a prolonged activation of ERK induced cell apoptosis [170]. Moreover, when EGCG was applied to a hepatocarcinoma cell line (HepG2), its proapoptotic activity was stimulated by the JNK pathway [190]. However, in a study on human bladder cancer cells (T24), a 30 min pretreatment with EGCG before the application of IL-1β caused the suppression of both the ERK1/2 and JNK pathways by silencing their transcriptional activity, while EGCG, in a concentration range of 0–50 μM, did not affect the T24 cells’ viability [191]. In a study by Chen et al., EGCG was indicated as an estrogen receptor (ERα36) inhibitor in Hep3B cells, where it showed both anti-proliferative and pro-apoptotic properties, leading to the inhibition of the ERα36-EGFR-HER-2 feedback loop or the PI3K/AKT and MAPK/ERK pathways [170].

2.4.4. PI3K/AKT/mTOR Pathway

The PI3K/AKT/mTOR signaling hub regulates a wide variety of cellular activities, including cell survival, proliferation, growth, metabolism, angiogenesis, and metastasis. This pathway is hyperactivated or disrupted in many kinds of cancer [192]. Signaling from receptor tyrosine kinases or the binding of growth factors to PI3Ks trigger the phosphorylation of the hydroxyl groups (3′OH) of phosphatidylinositol. Phosphorylation of phosphatidyl-1,4-bisphosphate (PIP2) leads to the formation of phosphatidylinositol-1,4,5-triphosphate (PIP3), which recruits phosphoinositol-dependent kinase 1 (PDK1) and AKT kinase to the vicinity of the cell membrane. The phosphorylated, active AKT phosphorylates proteins involved in cell growth and survival and induces the expression of many proteins with proliferative and anti-apoptotic functions, including BCL-2, BCL-xL, and NF-Kb. Another growth regulator activated downstream of PI3K/AKT is mTOR serine/threonine kinase. AKT activates mTOR, which then phosphorylates ribosomal protein kinase S6 (S6K) and inactivates eukaryotic translation initiation factor 4E binding protein (4E-BP1). mTOR coordinates cell growth mainly through increased mRNA translation. An important regulator of the pathway is the phosphatidylinositol 3,4,5-trisphosphate 3-phosphatase and dual-specificity protein phosphatase (PTEN) protein, which dephosphorylates PIP3 and thus prevents the activation of the AKT protein. The details of the PI3K-AKT-mTOR signaling network and its targeting were previously and extensively described by other authors [193,194,195,196,197,198] and will not be discussed here in detail.
EGCG can inhibit the proliferation of pancreatic carcinoma PANC-1 cells and bladder cancer T24 and 5637 cells or promote their apoptosis through the up-regulation of PTEN expression while simultaneously suppressing the phosphorylation and expression of AKT and mTOR [38,199]. Similar observations were made when the effect of EGCG was investigated in lung cancer cells (H1299 cell line), where a down-regulation of phosphorylated PI3K and AKT was shown following EGCG treatment [22]. The study of Ding and Yang also examined whether EGCG inhibits colon cancer cell growth in vitro and nude mouse xenografts by targeting the sonic hedgehog (SHH) and PI3K pathways. EGCG inhibited colon cancer cell proliferation dose-dependently with negligible toxicity to normal colon epithelial cells. EGCG suppressed colon cancer cell migration and invasion and induced their apoptosis. Furthermore, the use of purmorphamine (smoothened (Smo) agonist) or IGF-1 (PI3K agonist) partially blocked EGCG’s inhibitory effects on cell proliferation, migration, and death [200]. Numerous other studies have indicated that EGCG induces apoptosis via the inhibition of the PI3K/AKT/mTOR pathway, which leads to the promotion of cytotoxic autophagy. The anti-cancer properties of EGCG have also been hypothesized to be caused by its pro-oxidant activity. When used in conjunction with chemotherapy treatment, EGCG suppressed both the proliferation of cancer cells and the drug’s induced pro-survival autophagy, as recently reviewed by Ferrari et al. [201]. An overview of the inhibitory effects of EGCG on the key signaling pathways associated with the development of cancer is shown in Figure 4.

2.4.5. Other Signaling Pathways

The 67-kDa laminin receptor (67-LR) is located on the cell surface and is a key factor in cell adhesion tumor growth, invasion, and metastasis [202,203]. Its overexpression is found in various types of tumors, and recently 67-LR has been appointed as a single-target molecule interacting with EGCG at physiologically achievable concentrations (0.1–1 μM) [204,205]. Hepatoma cells treated with radiotherapy can develop invasive potential due to the activation of hepatic stellate cells. It was observed that a 48 h EGCG pretreatment of LX-2 cells (human hepatic stellate cell line) inhibited the toll-like receptor 4 (TLR4) pathway by binding with 67-LR. This resulted in diminished invasive properties of the irradiated cells [206]. Lipid rafts are microdomains of the cell membrane that are rich in cholesterol and sphingolipids and play an important role in cell signaling and communication. Cellular functions such as adhesion, migration, and proliferation have all been linked to these microdomains. Lipid rafts have been linked to tumor development, invasion, and metastasis. Lipid raft disruption can alter the positioning and function of signaling molecules like EGFR and HER2 that play a role in cancer cell proliferation and survival. Through enhancing invadopodia formation and metalloproteinase activity, lipid rafts can also encourage cancer cell movement and invasion. It has been suggested that lipid rafts could be used as a target in the treatment of cancer [207,208,209,210]. Treating human multiple myeloma U266 cells with EGCG resulted in plasma membrane lipid raft disruption via the 67-LR/acid sphingomyelinase (ASM) pathway and the elicitation of EGCG-induced cell death. Moreover, the pretreatment of cells with anti-67-LR antibodies protected them from EGCG influence. As 67-LR is overexpressed on the surface of cancer cells, EGCG’s cell-killing properties appear to be cancer cell-specific. However, high doses of EGCG are hepatotoxic, and the concentration of EGCG needed for activating the lipid raft disruption process seems unachievable in therapy. Sphingosine kinase-1 inhibitor (safingol) significantly sensitizes multiple myeloma cells to EGCG and could help in reducing the dose of the phytochemical [211]. EGCG triggered the protein phosphatase 2A (PP2A) pathway via 76-LR interaction in melanoma cells (Hs294T, A375, MeWo cell lines), thus limiting tumor growth. As in earlier studies, primary normal human melanocyte cells were unaffected, confirming EGCG’s cancer cell discrimination [212].
Energy homeostasis in cells is controlled by the AMP-activated protein kinase (AMPK) pathway. The activation of AMP-activated protein kinase (AMPK) suppresses cell growth and proliferation by inhibiting anabolic pathways and stimulating catabolic pathways. The dysregulation of the AMPK pathway has been linked to cancer progression, metastasis, and drug resistance. The activation of AMPK has been proven in multiple studies to reduce the survival and growth of cancer cells. Important steps in the development of cancer, such as angiogenesis and inflammation, can be suppressed by activating AMPK. In addition, activating AMPK can make cancer cells more sensitive to chemotherapy and radiotherapy, making it a promising target for cancer treatment. However, other research has revealed that AMPK activation can increase tumor development and survival in specific circumstances, suggesting that AMPK’s role in cancer may be nuanced and situation-specific [213,214,215]. There is mounting evidence that EGCG can stimulate ROS production, which in turn leads to the phosphorylation and activation of AMPK. Some of the proteins that are involved in adipogenesis, lipogenesis, and lipolysis can be modulated by phosphorylated AMPK in response to EGCG treatment. This is known as the “AMPK hypothesis”, which was proposed by Yang et al. [216]. AMPK activation by EGCG can lead to the inhibition of the mTOR pathway, which is involved in regulating cell growth and proliferation, as well as the activation of autophagy, which can promote cell death. EGCG-mediated activation of AMPK has also been shown to sensitize cancer cells to chemotherapy and radiotherapy [217,218].
Focal adhesion kinase (FAK) is a protein that plays a critical role in cell signaling and the regulation of cell behavior. The FAK pathway is involved in several cellular processes including cell adhesion, migration, proliferation, and survival. The dysregulation of FAK signaling has been linked to the development and progression of several types of cancer including breast, lung, and pancreatic cancer. FAK is known to promote tumor growth, invasion, and metastasis as well as resistance to chemotherapy and radiotherapy. Targeting the FAK pathway has emerged as a promising approach for cancer therapy [219,220,221]. EGCG treatment can suppress FAK activity, leading to reduced cancer cell migration, invasion, and proliferation. Moreover, EGCG-mediated inhibition of the FAK pathway has been demonstrated to sensitize cancer cells to chemotherapy and radiotherapy. These findings suggest that EGCG has potential as an anti-cancer agent that targets FAK signaling in cancer cells [222,223,224].
The hedgehog (Hh) signaling pathway plays a crucial role in cell growth, differentiation, and tissue development. The aberrant activation of this pathway has been linked to various types of cancer, including basal cell carcinoma, medulloblastoma, and gastrointestinal, lung, breast, and prostate cancers. Activation of the Hh pathway in cancer leads to an increased expression of target genes that promote cell proliferation, survival, and invasion. Thus, targeting the Hh pathway has emerged as a promising strategy for cancer therapy [225,226]. The expression of Hh pathway components such as smoothened (SMO) and GLI1 have been discovered to be suppressed by EGCG, ultimately leading to the inhibition of the Hh pathway [227]. EGCG has been proven to reduce the proliferation, migration, and invasion of cancer cells through the blockade of the Hh pathway [41,200].
The NOTCH pathway is a signaling pathway that is important in determining cell fate, proliferation, and differentiation. Its malfunction has been implicated in the growth and spread of different types of cancer including leukemia, lung cancer, and breast cancer. The key components of the NOTCH pathway include NOTCH receptors (NOTCH1-4), ligands (JAGGED1, JAGGED2, DELTA-like1, DELTA-like3, and DELTA-like4), and downstream transcription factors such as HES and HEY. The abnormal activation of the NOTCH pathway has been linked to increased cancer cell growth, survival, invasion, and resistance to chemotherapy and radiation therapy. On the other hand, blocking the NOTCH pathway has been shown to trigger cancer cell differentiation and cell death, and it makes them more responsive to therapy. Hence, inhibiting the NOTCH pathway has become a promising approach for treating cancer, and various NOTCH pathway inhibitors are currently under development for cancer treatment [228,229,230]. EGCG was found to be a NOTCH signaling inhibitor [231]. For example, Wei et al. found that by preventing the activation of the NOTCH signaling system, EGCG dramatically reduced the proliferation of and promoted apoptosis in tongue cancer cells [232]. In another study, EGCG was shown to influence the cell cycle, trigger apoptosis, and suppress the growth of colorectal cancer cells in vitro and in vivo. The expressions of HES1 and NOTCH2 were suppressed after EGCG treatment [65].
Mutations in the genes encoding components of the WNT/β-catenin pathway, such as APC and β-catenin, can cause the accumulation of β-catenin in cancer cells. This accumulation results in the activation of the WNT/β-catenin pathway, which encourages cell proliferation, survival, and invasion. This pathway’s dysfunction has been linked to various cancers, including colon, liver, breast, and lung cancers [233,234]. EGCG was found to exercise its anticancer action by boosting the phosphorylation and proteasomal degradation of β-catenin through a mechanism that is independent of both glycogen synthase kinase-3 (GSK-3) and protein phosphatase 2A (PP2A). Zhu et al. found that EGCG down-regulated WNT/β-catenin activation in lung cancer stem cells [235].

3. Emerging EGCG Targets

In the early phases of the drug design process, computational techniques supplement experimental analysis by providing important insights into the comprehension of the molecular mechanisms of compound activity. Molecular docking is a computational strategy used to anticipate the binding modes of small compounds or macromolecules with a receptor. This methodology allows the ranking of compounds based on a hierarchy established by specific scoring functions. When coupled with the more expensive but also more precise molecular dynamics (MD) techniques, docking gains a major complementary tool [236]. This approach has been found to be useful in the identification and optimization of many currently approved drugs including captopril, oseltamivir, and orzanamivir [237].
Glucose-regulated protein 78 (GRP78) is a crucial component of the unfolded protein response. It works as a chaperone protein and is up-regulated in cancer cells, conferring the inhibition of apoptosis and the promotion of chemoresistance in treated cancer cells. GRP78 is composed of an ATPase domain, a substrate-binding domain, and a linker region. ATP-competitive inhibitors such as EGCG suppress GRP78 activity and hinder its expression in glioblastomas. The in silico studies of Bhattacharjee et al. revealed crucial amino acids involved in GRP78-EGCG binding, including Ile61, Glu293, Arg297, and Arg367 [238]. Furthermore, apoptosis induction, MMP changes, and ROS reduction were observed in colorectal cancer cells treated with EGCG alone or in conjunction with irinotecan. When given alongside irinotecan, EGCG has been shown to increase the chemosensitivity of colon cancer cells by inducing GRP78-mediated endoplasmic reticulum stress [63]. Similar observations were made in breast cancer cells treated with microtubule-interfering agents (taxol and vinblastine), further confirming the role of EGCG as a GRP78 inhibitor [239,240].
Khan et al. used a structure-based virtual screening of phytocompounds to identify possible inhibitors of hexokinase 2 (HK2) [241]. HK2 is the most active of the family’s isozymes and is primarily expressed in insulin-sensitive tissues. The induction of HK2 in the majority of neoplastic cells leads to their metabolic shift towards aerobic glycolysis, and the genetic deletion of HK2 prevented the progression of malignant growth in mouse models [242]. EGCG and quercitin were identified as two compounds with a significant capability of HK2 inhibition in docking and MD simulations. Based on the outcomes of MD simulations, it was discovered that HK2 consistently creates stable protein–ligand complexes with EGCG and quercitrin throughout the simulation process. In general, these findings point to the possibility that EGCG and quercitrin could be further investigated as potential scaffolds in the process of developing drugs that target HK2 [241]. This is consistent with earlier studies on the impact of EGCG on glycolysis, where the treatment of human tongue carcinoma cells resulted in the down-regulation of HK2 [243]. Other studies suggested a more versatile function of EGCG in glycolysis, where it reduced the activity and expression of several enzymes including phosphofructokinase (PFK) and pyruvate kinase (PK) in pancreatic cells in vitro; however, this was without the influence of HK expression. Furthermore, EGCG treatment resulted in a decrease in the levels of platelet-type phosphofructokinase (PFKP) and pyruvate kinase M2 (PKM2) in pancreatic tumor xenograft homogenates acquired from mice [244].
In chronic lymphocytic leukemia (CLL), the expression of zeta-chain-associated protein kinase-70 (ZAP-70) is a significant component in the development of a poor prognosis of the disease. This protein tyrosine kinase is an important mediator in the signaling process that is mediated by the T-cell receptor (TCR). Structurally, it is homologous to spleen tyrosine kinase (SYK), which plays a similar function in the signaling process that is mediated by the B-cell receptor (BCR) [245]. Recent evidence also indicates that ZAP-70, which is intrinsically expressed in tumor cells, may influence the crosstalk between malignant B cells and the immune microenvironment. This suggests that ZAP-70 plays a more complicated role in the pathogenesis of B cell malignancies than was previously thought [246]. Additional data showed that EGCG effectively suppressed ZAP-70 activity in CD3-activated T-cell leukemia, demonstrating a high binding affinity (Kd = 0.6207 μM). Furthermore, EGCG therapy dose-dependently suppressed the CD3-induced activation of activator protein-1 (AP-1) and IL-2. In particular, EGCG triggered caspase-mediated apoptosis in ZAP-70-expressing leukemia cells in a dose-dependent manner, whereas EGCG treatment was ineffective against P116 ZAP-70-deficient cells. The stability of the ZAP-70-EGCG complex may be aided by a network of intermolecular hydrogen bonds and hydrophobic contacts, as demonstrated by site-directed mutagenesis and molecular docking studies [247].
Wang et al. [248] employed a reverse docking methodology to identify new targets of EGCG and comprehend the pharmacological mechanism of action of this compound. The research revealed that EGCG may impact a total of 12 signaling pathways and 33 proteins. EGCG was found to exert binding properties with many proteins previously investigated in in vitro studies including proto-oncogene tyrosine-protein kinase FYN (FYN) [249], NOS2 [250], cyclin-dependent kinase 2 (CDK2) [251,252], tyrosine-protein kinase ABL1 (ABL1) [253], SYK [254], AKT1/2 [18,255], MAPK8, [256], interleukin-1 receptor-associated kinase 4 (IRAK4) [257], and apoptotic protease-activating factor 1 (APAF1) [258]. Furthermore, using network pharmacology, molecular dynamics simulation, and enzymatic assays, authors have investigated the inhibition of novel potential EGCG targets that were previously not investigated, including IKKβ, GTPase KRas (KRAS), high-affinity nerve growth factor receptor (NTRK1), and Wee1-like protein kinase (WEE1). EGCG has been shown to moderately suppress WEE1 activity while efficiently inhibiting IKKβ, KRAS, and NTRK1 in enzymatic assays [248]. The WEE1 kinase is essential for DNA repair at the G2-M cell-cycle checkpoint and is overexpressed in a wide range of malignancies, including, but not limited to, breast cancer, leukemias, melanomas, adult, and brain tumors [259]. In contrast, mutations in KRAS result in the permanent activation of this protein that acts as a molecular switch to consistently stimulate downstream signaling pathways controlling cell proliferation and survival [260,261]. On the other hand, NTRK1 is involved in the control of neural development and differentiation. In multiple tumor types, a constitutive activation of NTRK1 was observed. Tumor progression in both prostate carcinoma and breast cancer has been linked to an autocrine cycle involving NTRK1. Recently, a novel alternative splicing variant in neuroblastoma was characterized. Moreover, a significant proportion of papillary thyroid cancers harbor NTRK1 somatic rearrangements that generate chimeric oncogenes with constitutive tyrosine kinase activity, as reviewed by Pierotti and Greco [262]. Several other computational studies have investigated the effects of EGCG on other cancer-related factors including trypsin [263], proteasome [264], PTP1B phosphatase [265], glutathione S-transferase (GST P1-1) [266], or the epigenetic modulators DNMT [267], HDAC [57], and sirtuin-6 (SIRT6) [268], as recently reviewed [269,270].

4. EGCG in Cancer Prevention and Therapy?

4.1. Epidemiological Studies

The results of epidemiological studies on the relationship between drinking tea and the risk of developing cancer in human subjects have been inconclusive. According to Yuan et al., drinking black tea did not seem to reduce the incidence of cancer. After accounting for any confounding factors, a high consumption of green tea was consistently related to a lower risk of malignancies of the upper gastrointestinal tract and could help to prevent lung and liver cancer. No conclusive data suggest that tea consumption helps to prevent the onset of malignancies of the colorectum, pancreas, and urinary system or the development of glioma, lymphoma, or leukemia [271]. According to Zhang et al., consuming additional quantities of tea did not significantly reduce the chance of developing gliomas, brain tumors, stomach cancer, colon cancer, lung cancer, pancreatic cancer, liver cancer, breast cancer, prostate cancer, ovarian cancer, or bladder cancer. On the other hand, heavy tea consumption was connected to a decreased incidence of oral cancer. In addition, in western countries, an increase in tea consumption was associated with a lower bladder cancer risk [272]. While some studies in Asian populations indicated positive associations between green tea consumption and mortality [273], unclear results were obtained in epidemiological studies investigating the impact of green tea consumption on colorectal cancer [274], esophageal cancer [275,276], and gastric cancer [277] incidence in other populations.

4.2. Clinical Studies

The clinical trial results obtained in the last ten years of research indicate the beneficial influence of EGCG on the human body. Studies have focused on the influence of this phenolic compound on amyloidosis [278] atherosclerosis [279], cystic fibrosis [280,281], Barrett’s oesophagus [282], diabetes [283,284,285], Down’s syndrome [286,287,288,289], dystrophic epidermolysis bullosa [290], hypertension [291,292], multiple sclerosis [293,294,295,296,297,298,299,300], obesity [301,302,303,304,305,306,307,308], radiation-induced esophagitis [309,310,311,312], and ulcerative colitis [313].
Importantly, green tea extracts and tea consumption cannot be considered equal for several reasons. Green tea extracts are highly concentrated forms of the active compounds found in green tea, whereas tea consumption involves drinking brewed green tea, which typically contains lower concentrations of these compounds. Extracts are often made by concentrating the beneficial components of green tea, such as catechins, into a supplement form. Green tea extracts can be standardized to contain specific amounts of active compounds, ensuring consistency in dosage and potency. On the other hand, the composition of brewed tea can vary depending on factors such as brewing time, temperature, and the specific tea leaves used, making it challenging to determine the exact concentration of active compounds. Similarly, the extraction conditions may strongly impact the content of bioactive compounds in the final product. Furthermore, when drinking brewed green tea, the consumption of other components present in the tea leaves occurs. These include caffeine and various antioxidants, which can have additional effects on the body. Green tea extracts, on the other hand, may be formulated to remove or reduce certain components, allowing for the targeted supplementation of specific compounds [314,315,316,317].
Liu et al. investigated the effect of an intravesical irrigation of EGCG (dosage: 4 mg/kg body weight once per week with four treatments in total) in patients diagnosed with interstitial cystitis. The anticancer activities of this chemical against bladder cancer were attributed to the molecule’s ability to block the expression of purinergic receptors in urothelial cells as well as the release of ATP from those cells. In addition, EGCG inhibited the induction of iNOS and phosphorylated NF-κB, which resulted in the significant antioxidative and anti-inflammatory activity of the compounds [318]. An EGCG-enriched tea drink—double-brewed green tea (DBGT)—was the subject of a phase II study that was designed to evaluate the efficacy and safety of the beverage as a maintenance treatment for women who had advanced-stage serous or endometrioid ovarian cancer. After standard treatment for ovarian cancer, DBGT supplementation did not appear to be a promising maintenance option for patients who had reached an advanced stage of the disease [319].
Some data from preclinical studies, epidemiologic studies, and previous clinical trials suggest that catechins found in green tea may lessen the chance of developing prostate cancer. Although being well tolerated and accumulating in plasma over a year, a standardized, decaffeinated catechin mixture comprising 400 mg of EGCG per day did not diminish the risk of prostate cancer in men who had high-grade prostatic intraepithelial neoplasia (HGPIN) and/or atypical small acinar proliferation (ASAP) at the start of the study [320]. However, ASAP was never identified as a pre-neoplastic lesion in the context of prostate cancer [321,322]. When the same data from Kumar et al. [320] was analyzed without the ASAP lesions but with the HGPIN lesions, the reduction in PCa progression was approximately 50% according to a Kaplan-Meyer analysis. Furthermore, Bettuzzi et al. [323]. found a 90% reduction in prostate cancer in men with HGPIN who received green tea catechins for a year. Furthermore, the authors found no toxicity with green tea catechins in men with HGPIN who were randomly assigned to receive 600 mg of EGCG per day for a year. This chemoprevention strategy in high-risk patients would address a huge therapeutic and societal need, paving the way for a unique and effective clinical approach [323,324,325]. This is consistent with the findings of Parletti et al. [326], who conducted a single-outcome meta-analysis on the preventive role of green tea catechin extracts towards prostate cancer. The total number of patients in the study was 223; 114 and 109 were assigned to the catechin and placebo groups, respectively. There were nine incidences of prostate cancer in the catechin arm (7.9%) and twenty-four cases in the placebo arm (22%). Pooled analysis revealed that the catechin arm had a significantly lower cancer risk (risk-ratio = 0.41; 95% CI: 0.19–0.86; I2 = 0). These findings indicate that consuming concentrated green tea catechin formulations may provide significant protection to carriers of early neoplastic lesions of the prostate. The evidence is of average quality, and further large-scale research is needed to back up these preliminary conclusions [326].
Another study indicated that tea polyphenols and theaflavins (including EGCG) are bioavailable in the prostates of cancer patients, where they may play a role in prostate cancer prevention. For 5 days before radical prostatectomy, 20 men were randomized to consume 1.42 L of green tea, black tea, or a caffeine-matched soda control daily. Tea polyphenol levels were higher in the prostate samples from the individuals who consumed black tea and green tea versus the men who consumed the soda control. While tea polyphenols were not detected in serum, ex vivo LNCaP prostate cancer cell proliferation was reduced when cells were cultured in conditions containing patient sera collected after black tea and green tea consumption compared to baseline serum [327]. Additionally, men with clinically localized prostate cancer were given six cups of green tea or water daily for three to six weeks before undergoing radical prostatectomy to explore the metabolism and bioactivity of green tea polyphenols in human prostate tissue. Following a short-term green tea treatment, methylated and nonmethylated forms of EGCG were detected in prostate tissue, and the degree of methylation of EGCG might modulate its preventive effect on prostate cancer, probably involving genetic polymorphisms of catechol O-methyltransferase [328]. McLarty et al. investigated the short-term supplementation of polyphenon E (PPE) comprising 800 mg of (−)-epigallocatechin-3-gallate (EGCG) on the protein levels of hepatocyte growth factor (HGF), VEGF, IGF-I, IGF binding protein-3 (IGFBP-3), and prostate-specific antigen (PSA) in prostate cancer patients’ sera. The results demonstrated a significant reduction in PSA, HGF, and VEGF serum levels in males with prostate cancer after a short course of treatment with EGCG (PPE) with no increase in liver enzymes. These data suggest the possible role of PPE in the therapy or prevention of prostate cancer [329].
Fatty acid synthase (FAS) is essential for the development and maintenance of tumors with lipogenic characteristics because of its role as a master regulator of lipid metabolism. Evidence suggests that it can rewire tumor cells to have more energy flexibility, which is necessary for them to meet their high energy needs [330]. In contrast, the Ki-67 protein has served as a standard proliferation marker for human tumor cells for many years. Multiple molecular roles of this immense protein have been elucidated in recent research [331]. According to the findings of another clinical trial, supplementation with fish oil (1.9 g pf docosahexaenoic acid (DHA) + eicosapentanoic acid (EPA)/day) or EGCG (600 mg/day) for 90 days may not be adequate to generate biologically significant changes in FAS or Ki-67 protein levels in prostate tissue [332]. It was also found that a daily intake of a predefined, decaffeinated catechin mixture containing 200 mg of EGCG taken with food for one year produced an accumulation of the phenolic compound in plasma, and it was well tolerated and did not produce any treatment-related adverse effects in men who had baseline HGPIN or ASAP [325]. The above-mentioned studies investigated the effects of relatively short supplementation. In another study in men with baseline HGPIN or ASAP, daily intake of a uniform, decaffeinated catechin PPE composition comprising 200 mg/twice a day, EGCG that was taken with food for 1 year accumulated in plasma, was well tolerated, and did not elicit treatment-related side effects [325].
Also, the administration of GTE capsules (800 mg EGCG) seems to regulate biomarkers related to colorectal cancer pathogenesis, especially genes associated with selenoproteins, WNT signaling (β-catenin), inflammation (NF-κB), and methylation (DNMT1) [333]. Based on the synergistic effects seen in preclinical studies between green tea PPE and EGFR-tyrosine kinase inhibitors [334], a 6-month phase IB study of PPE (200 mg three times a day) and erlotinib (dose escalation of erlotinib (50, 75, and 100 mg daily)) was conducted in patients with head and neck cancer who had advanced premalignant lesions (APL) of the oral cavity and larynx. Patients diagnosed with APLs of the head and neck who received treatment with the combination of PPE and erlotinib were reported to have a prolonged cancer-free survival. The regimen was well tolerated by the patients, indicating that this combination should be investigated further for its potential to be used as a chemoprevention of recurrent primary tumors in early-stage head and neck cancer patients [335]. Several other clinical studies are planned to investigate EGCG’s anticancer activity and safety in cancer patients, as discussed in Table 1.

4.3. Bioavailability of EGCG

One of the obstacles to the clinical use of EGCG is its poor oral bioavailability that is related to its low intestinal absorption and short retention time or lack of stability. Overcoming these challenges related to its physicochemical properties will realistically impact the conversion of EGCG from basic studies to widespread use in clinical settings [336]. Improvements in EGCG’s pharmacokinetics and pharmacological properties, and thus its therapeutic possibilities, have been achieved by a variety of approaches, including structural alterations of the compound, the use of nano-carriers including carbohydrate-based carriers (chitosan-tripolyphosphate nanoparticles [337], γ-cyclodextrin (γ-CD), or carriers coated with hydroxypropylmethyl cellulose phthalate [338,339]), lipid carriers (liposomes) [340,341,342], protein carriers [343,344,345,346], and other effective drug delivery methods, as recently discussed in excellent reviews [347,348]. These approaches increase the bioavailability of EGCG and prolong the dosing interval [349].
EGCG has been investigated in doses ranging from 150 to 400 mg per day, with oral administration being the most extensively utilized delivery modality in animal models as well as clinical studies [350,351,352]. In contrast, it was reported that doses of 140 mg to 1000 mg per day may exhibit hepatotoxic effects [353]. However, these finding seems ambiguous [354]. Lambert et al. investigated the hepatotoxicity of large doses of EGCG in male CF-1 mice. A single dose of EGCG (1500 mg/kg, i.g.) elevated plasma ALT levels by 138-fold and lowered mice survival by 85%. Once-daily EGCG administration increased the level of hepatotoxic reaction. Following two once-daily dosages of 750 mg/kg EGCG, plasma ALT levels were elevated 184-fold. Following EGCG therapy, moderate-to-severe hepatic necrosis was found; elevated hepatic lipid peroxidation (5-fold increase), plasma 8-isoprostane (9.5-fold increase), and elevated hepatic metallothionein and gamma-histone 2AX protein expression were all related with EGCG hepatotoxicity. EGCG also elevated interleukin-6 and monocyte chemoattractant protein-1 levels in the blood [355]. A series of tests were conducted to determine the toxicity of purified green tea extracts with high concentrations of EGCG to determine the safety of Teavigo, a high-concentration EGCG extract. Topical administration of EGCG extract had minor dermal irritation properties in rats and guinea pigs but not rabbits. In the guinea pig maximization test, topical EGCG formulations generated modest cutaneous irritation. An eye irritation test on rabbits revealed a strong enough response to rule out further testing in this assay. An oral dose of a 2000 mg EGCG preparation/kg killed rats, whereas a dose of 200 mg EGCG/kg had no adverse effects. At doses of up to 500 mg/kg/day, food feeding of an EGCG preparation to rats for 13 weeks was not hazardous. Similarly, when a 500 mg EGCG preparation/kg/day was delivered in divided doses to pre-fed dogs, no adverse effects were seen. When administered to starved dogs as a single bolus dosage, this dose induced morbidity, although this model was regarded as an impractical analogy to human consumption. These trials established a non-observed adverse impact limit of 500 mg of EGCG preparation/kg/day [356]. Similarly, following oral dosage using capsules, a standardized green tea extract was tested for exposure and toxicity in beagles. EGCG was the major component of the preparation, accounting for 56–72% of the material. To take advantage of the reported enhanced catechin bioavailability with fasting, a 9-month chronic trial (0, 200, 500, and 1000 mg/kg/day) was conducted in fasted dogs. Extensive morbidity, mortality, and pathology of several major organs resulted in its premature conclusion at 6.5 months, preventing the elucidation of the toxicity reasons. A 13-week follow-up study looked at the extract’s toxicity and exposure. The toxicities were generally less severe than in the chronic study at the same time. Dosing in a fed state resulted in significantly lower and less varied exposure than in fasted settings. Toxicity was less common and less severe with reduced exposure, but the small sample size and high variability precluded a definitive conclusion from being reached [357]. Micronuclei production in bone marrow cells was not induced by oral administration of 500, 1000, or 2000 mg EGCG/kg to mice. Similarly, exposing them to 400, 800, or 1200 mg EGCG/kg/day for 10 days did not induce the bone marrow cell micronuclei while generating plasma EGCG concentrations comparable to those seen in human investigations. The intravenous administration of 10, 25, and 50 mg EGCG/kg/day to rats resulted in significantly greater plasma concentrations and the absence of EGCG’s genotoxic effects. According to the findings of these investigations, Teavigo (EGCG) is not genotoxic [358]. Feeding pregnant rats meals supplemented with 1400, 4200, or 14,000 ppm throughout organogenesis was not hazardous to dams or fetuses in a major teratogenicity investigation. A two-generation study in rats administered with EGCG preparations at 1200, 3600, or 12,000 ppm found no negative effects on reproduction or fertility. The greatest dose lowered the progeny growth rate and caused a modest increase in pup loss. At 3600 ppm, there was also a growth effect in pups, but this was only observed in the second generation. The lowest dose was determined to be the overall non-observed adverse effect level (NOAEL). The NOAEL was similar to the 200 mg/kg/day EGCG preparation because the dams consumed twice as much feed during the critical lactation period [359]. A review of adverse event (AE) data from 159 human intervention studies found that only a few kinds of concentrated catechin-rich green tea preparations caused hepatic AEs in a dose-dependent manner when consumed in large bolus doses but not when utilized as brewed tea or extracts in beverages or as part of food. Toxico- and pharmacokinetic research indicates that the internal dose of catechins is a critical variable in the prevalence and severity of hepatotoxicity. Toxological and human safety data for tea preparations consumed as a solid bolus dose were used to calculate a safe intake level of 338 mg of EGCG/day for adults. Based on human AE data, an observed safe level (OSL) of 704 mg of EGCG/day could be proposed for tea preparations in a drinkable form [360].
Nevertheless, it is crucial to understand the fate of EGCG following this route of administration. The stomach’s acid (pH < 3) ensures the stabilization of EGCG following oral ingestion. The bioactivities and health effects of EGCG rely on its absorption and metabolism in the intestine. The small intestine absorbs some EGCG; however, a major fraction of EGCG passes from the small to the large intestine, where enterocytes and microbes convert it into up to eleven catechin ring-fission products [350]. EGCG is transported across the epithelium via passive diffusion, which explains the poor absorption of this catechin. Moreover, the compound may act as a substrate for multiple efflux transporters including multidrug-resistance-associated protein (MRP) efflux pumps [361] and P-glycoprotein (P-gp) [362], which limit the cellular uptake of the phenolic compound. This topic was previously reviewed [348,363].
As evidenced by clinical studies, the significant degree of variability in the pharmacokinetic parameters of EGCG might be explained by genetic variations in the genes encoding the drug transporters MRP2 and organic anion transporter SLC21A6 (OATP1B1). It is possible that this interindividual heterogeneity could account for the quantities of green tea polyphenols in human plasma or their potential to induce beneficial effects [364] and their potential to trigger pharmacokinetic interaction with other drugs [365,366,367,368].
EGCG is susceptible to degradation or oxidation, thus conferring low EGCG concentrations in peripheral blood. Thereby, human plasma contains free and conjugated intestinal-microbiota-produced EGCG ring-fission metabolites [350]. Phase II enzymes, including UDP-glucuronosyltransferases (UGTs), sulphotransferases (SULTs), and catechol-O-methyltransferase (COMT), are responsible for the extensive metabolism of a portion of the ingested tea catechins before and after absorption, which occurs primarily in the small intestine and the liver [369]. Following ingestion, EGCG reaches a maximal plasma concentration after 90 min and is removed from the organism in the following 24 h [350]. Another study suggested that the elimination half-life of this phenolic compound is between 3.4 ± 0.3 h [370]. EGCG has been the subject of several more studies in both animal and human subjects aimed at assessing the bioavailability of this compound, some of which have shown conflicting findings [351,371]. An overview of EGCG’s metabolism is presented in Figure 5.
There is also no general agreement on the optimal level of EGCG that should be present in the human serum to achieve the greatest possible therapeutic benefits [350]. Several in vitro investigations have found that concentrations between 1–200 µM of EGCG are optimal for achieving the desired biological effects. Several pharmacokinetics investigations have shown that only a tiny percentage of unmodified EGCG reaches the bloodstream or tissues, making it challenging to achieve this effective concentration in vivo or in clinical research [347].
Recent investigations suggest the superior bioavailability of Teavigo® (Healthy Origins, Pittsburgh, PA, USA) (capsules made from a refined green tea extract that have an EGCG level of around 94% with a greater solubility and stability compared to non-formulated EGCG extracts) compared to other available formulations such as FontUp® (Grand Fontaine Laboratories, Barcelona, Spain) (Teavigo® formulation combined with fats, carbohydrates, proteins, vitamins, and minerals). Moreover, the greatest bioavailability was achieved when the formulated extract was administered following fasting conditions (overnight) [350]. This is consistent with the earlier reports of EGCG’s bioavailability [375]. In contrast, taking EGCG in combination with dietary supplements (FontUp®) increased the molecule’s stability in the body [350]. The bioavailability and stability of EGCG have also been investigated in combination with other dietary components such as quercetin [376] or vitamin C [377]. Lazzeroni et al. investigated the effects of a lecithin formulation of a caffeine-free green tea catechin extract (Greenselect Phytosome (GSP)) on EGCG tissue distribution and its effect on cell proliferation and circulating biomarkers in breast cancer patients. Oral GSP was found to increase the bioavailability of EGCG [378].
Interesting clinical studies were performed to investigate whether or not the combined effects of drinking green tea and bathing in hot springs can lead to an increase in the yields of catechins that are absorbed into the body. The study compared two different interventions: drinking green tea while also taking a bath in a hot spring and drinking green tea on its own. Four healthy individuals had their plasma levels of EGCG measured. It was discovered that the combined practice of drinking green tea and taking a bath in hot springs resulted in a higher concentration of EGCG in the plasma [379].

4.4. Other Concerns

Despite the fact that the bioavailability of EGCG can strongly impact its conversion for clinical use, several other concerns need to be addressed:
  • Standardization: Clinical research outcomes are difficult to evaluate and duplicate because of variations in the quality and purity of the EGCG preparations used.
  • Safety concerns: Even though EGCG is usually believed to be safe, it might lead to liver damage in high doses, and there is a possibility that it could interact with some drugs. Additional research is required to completely understand the safety profile of this compound.
  • Lack of clinical data: Although there is some evidence from preclinical research to show that EGCG may have a cancer-preventative effect, there is very little evidence from clinical studies that support its use in humans. More clinical tests that are carefully designed are required to assess the effectiveness and safety of this flavonoid.

5. Conclusions

EGCG, the main bioactive constituent of green tea, is a promising natural anti-cancer agent with a wide spectrum of health benefits. This phytochemical was shown to inhibit the growth, proliferation, and metastasis and trigger the apoptosis of multiple cancer cells derived from different tissues and organs in both in vitro and in vivo settings. The understanding of the molecular mechanisms of the anticancer activity of this compound revealed the versatile capability of this catechin to influence multiple pivotal signaling pathways inside the cell including, but not limited to, EGFR, JAK/STAT, PI3K/AKT/mTOR, and MAPK. EGCG has been found to modulate epigenetic mechanisms such as DNA methylation and histone modifications by targeting epigenetic modulators such as DNMTs, HATs, and HDACs (Figure 6).
Moreover, new emerging targets of EGCG have been identified. For example, EGCG acts as an ATP-competitive inhibitor of GRP78, a chaperone protein that promotes chemoresistance and inhibits apoptosis in cancer cells. EGCG has been found to suppress GRP78 activity and hinder its expression, particularly in glioblastomas. EGCG has been identified as a compound with a significant capability of HK2 inhibition. HK2 plays a role in the metabolic shift towards aerobic glycolysis in cancer cells. EGCG effectively suppresses ZAP-70 activity, a protein tyrosine kinase involved in the signaling process mediated by T-cell receptors. Through reverse docking methodology and network pharmacology, EGCG has been found to impact multiple signaling pathways and proteins. These include FYN, NOS2, CDK2, ABL1, SYK, AKT1/2, MAPK8, IRAK4, and APAF1, which have been previously investigated in in vitro studies. Additionally, novel targets including IKKβ, KRAS, NTRK1, and WEE1 have been identified. EGCG has shown a moderate suppression of WEE1 activity and an efficient inhibition of IKKβ, KRAS, and NTRK1 in enzymatic assays. Future studies may also elucidate new mechanisms through which EGCG may suppress the development of tumors or restrict their growth.
Clinical trial results from the last ten years suggest that EGCG may have beneficial effects on several health conditions. These conditions include amyloidosis, atherosclerosis, cystic fibrosis, Barrett’s esophagus, diabetes, Down’s syndrome, dystrophic epidermolysis bullosa, hypertension, multiple sclerosis, obesity, radiation-induced esophagitis, and ulcerative colitis. EGCG has shown anticancer activities against bladder cancer. It blocks the expression of purinergic receptors in urothelial cells and inhibits the release of ATP from those cells. Additionally, EGCG has demonstrated anti-inflammatory and antioxidant activity, as it inhibits the induction of iNOS and phosphorylated NF-κB. Some preclinical studies, epidemiologic studies, and previous clinical trials suggest that catechins, including EGCG, found in green tea may reduce the risk of developing prostate cancer. However, a clinical trial with a decaffeinated catechin mixture containing 400 mg of EGCG per day did not diminish the risk of prostate cancer in men with HGPIN and/or ASAP. The administration of GTE capsules containing 800 mg of EGCG seems to regulate biomarkers related to colorectal cancer pathogenesis, including selenoproteins, WNT signaling (β-catenin), inflammation (NF-κB), and methylation (DNMT1). A phase IB study conducted in patients with APL of the oral cavity and larynx showed that a combination of green tea polyphenon E (PPE) and erlotinib prolonged cancer-free survival. This combination was well tolerated, suggesting its potential as a chemopreventive measure for recurrent primary tumors in early-stage head and neck cancer patients.
Despite the initial enthusiasm for EGCG use in cancer treatment, conflicting epidemiological data were obtained considering EGCG consumption and the decreased risk of many cancers. No clear evidence suggests that tea consumption reduces the overall risk of developing cancer.
Furthermore, the clinical use of EGCG is hindered by its poor oral bioavailability, low intestinal absorption, and short retention time. Overcoming these challenges is necessary for its widespread use in clinical settings. Various approaches, such as structural alterations, nano-carriers, lipid carriers, and protein carriers, have been employed to improve the pharmacokinetics and pharmacological properties of EGCG, increasing its bioavailability and dosing interval. Moreover, the effects of nutrients, timing, and fasting were investigated to increase the intestinal absorption and bioavailability of EGCG. Additionally, the wide-spread access to EGCG and EGCG sources increases the toxicities in humans when its use is abused. Despite the fact that EGCG is usually believed to be safe, it might lead to liver damage in high doses, and there is a possibility that it could interact with some drugs. Additional research is required to completely understand the safety profile of this compound and establish an optimal dosage that will provide the most health benefits and decrease the chance of developing toxicity effects.
Although there is some evidence from preclinical research to show that EGCG may have a cancer-preventative effect, there is very little evidence from clinical studies that supports its use in humans. More clinical tests that are carefully designed are required to assess the effectiveness and safety of this compound. Clinical research outcomes are often difficult to evaluate and duplicate because of variations in the quality and purity of EGCG preparations used. Therefore, standardization is required.

Author Contributions

Conceptualization, M.K., M.A., N.A. and S.R.; supervision, M.K., M.A., N.A., S.R., R.S., I.C., E.B.Y., A.D., E.Z. and R.K.; visualization, M.K., M.A., N.A., S.R. and P.G.; writing—original draft, M.K., M.A., N.A., S.R. and P.G.; writing—review and editing, M.K., M.A., N.A., S.R., P.G., R.S., I.C., E.B.Y., A.D., E.Z. and R.K. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported via funding from Prince sattam bin Abdulaziz University, project number PSAU/2023/R/1444, and was founded by the University of Lodz: order no. 18 of the Rector of the University of Lodz of 23 October 2019.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created.

Acknowledgments

The authors acknowledge the support via funding from Prince sattam bin Abdulaziz University, project number PSAU/2023/R/1444, and was founded by the University of Lodz: order no. 18 of the Rector of the University of Lodz of 23 October 2019.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

4E-BP1Eukaryotic-translation-initiation-factor-4E-binding protein 1
67LR67-kD laminin receptor
ABL1Tyrosine-protein kinase ABL1
AIFApoptosis-inducing factor
AKTProtein kinase B
AP-1Activator protein-1
APAF1Apoptotic-protease-activating factor 1
APLAcute promyelocytic leukemia
ARAndrogen receptor
ASAPAtypical small acinar proliferation
ASMAcid sphingomyelinase
BAXBcl-2-associated X protein
BCL-2B-cell lymphoma 2
BCL-xLB-cell lymphoma-extra large
BCRB-cell receptor
CAFsCancer-associated fibroblasts
CD3Cluster of differentiation 3
CDH1Cadherin 1
CDK2Cyclin-dependent kinase 2
CDK4/6Cyclin-dependent kinases 4/6
c-IAP1/2Cellular inhibitor of apoptosis protein 1/2
CLLChronic lymphocytic leukemia
COMTCatechol-O-methyltransferase
COX-1Cyclooxygenase-1
COX-2Cyclooxygenase-2
COXsCyclooxygenases
CpGCytosine-phosphate-guanine
DAPK1Death-associated protein kinase 1
DBGTDouble-brewed green tea
DNA-PKDNA-dependent protein kinase
DNMTDNA methyltransferase
DR5Death receptor 5
XIAPE3 ubiquitin-protein-ligase-X-linked inhibitor of apoptosis protein
ECEpicatechin
ECGEpicatechin-3-gallate
ECMExtracellular matrix
EGCEpigallocatechin
EGCGEpigallocatechin-3-gallate
EGFEpidermal growth factor
EGFREpidermal growth factor receptor
EREndoplasmic reticulum
ERK1/2Extracellular-signal-regulated kinases 1 and 2
ETCElectron transport chain
EZH2Enhancer of zeste homolog 2
FAKFocal adhesion kinase
FASFatty acid synthase
FLIPCellular FLICE-like inhibitory protein
GADD153Growth arrest and DNA-damage-inducible gene 153
GLI1GLI-Kruppel family member 1
GLUTsGlucose transporters
GRP78Glucose-regulated protein 78
GST P1-1Glutathione S-transferase P1-1
GTEGreen tea extract
HDACHistone deacetylase
HER-2Human epidermal growth factor receptor 2
HGFHepatocyte growth factor
HGPINHigh-grade prostatic intraepithelial neoplasia
HhHedgehog
HIF-1αHypoxia-inducible factor 1-alpha
HK2Hexokinase 2
HUVECsHuman umbilical vein endothelial cells
ICAM-1Intercellular adhesion molecule-1
IFNsInterferons
IGF-1Insulin-like growth factor 1
IGFBP-3IGF-binding protein-3
IGF-1RInsulin-like growth factor-1 receptor
IKKβInhibitor of nuclear factor kappa-B kinase subunit beta
IL-1βInterleukin 1β
IL-2Interleukin 2
IL-4Interleukin 4
IL-6Interleukin 6
IL-8Interleukin 8
iNOSInducible nitric oxide synthase
IRAK4Interleukin-1-receptor-associated kinase 4
IRF1Interferon regulatory factor 1
JAKTyrosine-protein kinase JAK
JNKc-Jun N-terminal kinase
Ki-67Antigen Ki-67
KRASGTPase KRas
MAPK8Mitogen-activated protein kinase 8
MCP-1Monocyte chemoattractant protein-1
MDMolecular dynamics
MMPsMatrix metalloproteinases
MRPMultidrug-resistance-associated protein
mTORSerine/threonine-protein kinase mTOR
NADPHNicotinamide adenine dinucleotide phosphate
NF-KbNuclear factor kappa-light-chain-enhancer of activated B cells
NONitric oxide
NOSNitric oxide synthase
NOTCHNeurogenic locus notch homolog protein
NPsNatural products
NSCLCNon-small-cell lung cancer
NTRK1High-affinity nerve growth factor receptor
OATP1B1Organic anion-transporting polypeptide 1B1
P300/CBPCREB-binding protein/p300
PARPPoly (ADP-ribose) polymerase
PcGPolycomb group
PDK1Phosphoinositide-dependent kinase-1
PD-L1/2Programmed death-ligand 1/2
PFKPhosphofructokinase
PFKPPlatelet-type phosphofructokinase
PGE2Prostaglandin E2
P-gpP-glycoprotein
PI3KPhosphatidylinositol-3-kinase
PIP2Phosphatidylinositol 4,5-bisphosphate
PIP3Phosphatidylinositol 3,4,5-triphosphate
PKPyruvate kinase
PKCProtein kinase C
PKM2Pyruvate kinase M2
PP2AProtein phosphatase 2A
PPEGreen tea polyphenon E
PRC2Polycomb-repressive complex 2
PSAProstate-specific antigen
PTENPhosphatidylinositol-3,4,5-trisphosphate 3-phosphatase and dual-specificity protein phosphatase
RARβRetinoic acid receptor beta
RAS/MAPKRAS/mitogen-activated protein kinase
ROSReactive oxygen species
S6KRibosomal protein kinase S6
SAMS-adenosylmethionine
SHHSonic hedgehog
SIRT6Sirtuin-6
SLC21A6Organic anion transporter
SMOSmoothened
SPRSurface plasmon resonance
STATSignal transducer and activator of transcription
SULTsSulphotransferases
SYKSpleen tyrosine kinase
TAMsTumor-associated macrophages
TCRT-cell receptor
TIMP-3Tissue inhibitor of matrix metalloproteinase-3
TLR4Toll-like receptor 4
TMETumor microenvironment
TNFR-1/2TNF receptor 1/2
TNF-αTumor necrosis factor α
TP53Tumor protein p53
UGTsUDP-glucuronosyltransferases
VEGFVascular endothelial growth factor
VEGFRVascular endothelial growth factor receptor
WEE1Wee1-like protein kinase
ZAP-70Zeta-chain-associated protein kinase-70
γ-CDγ-cyclodextrin

References

  1. Hayat, K.; Iqbal, H.; Malik, U.; Bilal, U.; Mushtaq, S. Tea and Its Consumption: Benefits and Risks. Crit. Rev. Food Sci. Nutr. 2015, 55, 939–954. [Google Scholar] [CrossRef]
  2. Khan, N.; Mukhtar, H. Tea and Health: Studies in Humans. Curr. Pharm. Des. 2013, 19, 6141–6147. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Chakrawarti, L.; Agrawal, R.; Dang, S.; Gupta, S.; Gabrani, R. Therapeutic Effects of EGCG: A Patent Review. Expert Opin. Ther. Pat. 2016, 26, 907–916. [Google Scholar] [CrossRef]
  4. Alam, M.; Ali, S.; Ashraf, G.M.; Bilgrami, A.L.; Yadav, D.K.; Hassan, M.I. Epigallocatechin 3-Gallate: From Green Tea to Cancer Therapeutics. Food Chem. 2022, 379, 132135. [Google Scholar] [CrossRef] [PubMed]
  5. Lambert, J.D.; Yang, C.S. Cancer Chemopreventive Activity and Bioavailability of Tea and Tea Polyphenols. Mutat. Res. 2003, 523–524, 201–208. [Google Scholar] [CrossRef]
  6. Yang, C.S.; Lambert, J.D.; Ju, J.; Lu, G.; Sang, S. Tea and Cancer Prevention: Molecular Mechanisms and Human Relevance. Toxicol. Appl. Pharmacol. 2007, 224, 265–273. [Google Scholar] [CrossRef] [Green Version]
  7. Mukhtar, H.; Ahmad, N. Tea Polyphenols: Prevention of Cancer and Optimizing Health. Am. J. Clin. Nutr. 2000, 71, 1698S–1702S, discussion 1703S–1704S. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Ohishi, T.; Goto, S.; Monira, P.; Isemura, M.; Nakamura, Y. Anti-Inflammatory Action of Green Tea. Anti-Inflamm. Anti-Allergy Agents Med. Chem. 2016, 15, 74–90. [Google Scholar] [CrossRef] [PubMed]
  9. Oz, H.S. Chronic Inflammatory Diseases and Green Tea Polyphenols. Nutrients 2017, 9, 561. [Google Scholar] [CrossRef] [Green Version]
  10. Sah, J.F.; Balasubramanian, S.; Eckert, R.L.; Rorke, E.A. Epigallocatechin-3-Gallate Inhibits Epidermal Growth Factor Receptor Signaling Pathway. Evidence for Direct Inhibition of ERK1/2 and AKT Kinases. J. Biol. Chem. 2004, 279, 12755–12762. [Google Scholar] [CrossRef] [Green Version]
  11. Minnelli, C.; Cianfruglia, L.; Laudadio, E.; Mobbili, G.; Galeazzi, R.; Armeni, T. Effect of Epigallocatechin-3-Gallate on EGFR Signaling and Migration in Non-Small Cell Lung Cancer. Int. J. Mol. Sci. 2021, 22, 11833. [Google Scholar] [CrossRef] [PubMed]
  12. Shimizu, M.; Deguchi, A.; Lim, J.T.E.; Moriwaki, H.; Kopelovich, L.; Weinstein, I.B. (−)-Epigallocatechin Gallate and Polyphenon E Inhibit Growth and Activation of the Epidermal Growth Factor Receptor and Human Epidermal Growth Factor Receptor-2 Signaling Pathways in Human Colon Cancer Cells. Clin. Cancer Res. 2005, 11, 2735–2746. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Masuda, M.; Suzui, M.; Weinstein, I.B. Effects of Epigallocatechin-3-Gallate on Growth, Epidermal Growth Factor Receptor Signaling Pathways, Gene Expression, and Chemosensitivity in Human Head and Neck Squamous Cell Carcinoma Cell Lines. Clin. Cancer Res. 2001, 7, 4220–4229. [Google Scholar] [PubMed]
  14. Hamed, F.N.; McDonagh, A.J.G.; Almaghrabi, S.; Bakri, Y.; Messenger, A.G.; Tazi-Ahnini, R. Epigallocatechin-3 Gallate Inhibits STAT-1/JAK2/IRF-1/HLA-DR/HLA-B and Reduces CD8 MKG2D Lymphocytes of Alopecia Areata Patients. Int. J. Environ. Res. Public Health 2018, 15, 2882. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Ravindran Menon, D.; Li, Y.; Yamauchi, T.; Osborne, D.G.; Vaddi, P.K.; Wempe, M.F.; Zhai, Z.; Fujita, M. EGCG Inhibits Tumor Growth in Melanoma by Targeting JAK-STAT Signaling and Its Downstream PD-L1/PD-L2-PD1 Axis in Tumors and Enhancing Cytotoxic T-Cell Responses. Pharmaceuticals 2021, 14, 1081. [Google Scholar] [CrossRef]
  16. Al-Maghrebi, M.; Alnajem, A.S.; Esmaeil, A. Epigallocatechin-3-Gallate Modulates Germ Cell Apoptosis through the SAFE/Nrf2 Signaling Pathway. Naunyn-Schmiedeberg’s Arch. Pharmacol. 2020, 393, 663–671. [Google Scholar] [CrossRef]
  17. Barberino, R.S.; Santos, J.M.S.; Lins, T.L.B.G.; Menezes, V.G.; Monte, A.P.O.; Gouveia, B.B.; Palheta, R.C.; Matos, M.H.T. Epigallocatechin-3-Gallate (EGCG) Reduces Apoptosis of Preantral Follicles through the Phosphatidylinositol-3-Kinase/Protein Kinase B (PI3K/AKT) Signaling Pathway after in Vitro Culture of Sheep Ovarian Tissue. Theriogenology 2020, 155, 25–32. [Google Scholar] [CrossRef]
  18. Van Aller, G.S.; Carson, J.D.; Tang, W.; Peng, H.; Zhao, L.; Copeland, R.A.; Tummino, P.J.; Luo, L. Epigallocatechin Gallate (EGCG), a Major Component of Green Tea, Is a Dual Phosphoinositide-3-Kinase/MTOR Inhibitor. Biochem. Biophys. Res. Commun. 2011, 406, 194–199. [Google Scholar] [CrossRef]
  19. Tang, G.; Zhang, Z.; Qian, H.; Chen, J.; Wang, Y.; Chen, X.; Chen, B.; Chen, Y. (−)-Epigallocatechin-3-Gallate Inhibits Osteosarcoma Cell Invasiveness by Inhibiting the MEK/ERK Signaling Pathway in Human Osteosarcoma Cells. J. Environ. Pathol. Toxicol. Oncol. 2015, 34, 85–93. [Google Scholar] [CrossRef]
  20. Wei, R.; Wirkus, J.; Yang, Z.; Machuca, J.; Esparza, Y.; Mackenzie, G.G. EGCG Sensitizes Chemotherapeutic-Induced Cytotoxicity by Targeting the ERK Pathway in Multiple Cancer Cell Lines. Arch. Biochem. Biophys. 2020, 692, 108546. [Google Scholar] [CrossRef]
  21. Steed, K.L.; Jordan, H.R.; Tollefsbol, T.O. SAHA and EGCG Promote Apoptosis in Triple-Negative Breast Cancer Cells, Possibly Through the Modulation of CIAP2. Anticancer Res. 2020, 40, 9–26. [Google Scholar] [CrossRef]
  22. 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]
  23. Liu, L.; Ju, Y.; Wang, J.; Zhou, R. Epigallocatechin-3-Gallate Promotes Apoptosis and Reversal of Multidrug Resistance in Esophageal Cancer Cells. Pathol. Res. Pract. 2017, 213, 1242–1250. [Google Scholar] [CrossRef] [PubMed]
  24. Iwasaki, R.; Ito, K.; Ishida, T.; Hamanoue, M.; Adachi, S.; Watanabe, T.; Sato, Y. Catechin, Green Tea Component, Causes Caspase-Independent Necrosis-like Cell Death in Chronic Myelogenous Leukemia. Cancer Sci. 2009, 100, 349–356. [Google Scholar] [CrossRef]
  25. Alam, M.; Ali, S.; Ahmed, S.; Elasbali, A.M.; Adnan, M.; Islam, A.; Hassan, M.I.; Yadav, D.K. Therapeutic Potential of Ursolic Acid in Cancer and Diabetic Neuropathy Diseases. Int. J. Mol. Sci. 2021, 22, 12162. [Google Scholar] [CrossRef] [PubMed]
  26. Alam, M.; Ahmed, S.; Elasbali, A.M.; Adnan, M.; Alam, S.; Hassan, M.I.; Pasupuleti, V.R. Therapeutic Implications of Caffeic Acid in Cancer and Neurological Diseases. Front. Oncol. 2022, 12, 860508. [Google Scholar] [CrossRef] [PubMed]
  27. Butler, M.S.; Robertson, A.A.B.; Cooper, M.A. Natural Product and Natural Product Derived Drugs in Clinical Trials. Nat. Prod. Rep. 2014, 31, 1612–1661. [Google Scholar] [CrossRef] [PubMed]
  28. Gordaliza, M. Natural Products as Leads to Anticancer Drugs. Clin. Transl. Oncol. 2007, 9, 767–776. [Google Scholar] [CrossRef] [PubMed]
  29. Solowey, E.; Lichtenstein, M.; Sallon, S.; Paavilainen, H.; Solowey, E.; Lorberboum-Galski, H. Evaluating Medicinal Plants for Anticancer Activity. Sci. World J. 2014, 2014, 721402. [Google Scholar] [CrossRef] [Green Version]
  30. Kebebe, D.; Liu, Y.; Wu, Y.; Vilakhamxay, M.; Liu, Z.; Li, J. Tumor-Targeting Delivery of Herb-Based Drugs with Cell-Penetrating/Tumor-Targeting Peptide-Modified Nanocarriers. Int. J. Nanomed. 2018, 13, 1425–1442. [Google Scholar] [CrossRef] [Green Version]
  31. Hossain, R.; Jain, D.; Khan, R.A.; Islam, M.T.; Mubarak, M.S.; Mohammad Saikat, A.S. Natural-Derived Molecules as a Potential Adjuvant in Chemotherapy: Normal Cell Protectors and Cancer Cell Sensitizers. Anti-Cancer Agents Med. Chem. 2022, 22, 836–850. [Google Scholar] [CrossRef]
  32. Diederich, M. Natural Compound Inducers of Immunogenic Cell Death. Arch. Pharm. Res. 2019, 42, 629–645. [Google Scholar] [CrossRef]
  33. Luo, K.-W.; Chen, W.; Lung, W.-Y.; Wei, X.-Y.; Cheng, B.-H.; Cai, Z.-M.; Huang, W.-R. EGCG Inhibited Bladder Cancer SW780 Cell Proliferation and Migration Both in Vitro and in Vivo via Down-Regulation of NF-ΚB and MMP-9. J. Nutr. Biochem. 2017, 41, 56–64. [Google Scholar] [CrossRef]
  34. Luo, K.-W.; Zhu, X.-H.; Zhao, T.; Zhong, J.; Gao, H.-C.; Luo, X.-L.; Huang, W.-R. EGCG Enhanced the Anti-Tumor Effect of Doxorubicine in Bladder Cancer via NF-ΚB/MDM2/P53 Pathway. Front. Cell Dev. Biol. 2020, 8, 606123. [Google Scholar] [CrossRef] [PubMed]
  35. Yin, Z.; Li, J.; Kang, L.; Liu, X.; Luo, J.; Zhang, L.; Li, Y.; Cai, J. Epigallocatechin-3-Gallate Induces Autophagy-Related Apoptosis Associated with LC3B II and Beclin Expression of Bladder Cancer Cells. J. Food Biochem. 2021, 45, e13758. [Google Scholar] [CrossRef] [PubMed]
  36. Lee, H.-Y.; Chen, Y.-J.; Chang, W.-A.; Li, W.-M.; Ke, H.-L.; Wu, W.-J.; Kuo, P.-L. Effects of Epigallocatechin Gallate (EGCG) on Urinary Bladder Urothelial Carcinoma-Next-Generation Sequencing and Bioinformatics Approaches. Medicina 2019, 55, 768. [Google Scholar] [CrossRef] [Green Version]
  37. Philips, B.J.; Coyle, C.H.; Morrisroe, S.N.; Chancellor, M.B.; Yoshimura, N. Induction of Apoptosis in Human Bladder Cancer Cells by Green Tea Catechins. Biomed. Res. 2009, 30, 207–215. [Google Scholar] [CrossRef] [Green Version]
  38. Luo, K.-W.; Lung, W.-Y.; Chun-Xie; 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] [PubMed] [Green Version]
  39. Hsieh, D.-S.; Wang, H.; Tan, S.-W.; Huang, Y.-H.; Tsai, C.-Y.; Yeh, M.-K.; Wu, C.-J. The Treatment of Bladder Cancer in a Mouse Model by Epigallocatechin-3-Gallate-Gold Nanoparticles. Biomaterials 2011, 32, 7633–7640. [Google Scholar] [CrossRef]
  40. Qin, J.; Xie, L.-P.; Zheng, X.-Y.; Wang, Y.-B.; Bai, Y.; Shen, H.-F.; Li, L.-C.; Dahiya, R. A Component of Green Tea, (−)-Epigallocatechin-3-Gallate, Promotes Apoptosis in T24 Human Bladder Cancer Cells via Modulation of the PI3K/Akt Pathway and Bcl-2 Family Proteins. Biochem. Biophys. Res. Commun. 2007, 354, 852–857. [Google Scholar] [CrossRef]
  41. Sun, X.; Song, J.; Li, E.; Geng, H.; Li, Y.; Yu, D.; Zhong, C. (-)-Epigallocatechin-3-gallate Inhibits Bladder Cancer Stem Cells via Suppression of Sonic Hedgehog Pathway. Oncol. Rep. 2019, 42, 425–435. [Google Scholar] [CrossRef] [PubMed]
  42. Qin, J.; Wang, Y.; Bai, Y.; Yang, K.; Mao, Q.; Lin, Y.; Kong, D.; Zheng, X.; Xie, L. Epigallocatechin-3-Gallate Inhibits Bladder Cancer Cell Invasion via Suppression of NF-ΚB-mediated Matrix Metalloproteinase-9 Expression. Mol. Med. Rep. 2012, 6, 1040–1044. [Google Scholar] [CrossRef] [Green Version]
  43. Piwowarczyk, L.; Stawny, M.; Mlynarczyk, D.T.; Muszalska-Kolos, I.; Goslinski, T.; Jelińska, A. Role of Curcumin and (-)-Epigallocatechin-3-O-Gallate in Bladder Cancer Treatment: A Review. Cancers 2020, 12, 1801. [Google Scholar] [CrossRef]
  44. Zan, L.; Chen, Q.; Zhang, L.; Li, X. Epigallocatechin Gallate (EGCG) Suppresses Growth and Tumorigenicity in Breast Cancer Cells by Downregulation of MiR-25. Bioengineered 2019, 10, 374–382. [Google Scholar] [CrossRef] [Green Version]
  45. Wei, R.; Mao, L.; Xu, P.; Zheng, X.; Hackman, R.M.; Mackenzie, G.G.; Wang, Y. Suppressing Glucose Metabolism with Epigallocatechin-3-Gallate (EGCG) Reduces Breast Cancer Cell Growth in Preclinical Models. Food Funct. 2018, 9, 5682–5696. [Google Scholar] [CrossRef]
  46. 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] [PubMed]
  47. Gu, J.-W.; Makey, K.L.; Tucker, K.B.; Chinchar, E.; Mao, X.; Pei, I.; Thomas, E.Y.; Miele, L. EGCG, a Major Green Tea Catechin Suppresses Breast Tumor Angiogenesis and Growth via Inhibiting the Activation of HIF-1α and NFκB, and VEGF Expression. Vasc. Cell 2013, 5, 9. [Google Scholar] [CrossRef] [Green Version]
  48. Thangapazham, R.L.; Passi, N.; Maheshwari, R.K. Green Tea Polyphenol and Epigallocatechin Gallate Induce Apoptosis and Inhibit Invasion in Human Breast Cancer Cells. Cancer Biol. Ther. 2007, 6, 1938–1943. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Zhao, X.; Tian, H.; Ma, X.; Li, L. Epigallocatechin Gallate, the Main Ingredient of Green Tea Induces Apoptosis in Breast Cancer Cells. Front. Biosci. 2006, 11, 2428–2433. [Google Scholar] [CrossRef] [Green Version]
  50. Mittal, A.; Pate, M.S.; Wylie, R.C.; Tollefsbol, T.O.; Katiyar, S.K. EGCG Down-Regulates Telomerase in Human Breast Carcinoma MCF-7 Cells, Leading to Suppression of Cell Viability and Induction of Apoptosis. Int. J. Oncol. 2004, 24, 703–710. [Google Scholar] [CrossRef]
  51. Romano, A.; Martel, F. The Role of EGCG in Breast Cancer Prevention and Therapy. Mini Rev. Med. Chem. 2021, 21, 883–898. [Google Scholar] [CrossRef] [PubMed]
  52. Tang, Y.; Zhao, D.Y.; Elliott, S.; Zhao, W.; Curiel, T.J.; Beckman, B.S.; Burow, M.E. Epigallocatechin-3 Gallate Induces Growth Inhibition and Apoptosis in Human Breast Cancer Cells through Survivin Suppression. Int. J. Oncol. 2007, 31, 705–711. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Ahn, W.S.; Huh, S.W.; Bae, S.-M.; Lee, I.P.; Lee, J.M.; Namkoong, S.E.; Kim, C.K.; Sin, J.-I. A Major Constituent of Green Tea, EGCG, Inhibits the Growth of a Human Cervical Cancer Cell Line, CaSki Cells, through Apoptosis, G(1) Arrest, and Regulation of Gene Expression. DNA Cell Biol. 2003, 22, 217–224. [Google Scholar] [CrossRef] [PubMed]
  54. Zou, C.; Liu, H.; Feugang, J.M.; Hao, Z.; Chow, H.-H.S.; Garcia, F. Green Tea Compound in Chemoprevention of Cervical Cancer. Int. J. Gynecol. Cancer 2010, 20, 617–624. [Google Scholar] [CrossRef] [Green Version]
  55. Singh, M.; Singh, R.; Bhui, K.; Tyagi, S.; Mahmood, Z.; Shukla, Y. Tea Polyphenols Induce Apoptosis through Mitochondrial Pathway and by Inhibiting Nuclear Factor-KappaB and Akt Activation in Human Cervical Cancer Cells. Oncol. Res. 2011, 19, 245–257. [Google Scholar] [CrossRef]
  56. Sharma, C.; Nusri, Q.E.-A.; Begum, S.; Javed, E.; Rizvi, T.A.; Hussain, A. (−)-Epigallocatechin-3-Gallate Induces Apoptosis and Inhibits Invasion and Migration of Human Cervical Cancer Cells. Asian Pac. J. Cancer Prev. 2012, 13, 4815–4822. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Khan, M.A.; Hussain, A.; Sundaram, M.K.; Alalami, U.; Gunasekera, D.; Ramesh, L.; Hamza, A.; Quraishi, U. (−)-Epigallocatechin-3-Gallate Reverses the Expression of Various Tumor-Suppressor Genes by Inhibiting DNA Methyltransferases and Histone Deacetylases in Human Cervical Cancer Cells. Oncol. Rep. 2015, 33, 1976–1984. [Google Scholar] [CrossRef] [Green Version]
  58. Almatroodi, S.A.; Almatroudi, A.; Khan, A.A.; Alhumaydhi, F.A.; Alsahli, M.A.; Rahmani, A.H. Potential Therapeutic Targets of Epigallocatechin Gallate (EGCG), the Most Abundant Catechin in Green Tea, and Its Role in the Therapy of Various Types of Cancer. Molecules 2020, 25, 3146. [Google Scholar] [CrossRef]
  59. Panji, M.; Behmard, V.; Zare, Z.; Malekpour, M.; Nejadbiglari, H.; Yavari, S.; Nayerpour Dizaj, T.; Safaeian, A.; Maleki, N.; Abbasi, M.; et al. Suppressing Effects of Green Tea Extract and Epigallocatechin-3-Gallate (EGCG) on TGF-β- Induced Epithelial-to-Mesenchymal Transition via ROS/Smad Signaling in Human Cervical Cancer Cells. Gene 2021, 794, 145774. [Google Scholar] [CrossRef]
  60. La, X.; Zhang, L.; Li, Z.; Li, H.; Yang, Y. (−)-Epigallocatechin Gallate (EGCG) Enhances the Sensitivity of Colorectal Cancer Cells to 5-FU by Inhibiting GRP78/NF-ΚB/MiR-155-5p/MDR1 Pathway. J. Agric. Food Chem. 2019, 67, 2510–2518. [Google Scholar] [CrossRef]
  61. Wu, W.; Dong, J.; Gou, H.; Geng, R.; Yang, X.; Chen, D.; Xiang, B.; Zhang, Z.; Ren, S.; Chen, L.; et al. EGCG Synergizes the Therapeutic Effect of Irinotecan through Enhanced DNA Damage in Human Colorectal Cancer Cells. J. Cell Mol. Med. 2021, 25, 7913–7921. [Google Scholar] [CrossRef]
  62. Luo, K.-W.; Xia, J.; Cheng, B.-H.; Gao, H.-C.; Fu, L.-W.; Luo, X.-L. Tea Polyphenol EGCG Inhibited Colorectal-Cancer-Cell Proliferation and Migration via Downregulation of STAT3. Gastroenterol. Rep. 2021, 9, 59–70. [Google Scholar] [CrossRef]
  63. Wu, W.; Gou, H.; Xiang, B.; Geng, R.; Dong, J.; Yang, X.; Chen, D.; Dai, R.; Chen, L.; Liu, J. EGCG Enhances the Chemosensitivity of Colorectal Cancer to Irinotecan through GRP78-MediatedEndoplasmic Reticulum Stress. J. Oncol. 2022, 2022, 7099589. [Google Scholar] [CrossRef]
  64. Du, G.-J.; Zhang, Z.; Wen, X.-D.; Yu, C.; Calway, T.; Yuan, C.-S.; Wang, C.-Z. Epigallocatechin Gallate (EGCG) Is the Most Effective Cancer Chemopreventive Polyphenol in Green Tea. Nutrients 2012, 4, 1679–1691. [Google Scholar] [CrossRef]
  65. Jin, H.; Gong, W.; Zhang, C.; Wang, S. Epigallocatechin Gallate Inhibits the Proliferation of Colorectal Cancer Cells by Regulating Notch Signaling. OncoTargets Ther. 2013, 6, 145–153. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Fu, J.-D.; Yao, J.-J.; Wang, H.; Cui, W.-G.; Leng, J.; Ding, L.-Y.; Fan, K.-Y. Effects of EGCG on Proliferation and Apoptosis of Gastric Cancer SGC7901 Cells via Down-Regulation of HIF-1α and VEGF under a Hypoxic State. Eur. Rev. Med. Pharmacol. Sci. 2019, 23, 155–161. [Google Scholar] [CrossRef] [PubMed]
  67. Yang, C.; Du, W.; Yang, D. Inhibition of Green Tea Polyphenol EGCG((−)-Epigallocatechin-3-Gallate) on the Proliferation of Gastric Cancer Cells by Suppressing Canonical Wnt/β-Catenin Signalling Pathway. Int. J. Food Sci. Nutr. 2016, 67, 818–827. [Google Scholar] [CrossRef] [PubMed]
  68. Onoda, C.; Kuribayashi, K.; Nirasawa, S.; Tsuji, N.; Tanaka, M.; Kobayashi, D.; Watanabe, N. (−)-Epigallocatechin-3-Gallate Induces Apoptosis in Gastric Cancer Cell Lines by down-Regulating Survivin Expression. Int. J. Oncol. 2011, 38, 1403–1408. [Google Scholar] [CrossRef] [Green Version]
  69. Zhu, B.-H.; Zhan, W.-H.; Li, Z.-R.; Wang, Z.; He, Y.-L.; Peng, J.-S.; Cai, S.-R.; Ma, J.-P.; Zhang, C.-H. (−)-Epigallocatechin-3-Gallate Inhibits Growth of Gastric Cancer by Reducing VEGF Production and Angiogenesis. World J. Gastroenterol. 2007, 13, 1162–1169. [Google Scholar] [CrossRef] [Green Version]
  70. Sojoodi, M.; Wei, L.; Erstad, D.J.; Yamada, S.; Fujii, T.; Hirschfield, H.; Kim, R.S.; Lauwers, G.Y.; Lanuti, M.; Hoshida, Y.; et al. Epigallocatechin Gallate Induces Hepatic Stellate Cell Senescence and Attenuates Development of Hepatocellular Carcinoma. Cancer Prev. Res. 2020, 13, 497–508. [Google Scholar] [CrossRef] [Green Version]
  71. Bimonte, S.; Albino, V.; Piccirillo, M.; Nasto, A.; Molino, C.; Palaia, R.; Cascella, M. Epigallocatechin-3-Gallate in the Prevention and Treatment of Hepatocellular Carcinoma: Experimental Findings and Translational Perspectives. Drug Des. Dev. Ther. 2019, 13, 611–621. [Google Scholar] [CrossRef] [Green Version]
  72. Nishikawa, T.; Nakajima, T.; Moriguchi, M.; Jo, M.; Sekoguchi, S.; Ishii, M.; Takashima, H.; Katagishi, T.; Kimura, H.; Minami, M.; et al. A Green Tea Polyphenol, Epigalocatechin-3-Gallate, Induces Apoptosis of Human Hepatocellular Carcinoma, Possibly through Inhibition of Bcl-2 Family Proteins. J. Hepatol. 2006, 44, 1074–1082. [Google Scholar] [CrossRef]
  73. Bravi, F.; La Vecchia, C.; Turati, F. Green Tea and Liver Cancer. Hepatobiliary Surg. Nutr. 2017, 6, 127–129. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Sakamoto, Y.; Terashita, N.; Muraguchi, T.; Fukusato, T.; Kubota, S. Effects of Epigallocatechin-3-Gallate (EGCG) on A549 Lung Cancer Tumor Growth and Angiogenesis. Biosci. Biotechnol. Biochem. 2013, 77, 1799–1803. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Wang, H.; Bian, S.; Yang, C.S. Green Tea Polyphenol EGCG Suppresses Lung Cancer Cell Growth through Upregulating MiR-210 Expression Caused by Stabilizing HIF-1α. Carcinogenesis 2011, 32, 1881–1889. [Google Scholar] [CrossRef] [Green Version]
  76. Dhatwalia, S.K.; Kumar, M.; Dhawan, D.K. Role of EGCG in Containing the Progression of Lung Tumorigenesis—A Multistage Targeting Approach. Nutr. Cancer 2018, 70, 334–349. [Google Scholar] [CrossRef]
  77. Shin, Y.S.; Kang, S.U.; Park, J.K.; Kim, Y.E.; Kim, Y.S.; Baek, S.J.; Lee, S.-H.; Kim, C.-H. Anti-Cancer Effect of (−)-Epigallocatechin-3-Gallate (EGCG) in Head and Neck Cancer through Repression of Transactivation and Enhanced Degradation of β-Catenin. Phytomedicine 2016, 23, 1344–1355. [Google Scholar] [CrossRef] [PubMed]
  78. Katiyar, S.K. Emerging Phytochemicals for the Prevention and Treatment of Head and Neck Cancer. Molecules 2016, 21, 1610. [Google Scholar] [CrossRef] [Green Version]
  79. Wang, X.; Hao, M.-W.; Dong, K.; Lin, F.; Ren, J.-H.; Zhang, H.-Z. Apoptosis Induction Effects of EGCG in Laryngeal Squamous Cell Carcinoma Cells through Telomerase Repression. Arch. Pharm. Res. 2009, 32, 1263–1269. [Google Scholar] [CrossRef]
  80. Sarniak, A.; Lipińska, J.; Tytman, K.; Lipińska, S. Endogenous Mechanisms of Reactive Oxygen Species (ROS) Generation. Postepy Hig. Med. Dosw. 2016, 70, 1150–1165. [Google Scholar] [CrossRef]
  81. Sandalio, L.M.; Rodríguez-Serrano, M.; Romero-Puertas, M.C.; del Río, L.A. Role of Peroxisomes as a Source of Reactive Oxygen Species (ROS) Signaling Molecules. Subcell. Biochem. 2013, 69, 231–255. [Google Scholar] [CrossRef]
  82. Henle, E.S.; Linn, S. Formation, Prevention, and Repair of DNA Damage by Iron/Hydrogen Peroxide. J. Biol. Chem. 1997, 272, 19095–19098. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Schieber, M.; Chandel, N.S. ROS Function in Redox Signaling and Oxidative Stress. Curr. Biol. 2014, 24, R453–R462. [Google Scholar] [CrossRef] [Green Version]
  84. Cadet, J.; Douki, T.; Ravanat, J.-L. Measurement of Oxidatively Generated Base Damage in Cellular DNA. Mutat. Res. 2011, 711, 3–12. [Google Scholar] [CrossRef] [PubMed]
  85. Cadet, J.; Douki, T.; Ravanat, J.-L. Oxidatively Generated Base Damage to Cellular DNA. Free Radic. Biol. Med. 2010, 49, 9–21. [Google Scholar] [CrossRef] [PubMed]
  86. Cadet, J.; Ravanat, J.-L.; TavernaPorro, M.; Menoni, H.; Angelov, D. Oxidatively Generated Complex DNA Damage: Tandem and Clustered Lesions. Cancer Lett. 2012, 327, 5–15. [Google Scholar] [CrossRef]
  87. Nita, M.; Grzybowski, A. The Role of the Reactive Oxygen Species and Oxidative Stress in the Pathomechanism of the Age-Related Ocular Diseases and Other Pathologies of the Anterior and Posterior Eye Segments in Adults. Oxid. Med. Cell Longev. 2016, 2016, 3164734. [Google Scholar] [CrossRef] [Green Version]
  88. Snezhkina, A.V.; Kudryavtseva, A.V.; Kardymon, O.L.; Savvateeva, M.V.; Melnikova, N.V.; Krasnov, G.S.; Dmitriev, A.A. ROS Generation and Antioxidant Defense Systems in Normal and Malignant Cells. Oxid. Med. Cell Longev. 2019, 2019, 6175804. [Google Scholar] [CrossRef] [Green Version]
  89. Chatterjee, N.; Walker, G.C. Mechanisms of DNA Damage, Repair, and Mutagenesis. Environ. Mol. Mutagen. 2017, 58, 235–263. [Google Scholar] [CrossRef] [Green Version]
  90. Ljungman, M.; Hanawalt, P.C. Efficient Protection against Oxidative DNA Damage in Chromatin. Mol. Carcinog. 1992, 5, 264–269. [Google Scholar] [CrossRef]
  91. Kreuz, S.; Fischle, W. Oxidative Stress Signaling to Chromatin in Health and Disease. Epigenomics 2016, 8, 843–862. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Jakubczyk, K.; Dec, K.; Kałduńska, J.; Kawczuga, D.; Kochman, J.; Janda, K. Reactive Oxygen Species—Sources, Functions, Oxidative Damage. Pol. Merkur. Lekarski 2020, 48, 124–127. [Google Scholar] [PubMed]
  93. Bergamini, C.M.; Gambetti, S.; Dondi, A.; Cervellati, C. Oxygen, Reactive Oxygen Species and Tissue Damage. Curr. Pharm. Des. 2004, 10, 1611–1626. [Google Scholar] [CrossRef] [PubMed]
  94. Mokra, D.; Adamcakova, J.; Mokry, J. Green Tea Polyphenol (−)-Epigallocatechin-3-Gallate (EGCG): A Time for a New Player in the Treatment of Respiratory Diseases? Antioxidants 2022, 11, 1566. [Google Scholar] [CrossRef] [PubMed]
  95. Khan, N.; Adhami, V.M.; Mukhtar, H. Review: Green Tea Polyphenols in Chemoprevention of Prostate Cancer: Preclinical and Clinical Studies. Nutr. Cancer 2009, 61, 836–841. [Google Scholar] [CrossRef] [Green Version]
  96. Kim, H.-S.; Quon, M.J.; Kim, J. New Insights into the Mechanisms of Polyphenols beyond Antioxidant Properties; Lessons from the Green Tea Polyphenol, Epigallocatechin 3-Gallate. Redox Biol. 2014, 2, 187–195. [Google Scholar] [CrossRef] [Green Version]
  97. Legeay, S.; Rodier, M.; Fillon, L.; Faure, S.; Clere, N. Epigallocatechin Gallate: A Review of Its Beneficial Properties to Prevent Metabolic Syndrome. Nutrients 2015, 7, 5443. [Google Scholar] [CrossRef] [Green Version]
  98. Zhu, S.; Li, Y.; Li, Z.; Ma, C.; Lou, Z.; Yokoyama, W.; Wang, H. Lipase-Catalyzed Synthesis of Acetylated EGCG and Antioxidant Properties of the Acetylated Derivatives. Food Res. Int. 2014, 56, 279–286. [Google Scholar] [CrossRef]
  99. Greten, F.R.; Grivennikov, S.I. Inflammation and Cancer: Triggers, Mechanisms and Consequences. Immunity 2019, 51, 27–41. [Google Scholar] [CrossRef]
  100. Murata, M. Inflammation and Cancer. Environ. Health Prev. Med. 2018, 23, 50. [Google Scholar] [CrossRef] [Green Version]
  101. Coussens, L.M.; Werb, Z. Inflammation and Cancer. Nature 2002, 420, 860–867. [Google Scholar] [CrossRef] [PubMed]
  102. Kciuk, M.; Kołat, D.; Kałuzińska-Kołat, Ż.; Gawrysiak, M.; Drozda, R.; Celik, I.; Kontek, R. PD-1/PD-L1 and DNA Damage Response in Cancer. Cells 2023, 12, 530. [Google Scholar] [CrossRef] [PubMed]
  103. Wang, X.; Lin, Y. Tumor Necrosis Factor and Cancer, Buddies or Foes? Acta Pharmacol. Sin. 2008, 29, 1275–1288. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. van Horssen, R.; Ten Hagen, T.L.M.; Eggermont, A.M.M. TNF-Alpha in Cancer Treatment: Molecular Insights, Antitumor Effects, and Clinical Utility. Oncologist 2006, 11, 397–408. [Google Scholar] [CrossRef] [PubMed]
  105. Wajant, H. The Role of TNF in Cancer. Results Probl. Cell Differ. 2009, 49, 1–15. [Google Scholar] [CrossRef] [PubMed]
  106. Balkwill, F. TNF-Alpha in Promotion and Progression of Cancer. Cancer Metastasis Rev. 2006, 25, 409–416. [Google Scholar] [CrossRef]
  107. Shen, J.; Xiao, Z.; Zhao, Q.; Li, M.; Wu, X.; Zhang, L.; Hu, W.; Cho, C.H. Anti-Cancer Therapy with TNFα and IFNγ: A Comprehensive Review. Cell Prolif. 2018, 51, e12441. [Google Scholar] [CrossRef] [Green Version]
  108. Wang, Z.-M.; Gao, W.; Wang, H.; Zhao, D.; Nie, Z.-L.; Shi, J.-Q.; Zhao, S.; Lu, X.; Wang, L.-S.; Yang, Z.-J. Green Tea Polyphenol Epigallocatechin-3-Gallate Inhibits TNF-α-Induced Production of Monocyte Chemoattractant Protein-1 in Human Umbilical Vein Endothelial Cells. Cell Physiol. Biochem. 2014, 33, 1349–1358. [Google Scholar] [CrossRef]
  109. Shin, H.-Y.; Kim, S.-H.; Jeong, H.-J.; Kim, S.-Y.; Shin, T.-Y.; Um, J.-Y.; Hong, S.-H.; Kim, H.-M. Epigallocatechin-3-Gallate Inhibits Secretion of TNF-Alpha, IL-6 and IL-8 through the Attenuation of ERK and NF-KappaB in HMC-1 Cells. Int. Arch. Allergy Immunol. 2007, 142, 335–344. [Google Scholar] [CrossRef]
  110. Pang, L.Y.; Hurst, E.A.; Argyle, D.J. Cyclooxygenase-2: A Role in Cancer Stem Cell Survival and Repopulation of Cancer Cells during Therapy. Stem Cells Int. 2016, 2016, 2048731. [Google Scholar] [CrossRef] [Green Version]
  111. Hashemi Goradel, N.; Najafi, M.; Salehi, E.; Farhood, B.; Mortezaee, K. Cyclooxygenase-2 in Cancer: A Review. J. Cell Physiol. 2019, 234, 5683–5699. [Google Scholar] [CrossRef] [PubMed]
  112. Sinicrope, F.A. Targeting Cyclooxygenase-2 for Prevention and Therapy of Colorectal Cancer. Mol. Carcinog. 2006, 45, 447–454. [Google Scholar] [CrossRef] [PubMed]
  113. Xu, X.-C. COX-2 Inhibitors in Cancer Treatment and Prevention, a Recent Development. Anti-Cancer Drugs 2002, 13, 127–137. [Google Scholar] [CrossRef] [PubMed]
  114. Arun, B.; Goss, P. The Role of COX-2 Inhibition in Breast Cancer Treatment and Prevention. Semin. Oncol. 2004, 31, 22–29. [Google Scholar] [CrossRef]
  115. Sinicrope, F.A.; Gill, S. Role of Cyclooxygenase-2 in Colorectal Cancer. Cancer Metastasis Rev. 2004, 23, 63–75. [Google Scholar] [CrossRef]
  116. Peng, G.; Dixon, D.A.; Muga, S.J.; Smith, T.J.; Wargovich, M.J. Green Tea Polyphenol (−)-Epigallocatechin-3-Gallate Inhibits Cyclooxygenase-2 Expression in Colon Carcinogenesis. Mol. Carcinog. 2006, 45, 309–319. [Google Scholar] [CrossRef]
  117. Hussain, T.; Gupta, S.; Adhami, V.M.; Mukhtar, H. Green Tea Constituent Epigallocatechin-3-Gallate Selectively Inhibits COX-2 without Affecting COX-1 Expression in Human Prostate Carcinoma Cells. Int. J. Cancer 2005, 113, 660–669. [Google Scholar] [CrossRef]
  118. Adhami, V.M.; Malik, A.; Zaman, N.; Sarfaraz, S.; Siddiqui, I.A.; Syed, D.N.; Afaq, F.; Pasha, F.S.; Saleem, M.; Mukhtar, H. Combined Inhibitory Effects of Green Tea Polyphenols and Selective Cyclooxygenase-2 Inhibitors on the Growth of Human Prostate Cancer Cells Both in Vitro and in Vivo. Clin. Cancer Res. 2007, 13, 1611–1619. [Google Scholar] [CrossRef] [Green Version]
  119. Harper, C.E.; Patel, B.B.; Wang, J.; Eltoum, I.A.; Lamartiniere, C.A. Epigallocatechin-3-Gallate Suppresses Early Stage, but Not Late Stage Prostate Cancer in TRAMP Mice: Mechanisms of Action. Prostate 2007, 67, 1576–1589. [Google Scholar] [CrossRef]
  120. Xu, W.; Liu, L.Z.; Loizidou, M.; Ahmed, M.; Charles, I.G. The Role of Nitric Oxide in Cancer. Cell Res. 2002, 12, 311–320. [Google Scholar] [CrossRef] [Green Version]
  121. Khan, F.H.; Dervan, E.; Bhattacharyya, D.D.; McAuliffe, J.D.; Miranda, K.M.; Glynn, S.A. The Role of Nitric Oxide in Cancer: Master Regulator or NOt? Int. J. Mol. Sci. 2020, 21, 9393. [Google Scholar] [CrossRef] [PubMed]
  122. Somasundaram, V.; Basudhar, D.; Bharadwaj, G.; No, J.H.; Ridnour, L.A.; Cheng, R.Y.S.; Fujita, M.; Thomas, D.D.; Anderson, S.K.; McVicar, D.W.; et al. Molecular Mechanisms of Nitric Oxide in Cancer Progression, Signal Transduction, and Metabolism. Antioxid. Redox Signal. 2019, 30, 1124–1143. [Google Scholar] [CrossRef] [PubMed]
  123. Chan, M.M.; Fong, D.; Ho, C.T.; Huang, H.I. Inhibition of Inducible Nitric Oxide Synthase Gene Expression and Enzyme Activity by Epigallocatechin Gallate, a Natural Product from Green Tea. Biochem. Pharmacol. 1997, 54, 1281–1286. [Google Scholar] [CrossRef] [PubMed]
  124. Lin, Y.L.; Lin, J.K. (−)-Epigallocatechin-3-Gallate Blocks the Induction of Nitric Oxide Synthase by down-Regulating Lipopolysaccharide-Induced Activity of Transcription Factor Nuclear Factor-KappaB. Mol. Pharmacol. 1997, 52, 465–472. [Google Scholar]
  125. Xia, L.; Tan, S.; Zhou, Y.; Lin, J.; Wang, H.; Oyang, L.; Tian, Y.; Liu, L.; Su, M.; Wang, H.; et al. Role of the NFκB-Signaling Pathway in Cancer. OncoTargets Ther. 2018, 11, 2063–2073. [Google Scholar] [CrossRef] [Green Version]
  126. Xia, Y.; Shen, S.; Verma, I.M. NF-ΚB, an Active Player in Human Cancers. Cancer Immunol. Res. 2014, 2, 823. [Google Scholar] [CrossRef] [Green Version]
  127. Itoh, Y.; Nagase, H. Matrix Metalloproteinases in Cancer. Essays Biochem. 2002, 38, 21–36. [Google Scholar] [CrossRef] [Green Version]
  128. Gialeli, C.; Theocharis, A.D.; Karamanos, N.K. Roles of Matrix Metalloproteinases in Cancer Progression and Their Pharmacological Targeting. FEBS J. 2011, 278, 16–27. [Google Scholar] [CrossRef]
  129. Mustafa, S.; Koran, S.; AlOmair, L. Insights Into the Role of Matrix Metalloproteinases in Cancer and Its Various Therapeutic Aspects: A Review. Front. Mol. Biosci. 2022, 9, 896099. [Google Scholar] [CrossRef]
  130. Chowdhury, A.; Nandy, S.K.; Sarkar, J.; Chakraborti, T.; Chakraborti, S. Inhibition of Pro-/Active MMP-2 by Green Tea Catechins and Prediction of Their Interaction by Molecular Docking Studies. Mol. Cell Biochem. 2017, 427, 111–122. [Google Scholar] [CrossRef]
  131. Sarkar, J.; Nandy, S.K.; Chowdhury, A.; Chakraborti, T.; Chakraborti, S. Inhibition of MMP-9 by Green Tea Catechins and Prediction of Their Interaction by Molecular Docking Analysis. Biomed. Pharmacother. 2016, 84, 340–347. [Google Scholar] [CrossRef]
  132. Maeda-Yamamoto, M.; Kawahara, H.; Tahara, N.; Tsuji, K.; Hara, Y.; Isemura, M. Effects of Tea Polyphenols on the Invasion and Matrix Metalloproteinases Activities of Human Fibrosarcoma HT1080 Cells. J. Agric. Food Chem. 1999, 47, 2350–2354. [Google Scholar] [CrossRef]
  133. Ilango, S.; Paital, B.; Jayachandran, P.; Padma, P.R.; Nirmaladevi, R. Epigenetic Alterations in Cancer. Front. Biosci. 2020, 25, 1058–1109. [Google Scholar] [CrossRef]
  134. Kanwal, R.; Gupta, S. Epigenetic Modifications in Cancer. Clin. Genet. 2012, 81, 303–311. [Google Scholar] [CrossRef] [Green Version]
  135. Miranda Furtado, C.L.; Dos Santos Luciano, M.C.; Silva Santos, R.D.; Furtado, G.P.; Moraes, M.O.; Pessoa, C. Epidrugs: Targeting Epigenetic Marks in Cancer Treatment. Epigenetics 2019, 14, 1164–1176. [Google Scholar] [CrossRef]
  136. Zhang, J.; Yang, C.; Wu, C.; Cui, W.; Wang, L. DNA Methyltransferases in Cancer: Biology, Paradox, Aberrations, and Targeted Therapy. Cancers 2020, 12, 2123. [Google Scholar] [CrossRef] [PubMed]
  137. Peng, D.-F.; Kanai, Y.; Sawada, M.; Ushijima, S.; Hiraoka, N.; Kitazawa, S.; Hirohashi, S. DNA Methylation of Multiple Tumor-Related Genes in Association with Overexpression of DNA Methyltransferase 1 (DNMT1) during Multistage Carcinogenesis of the Pancreas. Carcinogenesis 2006, 27, 1160–1168. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Saito, Y.; Kanai, Y.; Nakagawa, T.; Sakamoto, M.; Saito, H.; Ishii, H.; Hirohashi, S. Increased Protein Expression of DNA Methyltransferase (DNMT) 1 Is Significantly Correlated with the Malignant Potential and Poor Prognosis of Human Hepatocellular Carcinomas. Int. J. Cancer 2003, 105, 527–532. [Google Scholar] [CrossRef] [PubMed]
  139. Butcher, D.T.; Rodenhiser, D.I. Epigenetic Inactivation of BRCA1 Is Associated with Aberrant Expression of CTCF and DNA Methyltransferase (DNMT3B) in Some Sporadic Breast Tumours. Eur. J. Cancer 2007, 43, 210–219. [Google Scholar] [CrossRef] [PubMed]
  140. Linhart, H.G.; Lin, H.; Yamada, Y.; Moran, E.; Steine, E.J.; Gokhale, S.; Lo, G.; Cantu, E.; Ehrich, M.; He, T.; et al. Dnmt3b Promotes Tumorigenesis in Vivo by Gene-Specific de Novo Methylation and Transcriptional Silencing. Genes. Dev. 2007, 21, 3110–3122. [Google Scholar] [CrossRef] [Green Version]
  141. Zhang, W.; Xu, J. DNA Methyltransferases and Their Roles in Tumorigenesis. Biomark. Res. 2017, 5, 1. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Tang, X.-H.; Gudas, L.J. Retinoids, Retinoic Acid Receptors, and Cancer. Annu. Rev. Pathol. 2011, 6, 345–364. [Google Scholar] [CrossRef] [PubMed]
  143. Tsuji, K.; Utsunomiya, H.; Miki, Y.; Hanihara, M.; Fue, M.; Takagi, K.; Nishimoto, M.; Suzuki, F.; Yaegashi, N.; Suzuki, T.; et al. Retinoic Acid Receptor β: A Potential Therapeutic Target in Retinoic Acid Treatment of Endometrial Cancer. Int. J. Gynecol. Cancer 2017, 27, 643–650. [Google Scholar] [CrossRef]
  144. Berx, G.; van Roy, F. Involvement of Members of the Cadherin Superfamily in Cancer. Cold Spring Harb. Perspect. Biol. 2009, 1, a003129. [Google Scholar] [CrossRef] [PubMed]
  145. Chen, D.; Zhou, X.Z.; Lee, T.H. Death-Associated Protein Kinase 1 as a Promising Drug Target in Cancer and Alzheimer’s Disease. Recent Pat. Anti-Cancer Drug Discov. 2019, 14, 144–157. [Google Scholar] [CrossRef]
  146. Nandakumar, V.; Vaid, M.; Katiyar, S.K. (−)-Epigallocatechin-3-Gallate Reactivates Silenced Tumor Suppressor Genes, Cip1/P21 and P16INK4a, by Reducing DNA Methylation and Increasing Histones Acetylation in Human Skin Cancer Cells. Carcinogenesis 2011, 32, 537–544. [Google Scholar] [CrossRef] [Green Version]
  147. Barneda-Zahonero, B.; Parra, M. Histone Deacetylases and Cancer. Mol. Oncol. 2012, 6, 579–589. [Google Scholar] [CrossRef] [Green Version]
  148. Lee, Y.-H.; Kwak, J.; Choi, H.-K.; Choi, K.-C.; Kim, S.; Lee, J.; Jun, W.; Park, H.-J.; Yoon, H.-G. EGCG Suppresses Prostate Cancer Cell Growth Modulating Acetylation of Androgen Receptor by Anti-Histone Acetyltransferase Activity. Int. J. Mol. Med. 2012, 30, 69–74. [Google Scholar] [CrossRef] [Green Version]
  149. Choi, K.-C.; Jung, M.G.; Lee, Y.-H.; Yoon, J.C.; Kwon, S.H.; Kang, H.-B.; Kim, M.-J.; Cha, J.-H.; Kim, Y.J.; Jun, W.J.; et al. Epigallocatechin-3-Gallate, a Histone Acetyltransferase Inhibitor, Inhibits EBV-Induced B Lymphocyte Transformation via Suppression of RelA Acetylation. Cancer Res. 2009, 69, 583–592. [Google Scholar] [CrossRef] [Green Version]
  150. Borutinskaitė, V.; Virkšaitė, A.; Gudelytė, G.; Navakauskienė, R. Green Tea Polyphenol EGCG Causes Anti-Cancerous Epigenetic Modulations in Acute Promyelocytic Leukemia Cells. Leuk. Lymphoma 2018, 59, 469–478. [Google Scholar] [CrossRef]
  151. Deb, G.; Thakur, V.S.; Limaye, A.M.; Gupta, S. Epigenetic Induction of Tissue Inhibitor of Matrix Metalloproteinase-3 by Green Tea Polyphenols in Breast Cancer Cells. Mol. Carcinog. 2015, 54, 485–499. [Google Scholar] [CrossRef]
  152. Deb, G.; Shankar, E.; Thakur, V.S.; Ponsky, L.E.; Bodner, D.R.; Fu, P.; Gupta, S. Green Tea-Induced Epigenetic Reactivation of Tissue Inhibitor of Matrix Metalloproteinase-3 Suppresses Prostate Cancer Progression through Histone-Modifying Enzymes. Mol. Carcinog. 2019, 58, 1194–1207. [Google Scholar] [CrossRef] [PubMed]
  153. Oya, Y.; Mondal, A.; Rawangkan, A.; Umsumarng, S.; Iida, K.; Watanabe, T.; Kanno, M.; Suzuki, K.; Li, Z.; Kagechika, H.; et al. Down-Regulation of Histone Deacetylase 4, -5 and -6 as a Mechanism of Synergistic Enhancement of Apoptosis in Human Lung Cancer Cells Treated with the Combination of a Synthetic Retinoid, Am80 and Green Tea Catechin. J. Nutr. Biochem. 2017, 42, 7–16. [Google Scholar] [CrossRef] [PubMed]
  154. Lemmon, M.A.; Schlessinger, J. Cell Signaling by Receptor Tyrosine Kinases. Cell 2010, 141, 1117–1134. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Schlessinger, J. Receptor Tyrosine Kinases: Legacy of the First Two Decades. Cold Spring Harb. Perspect. Biol. 2014, 6, a008912. [Google Scholar] [CrossRef] [Green Version]
  156. Sigismund, S.; Avanzato, D.; Lanzetti, L. Emerging Functions of the EGFR in Cancer. Mol. Oncol. 2018, 12, 3–20. [Google Scholar] [CrossRef] [Green Version]
  157. Tan, X.; Lambert, P.F.; Rapraeger, A.C.; Anderson, R.A. Stress-Induced EGFR Trafficking: Mechanisms, Functions, and Therapeutic Implications. Trends Cell Biol. 2016, 26, 352–366. [Google Scholar] [CrossRef] [Green Version]
  158. Roskoski, R. The ErbB/HER Family of Protein-Tyrosine Kinases and Cancer. Pharmacol. Res. 2014, 79, 34–74. [Google Scholar] [CrossRef]
  159. Wilson, K.J.; Mill, C.; Lambert, S.; Buchman, J.; Wilson, T.R.; Hernandez-Gordillo, V.; Gallo, R.M.; Ades, L.M.C.; Settleman, J.; Riese, D.J. EGFR Ligands Exhibit Functional Differences in Models of Paracrine and Autocrine Signaling. Growth Factors 2012, 30, 107–116. [Google Scholar] [CrossRef] [Green Version]
  160. Ferguson, K.M.; Berger, M.B.; Mendrola, J.M.; Cho, H.S.; Leahy, D.J.; Lemmon, M.A. EGF Activates Its Receptor by Removing Interactions That Autoinhibit Ectodomain Dimerization. Mol. Cell 2003, 11, 507–517. [Google Scholar] [CrossRef]
  161. Singh, B.; Carpenter, G.; Coffey, R.J. EGF Receptor Ligands: Recent Advances. F1000Research 2016, 5, F1000 Faculty Rev-2270. [Google Scholar] [CrossRef] [Green Version]
  162. Yarden, Y.; Sliwkowski, M.X. Untangling the ErbB Signalling Network. Nat. Rev. Mol. Cell Biol. 2001, 2, 127–137. [Google Scholar] [CrossRef] [PubMed]
  163. Lemmon, M.A.; Schlessinger, J.; Ferguson, K.M. The EGFR Family: Not so Prototypical Receptor Tyrosine Kinases. Cold Spring Harb. Perspect. Biol. 2014, 6, a020768. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Freed, D.M.; Bessman, N.J.; Kiyatkin, A.; Salazar-Cavazos, E.; Byrne, P.O.; Moore, J.O.; Valley, C.C.; Ferguson, K.M.; Leahy, D.J.; Lidke, D.S.; et al. EGFR Ligands Differentially Stabilize Receptor Dimers to Specify Signaling Kinetics. Cell 2017, 171, 683–695.e18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Garg, R.; Benedetti, L.G.; Abera, M.B.; Wang, H.; Abba, M.; Kazanietz, M.G. Protein Kinase C and Cancer: What We Know and What We Do Not. Oncogene 2014, 33, 5225–5237. [Google Scholar] [CrossRef] [Green Version]
  166. He, S.; Li, Q.; Huang, Q.; Cheng, J. Targeting Protein Kinase C for Cancer Therapy. Cancers 2022, 14, 1104. [Google Scholar] [CrossRef]
  167. Sun, X.-L.; Xiang, Z.-M.; Xie, Y.-R.; Zhang, N.; Wang, L.-X.; Wu, Y.-L.; Zhang, D.-Y.; Wang, X.-J.; Sheng, J.; Zi, C.-T. Dimeric-(−)-Epigallocatechin-3-Gallate Inhibits the Proliferation of Lung Cancer Cells by Inhibiting the EGFR Signaling Pathway. Chem. Biol. Interact. 2022, 365, 110084. [Google Scholar] [CrossRef]
  168. Honda, Y.; Takigawa, N.; Ichihara, E.; Ninomiya, T.; Kubo, T.; Ochi, N.; Yasugi, M.; Murakami, T.; Yamane, H.; Tanimoto, M.; et al. Effects of (−)-Epigallocatechin-3-Gallate on EGFR- or Fusion Gene-Driven Lung Cancer Cells. Acta Med. Okayama 2017, 71, 505–512. [Google Scholar] [CrossRef]
  169. Wu, D.; Liu, Z.; Li, J.; Zhang, Q.; Zhong, P.; Teng, T.; Chen, M.; Xie, Z.; Ji, A.; Li, Y. Epigallocatechin-3-Gallate Inhibits the Growth and Increases the Apoptosis of Human Thyroid Carcinoma Cells through Suppression of EGFR/RAS/RAF/MEK/ERK Signaling Pathway. Cancer Cell Int. 2019, 19, 43. [Google Scholar] [CrossRef] [Green Version]
  170. Chen, J.; Chen, L.; Lu, T.; Xie, Y.; Li, C.; Jia, Z.; Cao, J. ERα36 Is an Effective Target of Epigallocatechin-3-Gallate in Hepatocellular Carcinoma. Int. J. Clin. Exp. Pathol. 2019, 12, 3222–3234. [Google Scholar]
  171. Filippi, A.; Picot, T.; Aanei, C.M.; Nagy, P.; Szöllősi, J.; Campos, L.; Ganea, C.; Mocanu, M.-M. Epigallocatechin-3-O-Gallate Alleviates the Malignant Phenotype in A-431 Epidermoid and SK-BR-3 Breast Cancer Cell Lines. Int. J. Food Sci. Nutr. 2018, 69, 584–597. [Google Scholar] [CrossRef] [PubMed]
  172. Weng, L.X.; Wang, G.H.; Yao, H.; Yu, M.F.; Lin, J. Epigallocatechin Gallate Inhibits the Growth of Salivary Adenoid Cystic Carcinoma Cells via the EGFR/Erk Signal Transduction Pathway and the Mitochondria Apoptosis Pathway. Neoplasma 2017, 64, 563–570. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Hosseini, A.; Gharibi, T.; Marofi, F.; Javadian, M.; Babaloo, Z.; Baradaran, B. Janus Kinase Inhibitors: A Therapeutic Strategy for Cancer and Autoimmune Diseases. J. Cell Physiol. 2020, 235, 5903–5924. [Google Scholar] [CrossRef] [PubMed]
  174. Thoidingjam, L.K.; Blouin, C.M.; Gaillet, C.; Brion, A.; Solier, S.; Niyomchon, S.; El Marjou, A.; Mouasni, S.; Sepulveda, F.E.; de Saint Basile, G.; et al. Small Molecule Inhibitors of Interferon-Induced JAK-STAT Signalling. Angew. Chem. Int. Ed. Engl. 2022, 61, e202205231. [Google Scholar] [CrossRef]
  175. Hu, X.; Li, J.; Fu, M.; Zhao, X.; Wang, W. The JAK/STAT Signaling Pathway: From Bench to Clinic. Signal Transduct. Target. Ther. 2021, 6, 402. [Google Scholar] [CrossRef]
  176. Groner, B.; von Manstein, V. Jak Stat Signaling and Cancer: Opportunities, Benefits and Side Effects of Targeted Inhibition. Mol. Cell Endocrinol. 2017, 451, 1–14. [Google Scholar] [CrossRef]
  177. Ou, A.; Ott, M.; Fang, D.; Heimberger, A.B. The Role and Therapeutic Targeting of JAK/STAT Signaling in Glioblastoma. Cancers 2021, 13, 437. [Google Scholar] [CrossRef]
  178. Rah, B.; Rather, R.A.; Bhat, G.R.; Baba, A.B.; Mushtaq, I.; Farooq, M.; Yousuf, T.; Dar, S.B.; Parveen, S.; Hassan, R.; et al. JAK/STAT Signaling: Molecular Targets, Therapeutic Opportunities, and Limitations of Targeted Inhibitions in Solid Malignancies. Front. Pharmacol. 2022, 13, 821344. [Google Scholar] [CrossRef]
  179. Shao, F.; Pang, X.; Baeg, G.H. Targeting the JAK/STAT Signaling Pathway for Breast Cancer. Curr. Med. Chem. 2021, 28, 5137–5151. [Google Scholar] [CrossRef]
  180. Tang, S.-N.; Fu, J.; Shankar, S.; Srivastava, R.K. EGCG Enhances the Therapeutic Potential of Gemcitabine and CP690550 by Inhibiting STAT3 Signaling Pathway in Human Pancreatic Cancer. PLoS ONE 2012, 7, e31067. [Google Scholar] [CrossRef] [Green Version]
  181. Masuda, M.; Suzui, M.; Lim, J.T.E.; Weinstein, I.B. Epigallocatechin-3-Gallate Inhibits Activation of HER-2/Neu and Downstream Signaling Pathways in Human Head and Neck and Breast Carcinoma Cells. Clin. Cancer Res. 2003, 9, 3486–3491. [Google Scholar] [PubMed]
  182. Xiao, X.; Jiang, K.; Xu, Y.; Peng, H.; Wang, Z.; Liu, S.; Zhang, G. (−)-Epigallocatechin-3-Gallate Induces Cell Apoptosis in Chronic Myeloid Leukaemia by Regulating Bcr/Abl-Mediated P38-MAPK/JNK and JAK2/STAT3/AKT Signalling Pathways. Clin. Exp. Pharmacol. Physiol. 2019, 46, 126–136. [Google Scholar] [CrossRef] [PubMed]
  183. Senggunprai, L.; Kukongviriyapan, V.; Prawan, A.; Kukongviriyapan, U. Quercetin and EGCG Exhibit Chemopreventive Effects in Cholangiocarcinoma Cells via Suppression of JAK/STAT Signaling Pathway. Phytother. Res. 2014, 28, 841–848. [Google Scholar] [CrossRef]
  184. Wang, Y.; Ren, X.; Deng, C.; Yang, L.; Yan, E.; Guo, T.; Li, Y.; Xu, M.X. Mechanism of the Inhibition of the STAT3 Signaling Pathway by EGCG. Oncol. Rep. 2013, 30, 2691–2696. [Google Scholar] [CrossRef] [Green Version]
  185. Cheng, C.-W.; Shieh, P.-C.; Lin, Y.-C.; Chen, Y.-J.; Lin, Y.-H.; Kuo, D.-H.; Liu, J.-Y.; Kao, J.-Y.; Kao, M.-C.; Way, T.-D. Indoleamine 2,3-Dioxygenase, an Immunomodulatory Protein, Is Suppressed by (−)-Epigallocatechin-3-Gallate via Blocking of Gamma-Interferon-Induced JAK-PKC-Delta-STAT1 Signaling in Human Oral Cancer Cells. J. Agric. Food Chem. 2010, 58, 887–894. [Google Scholar] [CrossRef]
  186. Yuan, C.-H.; Horng, C.-T.; Lee, C.-F.; Chiang, N.-N.; Tsai, F.-J.; Lu, C.-C.; Chiang, J.-H.; Hsu, Y.-M.; Yang, J.-S.; Chen, F.-A. Epigallocatechin Gallate Sensitizes Cisplatin-Resistant Oral Cancer CAR Cell Apoptosis and Autophagy through Stimulating AKT/STAT3 Pathway and Suppressing Multidrug Resistance 1 Signaling. Environ. Toxicol. 2017, 32, 845–855. [Google Scholar] [CrossRef]
  187. Chang, L.; Karin, M. Mammalian MAP Kinase Signalling Cascades. Nature 2001, 410, 37–40. [Google Scholar] [CrossRef] [PubMed]
  188. Hepworth, E.M.W.; Hinton, S.D. Pseudophosphatases as Regulators of MAPK Signaling. Int. J. Mol. Sci. 2021, 22, 12595. [Google Scholar] [CrossRef]
  189. Santarpia, L.; Lippman, S.M.; El-Naggar, A.K. Targeting the MAPK-RAS-RAF Signaling Pathway in Cancer Therapy. Expert. Opin. Ther. Targets 2012, 16, 103–119. [Google Scholar] [CrossRef] [Green Version]
  190. Guan, Y.; Wu, Q.; Li, M.; Chen, D.; Su, J.; Zuo, L.; Zhu, B.; Li, Y. Epigallocatechin-3-Gallate Induced HepG2 Cells Apoptosis through ROS-Mediated AKT /JNK and P53 Signaling Pathway. Curr. Cancer Drug Targets 2022, 23, 447–460. [Google Scholar] [CrossRef]
  191. Sah, D.K.; Khoi, P.N.; Li, S.; Arjunan, A.; Jeong, J.-U.; Jung, Y.D. (−)-Epigallocatechin-3-Gallate Prevents IL-1β-Induced UPAR Expression and Invasiveness via the Suppression of NF-ΚB and AP-1 in Human Bladder Cancer Cells. Int. J. Mol. Sci. 2022, 23, 14008. [Google Scholar] [CrossRef] [PubMed]
  192. Ersahin, T.; Tuncbag, N.; Cetin-Atalay, R. The PI3K/AKT/MTOR Interactive Pathway. Mol. Biosyst. 2015, 11, 1946–1954. [Google Scholar] [CrossRef]
  193. LoRusso, P.M. Inhibition of the PI3K/AKT/MTOR Pathway in Solid Tumors. J. Clin. Oncol. 2016, 34, 3803. [Google Scholar] [CrossRef] [PubMed]
  194. Li, X.; Dai, D.; Chen, B.; Tang, H.; Xie, X.; Wei, W. Efficacy of PI3K/AKT/MTOR Pathway Inhibitors for the Treatment of Advanced Solid Cancers: A Literature-Based Meta-Analysis of 46 Randomised Control Trials. PLoS ONE 2018, 13, e0192464. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Morgensztern, D.; McLeod, H.L. PI3K/Akt/MTOR Pathway as a Target for Cancer Therapy. Anti-Cancer Drugs 2005, 16, 797–803. [Google Scholar] [CrossRef]
  196. Porta, C.; Paglino, C.; Mosca, A. Targeting PI3K/Akt/MTOR Signaling in Cancer. Front. Oncol. 2014, 4, 64. [Google Scholar] [CrossRef] [Green Version]
  197. Alzahrani, A.S. PI3K/Akt/MTOR Inhibitors in Cancer: At the Bench and Bedside. Semin. Cancer Biol. 2019, 59, 125–132. [Google Scholar] [CrossRef]
  198. Shorning, B.Y.; Dass, M.S.; Smalley, M.J.; Pearson, H.B. The PI3K-AKT-MTOR Pathway and Prostate Cancer: At the Crossroads of AR, MAPK, and WNT Signaling. Int. J. Mol. Sci. 2020, 21, 4507. [Google Scholar] [CrossRef]
  199. Liu, S.; Wang, X.-J.; Liu, Y.; Cui, Y.-F. PI3K/AKT/MTOR Signaling Is Involved in (−)-Epigallocatechin-3-Gallate-Induced Apoptosis of Human Pancreatic Carcinoma Cells. Am. J. Chin. Med. 2013, 41, 629–642. [Google Scholar] [CrossRef]
  200. Ding, F.; Yang, S. Epigallocatechin-3-Gallate Inhibits Proliferation and Triggers Apoptosis in Colon Cancer via the Hedgehog/Phosphoinositide 3-Kinase Pathways. Can. J. Physiol. Pharmacol. 2021, 99, 910–920. [Google Scholar] [CrossRef]
  201. Ferrari, E.; Bettuzzi, S.; Naponelli, V. The Potential of Epigallocatechin Gallate (EGCG) in Targeting Autophagy for Cancer Treatment: A Narrative Review. Int. J. Mol. Sci. 2022, 23, 6075. [Google Scholar] [CrossRef]
  202. Ménard, S.; Castronovo, V.; Tagliabue, E.; Sobel, M.E. New Insights into the Metastasis-Associated 67 KD Laminin Receptor. J. Cell Biochem. 1997, 67, 155–165. [Google Scholar] [CrossRef]
  203. Tanaka, M.; Narumi, K.; Isemura, M.; Abe, M.; Sato, Y.; Abe, T.; Saijo, Y.; Nukiwa, T.; Satoh, K. Expression of the 37-KDa Laminin Binding Protein in Murine Lung Tumor Cell Correlates with Tumor Angiogenesis. Cancer Lett. 2000, 153, 161–168. [Google Scholar] [CrossRef] [PubMed]
  204. Tachibana, H.; Koga, K.; Fujimura, Y.; Yamada, K. A Receptor for Green Tea Polyphenol EGCG. Nat. Struct. Mol. Biol. 2004, 11, 380–381. [Google Scholar] [CrossRef]
  205. Umeda, D.; Yano, S.; Yamada, K.; Tachibana, H. Green Tea Polyphenol Epigallocatechin-3-Gallate Signaling Pathway through 67-KDa Laminin Receptor. J. Biol. Chem. 2008, 283, 3050–3058. [Google Scholar] [CrossRef] [Green Version]
  206. Shen, X.; Zhao, J.; Wang, Q.; Chen, P.; Hong, Y.; He, X.; Chen, D.; Liu, H.; Wang, Y.; Cai, X. The Invasive Potential of Hepatoma Cells Induced by Radiotherapy Is Related to the Activation of Hepatic Stellate Cells and Could Be Inhibited by EGCG through the TLR4 Signaling Pathway. Radiat. Res. 2022, 197, 365–375. [Google Scholar] [CrossRef] [PubMed]
  207. Badana, A.; Chintala, M.; Varikuti, G.; Pudi, N.; Kumari, S.; Kappala, V.R.; Malla, R.R. Lipid Raft Integrity Is Required for Survival of Triple Negative Breast Cancer Cells. J. Breast Cancer 2016, 19, 372–384. [Google Scholar] [CrossRef]
  208. Greenlee, J.D.; Subramanian, T.; Liu, K.; King, M.R. Rafting down the Metastatic Cascade: The Role of Lipid Rafts in Cancer Metastasis, Cell Death, and Clinical Outcomes. Cancer Res. 2021, 81, 5–17. [Google Scholar] [CrossRef]
  209. Mollinedo, F.; Gajate, C. Lipid Rafts as Signaling Hubs in Cancer Cell Survival/Death and Invasion: Implications in Tumor Progression and Therapy. J. Lipid Res. 2020, 61, 611–635. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  210. Vona, R.; Iessi, E.; Matarrese, P. Role of Cholesterol and Lipid Rafts in Cancer Signaling: A Promising Therapeutic Opportunity? Front. Cell Dev. Biol. 2021, 9, 622908. [Google Scholar] [CrossRef]
  211. 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] [PubMed] [Green Version]
  212. Tsukamoto, S.; Huang, Y.; Umeda, D.; Yamada, S.; Yamashita, S.; Kumazoe, M.; Kim, Y.; Murata, M.; Yamada, K.; Tachibana, H. 67-KDa Laminin Receptor-Dependent Protein Phosphatase 2A (PP2A) Activation Elicits Melanoma-Specific Antitumor Activity Overcoming Drug Resistance. J. Biol. Chem. 2014, 289, 32671–32681. [Google Scholar] [CrossRef] [Green Version]
  213. Hardie, D.G. AMPK: A Target for Drugs and Natural Products with Effects on Both Diabetes and Cancer. Diabetes 2013, 62, 2164–2172. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  214. Salminen, A.; Kaarniranta, K. AMP-Activated Protein Kinase (AMPK) Controls the Aging Process via an Integrated Signaling Network. Ageing Res. Rev. 2012, 11, 230–241. [Google Scholar] [CrossRef]
  215. Winder, W.W.; Hardie, D.G. AMP-Activated Protein Kinase, a Metabolic Master Switch: Possible Roles in Type 2 Diabetes. Am. J. Physiol. 1999, 277, E1–E10. [Google Scholar] [CrossRef]
  216. Yang, C.S.; Zhang, J.; Zhang, L.; Huang, J.; Wang, Y. Mechanisms of Body Weight Reduction and Metabolic Syndrome Alleviation by Tea. Mol. Nutr. Food Res. 2016, 60, 160–174. [Google Scholar] [CrossRef] [Green Version]
  217. Mokra, D.; Joskova, M.; Mokry, J. Therapeutic Effects of Green Tea Polyphenol (−)-Epigallocatechin-3-Gallate (EGCG) in Relation to Molecular Pathways Controlling Inflammation, Oxidative Stress, and Apoptosis. Int. J. Mol. Sci. 2023, 24, 340. [Google Scholar] [CrossRef] [PubMed]
  218. Li, W.; Saud, S.M.; Young, M.R.; Chen, G.; Hua, B. Targeting AMPK for Cancer Prevention and Treatment. Oncotarget 2015, 6, 7365–7378. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  219. Schultze, A.; Fiedler, W. Therapeutic Potential and Limitations of New FAK Inhibitors in the Treatment of Cancer. Expert. Opin. Investig. Drugs 2010, 19, 777–788. [Google Scholar] [CrossRef]
  220. Lv, P.-C.; Jiang, A.-Q.; Zhang, W.-M.; Zhu, H.-L. FAK Inhibitors in Cancer, a Patent Review. Expert. Opin. Ther. Pat. 2018, 28, 139–145. [Google Scholar] [CrossRef]
  221. Schaller, M.D. Cellular Functions of FAK Kinases: Insight into Molecular Mechanisms and Novel Functions. J. Cell Sci. 2010, 123, 1007–1013. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  222. Kumazoe, M.; Kadomatsu, M.; Bae, J.; Otsuka, Y.; Fujimura, Y.; Tachibana, H. Src Mediates Epigallocatechin-3-O-Gallate-Elicited Acid Sphingomyelinase Activation. Molecules 2020, 25, 5481. [Google Scholar] [CrossRef] [PubMed]
  223. Sen, T.; Chatterjee, A. Epigallocatechin-3-Gallate (EGCG) Downregulates EGF-Induced MMP-9 in Breast Cancer Cells: Involvement of Integrin Receptor A5β1 in the Process. Eur. J. Nutr. 2011, 50, 465–478. [Google Scholar] [CrossRef] [PubMed]
  224. Liu, J.D.; Chen, S.H.; Lin, C.L.; Tsai, S.H.; Liang, Y.C. Inhibition of Melanoma Growth and Metastasis by Combination with (−)-Epigallocatechin-3-Gallate and Dacarbazine in Mice. J. Cell Biochem. 2001, 83, 631–642. [Google Scholar] [CrossRef]
  225. Jiang, J.; Hui, C.-C. Hedgehog Signaling in Development and Cancer. Dev. Cell 2008, 15, 801–812. [Google Scholar] [CrossRef] [Green Version]
  226. Pasca di Magliano, M.; Hebrok, M. Hedgehog Signalling in Cancer Formation and Maintenance. Nat. Rev. Cancer 2003, 3, 903–911. [Google Scholar] [CrossRef]
  227. Tang, S.-N.; Fu, J.; Nall, D.; Rodova, M.; Shankar, S.; Srivastava, R.K. Inhibition of Sonic Hedgehog Pathway and Pluripotency Maintaining Factors Regulate Human Pancreatic Cancer Stem Cell Characteristics. Int. J. Cancer 2012, 131, 30–40. [Google Scholar] [CrossRef] [Green Version]
  228. Andersson, E.R.; Lendahl, U. Therapeutic Modulation of Notch Signalling—Are We There Yet? Nat. Rev. Drug Discov. 2014, 13, 357–378. [Google Scholar] [CrossRef]
  229. Bray, S.J. Notch Signalling: A Simple Pathway Becomes Complex. Nat. Rev. Mol. Cell Biol. 2006, 7, 678–689. [Google Scholar] [CrossRef] [PubMed]
  230. Purow, B. NOTCH Inhibition as a Promising New Approach to Cancer Therapy. Adv. Exp. Med. Biol. 2012, 727, 305–319. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  231. Farooqi, A.A.; Pinheiro, M.; Granja, A.; Farabegoli, F.; Reis, S.; Attar, R.; Sabitaliyevich, U.Y.; Xu, B.; Ahmad, A. EGCG Mediated Targeting of Deregulated Signaling Pathways and Non-Coding RNAs in Different Cancers: Focus on JAK/STAT, Wnt/β-Catenin, TGF/SMAD, NOTCH, SHH/GLI, and TRAIL Mediated Signaling Pathways. Cancers 2020, 12, 951. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  232. Wei, H.; Ge, Q.; Zhang, L.-Y.; Xie, J.; Gan, R.-H.; Lu, Y.-G.; Zheng, D.-L. EGCG Inhibits Growth of Tumoral Lesions on Lip and Tongue of K-Ras Transgenic Mice through the Notch Pathway. J. Nutr. Biochem. 2022, 99, 108843. [Google Scholar] [CrossRef]
  233. Anastas, J.N.; Moon, R.T. WNT Signalling Pathways as Therapeutic Targets in Cancer. Nat. Rev. Cancer 2013, 13, 11–26. [Google Scholar] [CrossRef] [PubMed]
  234. Zhan, T.; Rindtorff, N.; Boutros, M. Wnt Signaling in Cancer. Oncogene 2017, 36, 1461–1473. [Google Scholar] [CrossRef]
  235. Zhu, J.; Jiang, Y.; Yang, X.; Wang, S.; Xie, C.; Li, X.; Li, Y.; Chen, Y.; Wang, X.; Meng, Y.; et al. Wnt/β-Catenin Pathway Mediates (−)-Epigallocatechin-3-Gallate (EGCG) Inhibition of Lung Cancer Stem Cells. Biochem. Biophys. Res. Commun. 2017, 482, 15–21. [Google Scholar] [CrossRef] [PubMed]
  236. Santos, L.H.S.; Ferreira, R.S.; Caffarena, E.R. Integrating Molecular Docking and Molecular Dynamics Simulations. Methods Mol. Biol. 2019, 2053, 13–34. [Google Scholar] [CrossRef]
  237. Talele, T.T.; Khedkar, S.A.; Rigby, A.C. Successful Applications of Computer Aided Drug Discovery: Moving Drugs from Concept to the Clinic. Curr. Top. Med. Chem. 2010, 10, 127–141. [Google Scholar] [CrossRef]
  238. Bhattacharjee, R.; Devi, A.; Mishra, S. Molecular Docking and Molecular Dynamics Studies Reveal Structural Basis of Inhibition and Selectivity of Inhibitors EGCG and OSU-03012 toward Glucose Regulated Protein-78 (GRP78) Overexpressed in Glioblastoma. J. Mol. Model. 2015, 21, 272. [Google Scholar] [CrossRef]
  239. Wang, J.; Yin, Y.; Hua, H.; Li, M.; Luo, T.; Xu, L.; Wang, R.; Liu, D.; Zhang, Y.; Jiang, Y. Blockade of GRP78 Sensitizes Breast Cancer Cells to Microtubules-Interfering Agents That Induce the Unfolded Protein Response. J. Cell Mol. Med. 2009, 13, 3888–3897. [Google Scholar] [CrossRef] [Green Version]
  240. Luo, T.; Wang, J.; Yin, Y.; Hua, H.; Jing, J.; Sun, X.; Li, M.; Zhang, Y.; Jiang, Y. (−)-Epigallocatechin Gallate Sensitizes Breast Cancer Cells to Paclitaxel in a Murine Model of Breast Carcinoma. Breast Cancer Res. 2010, 12, R8. [Google Scholar] [CrossRef]
  241. Khan, A.; Mohammad, T.; Shamsi, A.; Hussain, A.; Alajmi, M.F.; Husain, S.A.; Iqbal, M.A.; Hassan, M.I. Identification of Plant-Based Hexokinase 2 Inhibitors: Combined Molecular Docking and Dynamics Simulation Studies. J. Biomol. Struct. Dyn. 2022, 40, 10319–10331. [Google Scholar] [CrossRef] [PubMed]
  242. Ciscato, F.; Ferrone, L.; Masgras, I.; Laquatra, C.; Rasola, A. Hexokinase 2 in Cancer: A Prima Donna Playing Multiple Characters. Int. J. Mol. Sci. 2021, 22, 4716. [Google Scholar] [CrossRef] [PubMed]
  243. Gao, F.; Li, M.; Liu, W.-B.; Zhou, Z.-S.; Zhang, R.; Li, J.-L.; Zhou, K.-C. Epigallocatechin Gallate Inhibits Human Tongue Carcinoma Cells via HK2-mediated Glycolysis. Oncol. Rep. 2015, 33, 1533–1539. [Google Scholar] [CrossRef] [Green Version]
  244. Wei, R.; Hackman, R.M.; Wang, Y.; Mackenzie, G.G. Targeting Glycolysis with Epigallocatechin-3-Gallate Enhances the Efficacy of Chemotherapeutics in Pancreatic Cancer Cells and Xenografts. Cancers 2019, 11, 1496. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  245. Gobessi, S.; Laurenti, L.; Longo, P.G.; Sica, S.; Leone, G.; Efremov, D.G. ZAP-70 Enhances B-Cell-Receptor Signaling despite Absent or Inefficient Tyrosine Kinase Activation in Chronic Lymphocytic Leukemia and Lymphoma B Cells. Blood 2007, 109, 2032–2039. [Google Scholar] [CrossRef] [Green Version]
  246. Chen, J.; Moore, A.; Ringshausen, I. ZAP-70 Shapes the Immune Microenvironment in B Cell Malignancies. Front. Oncol. 2020, 10, 595832. [Google Scholar] [CrossRef]
  247. Shim, J.-H.; Choi, H.S.; Pugliese, A.; Lee, S.-Y.; Chae, J.-I.; Choi, B.Y.; Bode, A.M.; Dong, Z. (−)-Epigallocatechin Gallate Regulates CD3-Mediated T Cell Receptor Signaling in Leukemia through the Inhibition of ZAP-70 Kinase. J. Biol. Chem. 2008, 283, 28370–28379. [Google Scholar] [CrossRef] [Green Version]
  248. Wang, W.; Xiong, X.; Li, X.; Zhang, Q.; Yang, W.; Du, L. In Silico Investigation of the Anti-Tumor Mechanisms of Epigallocatechin-3-Gallate. Molecules 2019, 24, 1445. [Google Scholar] [CrossRef] [Green Version]
  249. He, Z.; Tang, F.; Ermakova, S.; Li, M.; Zhao, Q.; Cho, Y.-Y.; Ma, W.-Y.; Choi, H.-S.; Bode, A.M.; Yang, C.S.; et al. Fyn Is a Novel Target of (−)-Epigallocatechin Gallate in the Inhibition of JB6 Cl41 Cell Transformation. Mol. Carcinog. 2008, 47, 172–183. [Google Scholar] [CrossRef] [Green Version]
  250. Ahmed, S.; Rahman, A.; Hasnain, A.; Lalonde, M.; Goldberg, V.M.; Haqqi, T.M. Green Tea Polyphenol Epigallocatechin-3-Gallate Inhibits the IL-1 Beta-Induced Activity and Expression of Cyclooxygenase-2 and Nitric Oxide Synthase-2 in Human Chondrocytes. Free Radic. Biol. Med. 2002, 33, 1097–1105. [Google Scholar] [CrossRef]
  251. Hung, P.-F.; Wu, B.-T.; Chen, H.-C.; Chen, Y.-H.; Chen, C.-L.; Wu, M.-H.; Liu, H.-C.; Lee, M.-J.; Kao, Y.-H. Antimitogenic Effect of Green Tea (−)-Epigallocatechin Gallate on 3T3-L1 Preadipocytes Depends on the ERK and Cdk2 Pathways. Am. J. Physiol. Cell Physiol. 2005, 288, C1094–C1108. [Google Scholar] [CrossRef] [PubMed]
  252. Wu, B.-T.; Hung, P.-F.; Chen, H.-C.; Huang, R.-N.; Chang, H.-H.; Kao, Y.-H. The Apoptotic Effect of Green Tea (−)-Epigallocatechin Gallate on 3T3-L1 Preadipocytes Depends on the Cdk2 Pathway. J. Agric. Food Chem. 2005, 53, 5695–5701. [Google Scholar] [CrossRef] [PubMed]
  253. Chen, H.; Adams, E.; Van Schepdael, A. Study of Abl1 Tyrosine Kinase Inhibitors by Liquid Chromatography-Electrospray Ionization-Mass Spectrometry. Talanta 2013, 107, 88–94. [Google Scholar] [CrossRef] [PubMed]
  254. Deana, R.; Turetta, L.; Donella-Deana, A.; Donà, M.; Brunati, A.M.; De Michiel, L.; Garbisa, S. Green Tea Epigallocatechin-3-Gallate Inhibits Platelet Signalling Pathways Triggered by Both Proteolytic and Non-Proteolytic Agonists. Thromb. Haemost. 2003, 89, 866–874. [Google Scholar] [CrossRef] [Green Version]
  255. Park, S.Y.; Lee, Y.-K.; Kim, Y.-M.; Park, O.J.; Shin, J.-I. Control of AMP-Activated Protein Kinase, Akt, and MTOR in EGCG-Treated HT-29 Colon Cancer Cells. Food Sci. Biotechnol. 2013, 22, 147–151. [Google Scholar] [CrossRef]
  256. Park, S.B.; Bae, J.W.; Kim, J.M.; Lee, S.G.; Han, M. Antiproliferative and Apoptotic Effect of Epigallocatechin-3-Gallate on Ishikawa Cells Is Accompanied by Sex Steroid Receptor Downregulation. Int. J. Mol. Med. 2012, 30, 1211–1218. [Google Scholar] [CrossRef] [Green Version]
  257. Singh, A.K.; Umar, S.; Riegsecker, S.; Chourasia, M.; Ahmed, S. Regulation of Transforming Growth Factor β-Activated Kinase Activation by Epigallocatechin-3-Gallate in Rheumatoid Arthritis Synovial Fibroblasts: Suppression of K(63) -Linked Autoubiquitination of Tumor Necrosis Factor Receptor-Associated Factor 6. Arthritis Rheumatol. 2016, 68, 347–358. [Google Scholar] [CrossRef] [Green Version]
  258. Wu, P.-P.; Kuo, S.-C.; Huang, W.-W.; Yang, J.-S.; Lai, K.-C.; Chen, H.-J.; Lin, K.-L.; Chiu, Y.-J.; Huang, L.-J.; Chung, J.-G. (−)-Epigallocatechin Gallate Induced Apoptosis in Human Adrenal Cancer NCI-H295 Cells through Caspase-Dependent and Caspase-Independent Pathway. Anticancer Res. 2009, 29, 1435–1442. [Google Scholar]
  259. Matheson, C.J.; Backos, D.S.; Reigan, P. Targeting WEE1 Kinase in Cancer. Trends Pharmacol. Sci. 2016, 37, 872–881. [Google Scholar] [CrossRef]
  260. Zhu, G.; Pei, L.; Xia, H.; Tang, Q.; Bi, F. Role of Oncogenic KRAS in the Prognosis, Diagnosis and Treatment of Colorectal Cancer. Mol. Cancer 2021, 20, 143. [Google Scholar] [CrossRef]
  261. Huang, L.; Guo, Z.; Wang, F.; Fu, L. KRAS Mutation: From Undruggable to Druggable in Cancer. Signal Transduct. Target. Ther. 2021, 6, 386. [Google Scholar] [CrossRef] [PubMed]
  262. Pierotti, M.A.; Greco, A. Oncogenic Rearrangements of the NTRK1/NGF Receptor. Cancer Lett. 2006, 232, 90–98. [Google Scholar] [CrossRef]
  263. Cui, F.; Yang, K.; Li, Y. Investigate the Binding of Catechins to Trypsin Using Docking and Molecular Dynamics Simulation. PLoS ONE 2015, 10, e0125848. [Google Scholar] [CrossRef] [PubMed]
  264. Smith, D.M.; Daniel, K.G.; Wang, Z.; Guida, W.C.; Chan, T.-H.; Dou, Q.P. Docking Studies and Model Development of Tea Polyphenol Proteasome Inhibitors: Applications to Rational Drug Design. Proteins 2004, 54, 58–70. [Google Scholar] [CrossRef] [PubMed]
  265. Kuban-Jankowska, A.; Kostrzewa, T.; Musial, C.; Barone, G.; Lo-Bosco, G.; Lo-Celso, F.; Gorska-Ponikowska, M. Green Tea Catechins Induce Inhibition of PTP1B Phosphatase in Breast Cancer Cells with Potent Anti-Cancer Properties: In Vitro Assay, Molecular Docking, and Dynamics Studies. Antioxidants 2020, 9, 1208. [Google Scholar] [CrossRef]
  266. Artali, R.; Beretta, G.; Morazzoni, P.; Bombardelli, E.; Meneghetti, F. Green Tea Catechins in Chemoprevention of Cancer: A Molecular Docking Investigation into Their Interaction with Glutathione S-Transferase (GST P1-1). J. Enzyme Inhib. Med. Chem. 2009, 24, 287–295. [Google Scholar] [CrossRef]
  267. Fang, M.Z.; Wang, Y.; Ai, N.; Hou, Z.; Sun, Y.; Lu, H.; Welsh, W.; Yang, C.S. Tea Polyphenol (−)-Epigallocatechin-3-Gallate Inhibits DNA Methyltransferase and Reactivates Methylation-Silenced Genes in Cancer Cell Lines. Cancer Res. 2003, 63, 7563–7570. [Google Scholar]
  268. Rahnasto-Rilla, M.; Tyni, J.; Huovinen, M.; Jarho, E.; Kulikowicz, T.; Ravichandran, S.; Bohr, V.A.; Ferrucci, L.; Lahtela-Kakkonen, M.; Moaddel, R. Natural Polyphenols as Sirtuin 6 Modulators. Sci. Rep. 2018, 8, 4163. [Google Scholar] [CrossRef] [Green Version]
  269. Saeki, K.; Hayakawa, S.; Nakano, S.; Ito, S.; Oishi, Y.; Suzuki, Y.; Isemura, M. In Vitro and In Silico Studies of the Molecular Interactions of Epigallocatechin-3-O-Gallate (EGCG) with Proteins That Explain the Health Benefits of Green Tea. Molecules 2018, 23, 1295. [Google Scholar] [CrossRef] [Green Version]
  270. Nakano, S.; Megro, S.; Hase, T.; Suzuki, T.; Isemura, M.; Nakamura, Y.; Ito, S. Computational Molecular Docking and X-Ray Crystallographic Studies of Catechins in New Drug Design Strategies. Molecules 2018, 23, 2020. [Google Scholar] [CrossRef] [Green Version]
  271. Yuan, J.-M.; Sun, C.; Butler, L.M. Tea and Cancer Prevention: Epidemiological Studies. Pharmacol. Res. 2011, 64, 123–135. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  272. Zhang, Y.-F.; Xu, Q.; Lu, J.; Wang, P.; Zhang, H.-W.; Zhou, L.; Ma, X.-Q.; Zhou, Y.-H. Tea Consumption and the Incidence of Cancer: A Systematic Review and Meta-Analysis of Prospective Observational Studies. Eur. J. Cancer Prev. 2015, 24, 353–362. [Google Scholar] [CrossRef] [PubMed]
  273. Shin, S.; Lee, J.E.; Loftfield, E.; Shu, X.-O.; Abe, S.K.; Rahman, M.S.; Saito, E.; Islam, M.R.; Tsugane, S.; Sawada, N.; et al. Coffee and Tea Consumption and Mortality from All Causes, Cardiovascular Disease and Cancer: A Pooled Analysis of Prospective Studies from the Asia Cohort Consortium. Int. J. Epidemiol. 2022, 51, 626–640. [Google Scholar] [CrossRef] [PubMed]
  274. Kim, H.; Lee, J.; Oh, J.H.; Chang, H.J.; Sohn, D.K.; Shin, A.; Kim, J. Protective Effect of Green Tea Consumption on Colorectal Cancer Varies by Lifestyle Factors. Nutrients 2019, 11, 2612. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  275. Zheng, P.; Zheng, H.-M.; Deng, X.-M.; Zhang, Y. Green Tea Consumption and Risk of Esophageal Cancer: A Meta-Analysis of Epidemiologic Studies. BMC Gastroenterol. 2012, 12, 165. [Google Scholar] [CrossRef] [Green Version]
  276. Sang, L.-X.; Chang, B.; Li, X.-H.; Jiang, M. Green Tea Consumption and Risk of Esophageal Cancer: A Meta-Analysis of Published Epidemiological Studies. Nutr. Cancer 2013, 65, 802–812. [Google Scholar] [CrossRef]
  277. Zhou, Y.; Li, N.; Zhuang, W.; Liu, G.; Wu, T.; Yao, X.; Du, L.; Wei, M.; Wu, X. Green Tea and Gastric Cancer Risk: Meta-Analysis of Epidemiologic Studies. Asia Pac. J. Clin. Nutr. 2008, 17, 159–165. [Google Scholar]
  278. Meshitsuka, S.; Shingaki, S.; Hotta, M.; Goto, M.; Kobayashi, M.; Ukawa, Y.; Sagesaka, Y.M.; Wada, Y.; Nojima, M.; Suzuki, K. Phase 2 Trial of Daily, Oral Epigallocatechin Gallate in Patients with Light-Chain Amyloidosis. Int. J. Hematol. 2017, 105, 295–308. [Google Scholar] [CrossRef]
  279. Widmer, R.J.; Freund, M.A.; Flammer, A.J.; Sexton, J.; Lennon, R.; Romani, A.; Mulinacci, N.; Vinceri, F.F.; Lerman, L.O.; Lerman, A. Beneficial Effects of Polyphenol-Rich Olive Oil in Patients with Early Atherosclerosis. Eur. J. Nutr. 2013, 52, 1223–1231. [Google Scholar] [CrossRef] [Green Version]
  280. De Stefano, D.; Villella, V.R.; Esposito, S.; Tosco, A.; Sepe, A.; De Gregorio, F.; Salvadori, L.; Grassia, R.; Leone, C.A.; De Rosa, G.; et al. Restoration of CFTR Function in Patients with Cystic Fibrosis Carrying the F508del-CFTR Mutation. Autophagy 2014, 10, 2053–2074. [Google Scholar] [CrossRef] [Green Version]
  281. Tosco, A.; De Gregorio, F.; Esposito, S.; De Stefano, D.; Sana, I.; Ferrari, E.; Sepe, A.; Salvadori, L.; Buonpensiero, P.; Di Pasqua, A.; et al. A Novel Treatment of Cystic Fibrosis Acting On-Target: Cysteamine plus Epigallocatechin Gallate for the Autophagy-Dependent Rescue of Class II-Mutated CFTR. Cell Death Differ. 2016, 23, 1380–1393. [Google Scholar] [CrossRef] [Green Version]
  282. Joe, A.K.; Schnoll-Sussman, F.; Bresalier, R.S.; Abrams, J.A.; Hibshoosh, H.; Cheung, K.; Friedman, R.A.; Yang, C.S.; Milne, G.L.; Liu, D.D.; et al. Phase Ib Randomized, Double-Blinded, Placebo-Controlled, Dose Escalation Study of Polyphenon E in Patients with Barrett’s Esophagus. Cancer Prev. Res. 2015, 8, 1131–1137. [Google Scholar] [CrossRef] [Green Version]
  283. Huang, S.-M.; Chang, Y.-H.; Chao, Y.-C.; Lin, J.-A.; Wu, C.-H.; Lai, C.-Y.; Chan, K.-C.; Tseng, S.-T.; Yen, G.-C. EGCG-Rich Green Tea Extract Stimulates SRAGE Secretion to Inhibit S100A12-RAGE Axis through ADAM10-Mediated Ectodomain Shedding of Extracellular RAGE in Type 2 Diabetes. Mol. Nutr. Food Res. 2013, 57, 2264–2268. [Google Scholar] [CrossRef] [PubMed]
  284. Zhang, H.; Su, S.; Yu, X.; Li, Y. Dietary Epigallocatechin 3-Gallate Supplement Improves Maternal and Neonatal Treatment Outcome of Gestational Diabetes Mellitus: A Double-Blind Randomised Controlled Trial. J. Hum. Nutr. Diet. 2017, 30, 753–758. [Google Scholar] [CrossRef] [PubMed]
  285. Bazyar, H.; Hosseini, S.A.; Saradar, S.; Mombaini, D.; Allivand, M.; Labibzadeh, M.; Alipour, M. Effects of Epigallocatechin-3-Gallate of Camellia sinensis Leaves on Blood Pressure, Lipid Profile, Atherogenic Index of Plasma and Some Inflammatory and Antioxidant Markers in Type 2 Diabetes Mellitus Patients: A Clinical Trial. J. Complement. Integr. Med. 2020, 18, 405–411. [Google Scholar] [CrossRef] [PubMed]
  286. De la Torre, R.; De Sola, S.; Pons, M.; Duchon, A.; de Lagran, M.M.; Farré, M.; Fitó, M.; Benejam, B.; Langohr, K.; Rodriguez, J.; et al. Epigallocatechin-3-Gallate, a DYRK1A Inhibitor, Rescues Cognitive Deficits in Down Syndrome Mouse Models and in Humans. Mol. Nutr. Food Res. 2014, 58, 278–288. [Google Scholar] [CrossRef]
  287. de la Torre, R.; de Sola, S.; Hernandez, G.; Farré, M.; Pujol, J.; Rodriguez, J.; Espadaler, J.M.; Langohr, K.; Cuenca-Royo, A.; Principe, A.; et al. Safety and Efficacy of Cognitive Training plus Epigallocatechin-3-Gallate in Young Adults with Down’s Syndrome (TESDAD): A Double-Blind, Randomised, Placebo-Controlled, Phase 2 Trial. Lancet Neurol. 2016, 15, 801–810. [Google Scholar] [CrossRef]
  288. Xicota, L.; Rodríguez, J.; Langohr, K.; Fitó, M.; Dierssen, M.; de la Torre, R.; TESDAD Study Group. Effect of Epigallocatechin Gallate on the Body Composition and Lipid Profile of down Syndrome Individuals: Implications for Clinical Management. Clin. Nutr. 2020, 39, 1292–1300. [Google Scholar] [CrossRef]
  289. Cieuta-Walti, C.; Cuenca-Royo, A.; Langohr, K.; Rakic, C.; López-Vílchez, M.Á.; Lirio, J.; González-Lamuño Leguina, D.; González, T.B.; García, J.G.; Roure, M.R.; et al. Safety and Preliminary Efficacy on Cognitive Performance and Adaptive Functionality of Epigallocatechin Gallate (EGCG) in Children with Down Syndrome. A Randomized Phase Ib Clinical Trial (PERSEUS Study). Genet. Med. 2022, 24, 2004–2013. [Google Scholar] [CrossRef]
  290. Chiaverini, C.; Roger, C.; Fontas, E.; Bourrat, E.; Bourdon-Lanoy, E.; Labrèze, C.; Mazereeuw, J.; Vabres, P.; Bodemer, C.; Lacour, J.-P. Oral Epigallocatechin-3-Gallate for Treatment of Dystrophic Epidermolysis Bullosa: A Multicentre, Randomized, Crossover, Double-Blind, Placebo-Controlled Clinical Trial. Orphanet J. Rare Dis. 2016, 11, 31. [Google Scholar] [CrossRef] [Green Version]
  291. Arazi, H.; Samami, N.; Kheirkhah, J.; Taati, B. The Effect of Three Weeks Green Tea Extract Consumption on Blood Pressure, Heart Rate Responses to a Single Bout Resistance Exercise in Hypertensive Women. High Blood Press. Cardiovasc. Prev. 2014, 21, 213–219. [Google Scholar] [CrossRef]
  292. Maeda-Yamamoto, M.; Nishimura, M.; Kitaichi, N.; Nesumi, A.; Monobe, M.; Nomura, S.; Horie, Y.; Tachibana, H.; Nishihira, J. A Randomized, Placebo-Controlled Study on the Safety and Efficacy of Daily Ingestion of Green Tea (Camellia sinensis L.) Cv. “Yabukita” and “Sunrouge” on Eyestrain and Blood Pressure in Healthy Adults. Nutrients 2018, 10, 569. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  293. Mähler, A.; Steiniger, J.; Bock, M.; Klug, L.; Parreidt, N.; Lorenz, M.; Zimmermann, B.F.; Krannich, A.; Paul, F.; Boschmann, M. Metabolic Response to Epigallocatechin-3-Gallate in Relapsing-Remitting Multiple Sclerosis: A Randomized Clinical Trial. Am. J. Clin. Nutr. 2015, 101, 487–495. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  294. Lovera, J.; Ramos, A.; Devier, D.; Garrison, V.; Kovner, B.; Reza, T.; Koop, D.; Rooney, W.; Foundas, A.; Bourdette, D. Polyphenon E, Non-Futile at Neuroprotection in Multiple Sclerosis but Unpredictably Hepatotoxic: Phase I Single Group and Phase II Randomized Placebo-Controlled Studies. J. Neurol. Sci. 2015, 358, 46–52. [Google Scholar] [CrossRef] [Green Version]
  295. Platero, J.L.; Cuerda-Ballester, M.; Ibáñez, V.; Sancho, D.; Lopez-Rodríguez, M.M.; Drehmer, E.; de la Rubia Ortí, J.E. The Impact of Coconut Oil and Epigallocatechin Gallate on the Levels of IL-6, Anxiety and Disability in Multiple Sclerosis Patients. Nutrients 2020, 12, 305. [Google Scholar] [CrossRef] [Green Version]
  296. Benlloch, M.; Cuerda Ballester, M.; Drehmer, E.; Platero, J.L.; Carrera-Juliá, S.; López-Rodríguez, M.M.; Ceron, J.J.; Tvarijonaviciute, A.; Navarro, M.Á.; Moreno, M.L.; et al. Possible Reduction of Cardiac Risk after Supplementation with Epigallocatechin Gallate and Increase of Ketone Bodies in the Blood in Patients with Multiple Sclerosis. A Pilot Study. Nutrients 2020, 12, 3792. [Google Scholar] [CrossRef] [PubMed]
  297. Rust, R.; Chien, C.; Scheel, M.; Brandt, A.U.; Dörr, J.; Wuerfel, J.; Klumbies, K.; Zimmermann, H.; Lorenz, M.; Wernecke, K.-D.; et al. Epigallocatechin Gallate in Progressive MS: A Randomized, Placebo-Controlled Trial. Neurol. Neuroimmunol. Neuroinflamm. 2021, 8, e964. [Google Scholar] [CrossRef] [PubMed]
  298. Bellmann-Strobl, J.; Paul, F.; Wuerfel, J.; Dörr, J.; Infante-Duarte, C.; Heidrich, E.; Körtgen, B.; Brandt, A.; Pfüller, C.; Radbruch, H.; et al. Epigallocatechin Gallate in Relapsing-Remitting Multiple Sclerosis: A Randomized, Placebo-Controlled Trial. Neurol. Neuroimmunol. Neuroinflamm. 2021, 8, e981. [Google Scholar] [CrossRef]
  299. de la Rubia Ortí, J.E.; Platero, J.L.; Yang, I.H.; Ceron, J.J.; Tvarijonaviciute, A.; Sabater, P.S.; Benlloch, M.; Sancho-Cantus, D.; Sancho, S. Possible Role of Butyrylcholinesterase in Fat Loss and Decreases in Inflammatory Levels in Patients with Multiple Sclerosis after Treatment with Epigallocatechin Gallate and Coconut Oil: A Pilot Study. Nutrients 2021, 13, 3230. [Google Scholar] [CrossRef]
  300. Cuerda-Ballester, M.; Proaño, B.; Alarcón-Jimenez, J.; de Bernardo, N.; Villaron-Casales, C.; Lajara Romance, J.M.; de la Rubia Ortí, J.E. Improvements in Gait and Balance in Patients with Multiple Sclerosis after Treatment with Coconut Oil and Epigallocatechin Gallate. A Pilot Study. Food Funct. 2023, 14, 1062–1071. [Google Scholar] [CrossRef]
  301. Mielgo-Ayuso, J.; Barrenechea, L.; Alcorta, P.; Larrarte, E.; Margareto, J.; Labayen, I. Effects of Dietary Supplementation with Epigallocatechin-3-Gallate on Weight Loss, Energy Homeostasis, Cardiometabolic Risk Factors and Liver Function in Obese Women: Randomised, Double-Blind, Placebo-Controlled Clinical Trial. Br. J. Nutr. 2014, 111, 1263–1271. [Google Scholar] [CrossRef] [Green Version]
  302. Chen, I.-J.; Liu, C.-Y.; Chiu, J.-P.; Hsu, C.-H. Therapeutic Effect of High-Dose Green Tea Extract on Weight Reduction: A Randomized, Double-Blind, Placebo-Controlled Clinical Trial. Clin. Nutr. 2016, 35, 592–599. [Google Scholar] [CrossRef] [PubMed]
  303. Most, J.; Timmers, S.; Warnke, I.; Jocken, J.W.; van Boekschoten, M.; de Groot, P.; Bendik, I.; Schrauwen, P.; Goossens, G.H.; Blaak, E.E. Combined Epigallocatechin-3-Gallate and Resveratrol Supplementation for 12 Wk Increases Mitochondrial Capacity and Fat Oxidation, but Not Insulin Sensitivity, in Obese Humans: A Randomized Controlled Trial. Am. J. Clin. Nutr. 2016, 104, 215–227. [Google Scholar] [CrossRef] [Green Version]
  304. Most, J.; Warnke, I.; Boekschoten, M.V.; Jocken, J.W.E.; de Groot, P.; Friedel, A.; Bendik, I.; Goossens, G.H.; Blaak, E.E. The Effects of Polyphenol Supplementation on Adipose Tissue Morphology and Gene Expression in Overweight and Obese Humans. Adipocyte 2018, 7, 190–196. [Google Scholar] [CrossRef] [PubMed]
  305. Cazzola, R.; Rondanelli, M. N-Oleoyl-Phosphatidyl-Ethanolamine and Epigallo Catechin-3-Gallate Mitigate Oxidative Stress in Overweight and Class I Obese People on a Low-Calorie Diet. J. Med. Food 2020, 23, 319–325. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  306. Chatree, S.; Sitticharoon, C.; Maikaew, P.; Pongwattanapakin, K.; Keadkraichaiwat, I.; Churintaraphan, M.; Sripong, C.; Sririwichitchai, R.; Tapechum, S. Epigallocatechin Gallate Decreases Plasma Triglyceride, Blood Pressure, and Serum Kisspeptin in Obese Human Subjects. Exp. Biol. Med. 2021, 246, 163–176. [Google Scholar] [CrossRef]
  307. Roberts, J.D.; Willmott, A.G.B.; Beasley, L.; Boal, M.; Davies, R.; Martin, L.; Chichger, H.; Gautam, L.; Del Coso, J. The Impact of Decaffeinated Green Tea Extract on Fat Oxidation, Body Composition and Cardio-Metabolic Health in Overweight, Recreationally Active Individuals. Nutrients 2021, 13, 764. [Google Scholar] [CrossRef]
  308. Gu, Q.; Wang, X.; Xie, L.; Yao, X.; Qian, L.; Yu, Z.; Shen, X. Green Tea Catechin EGCG Could Prevent Obesity-Related Precocious Puberty through NKB/NK3R Signaling Pathway. J. Nutr. Biochem. 2022, 108, 109085. [Google Scholar] [CrossRef]
  309. Zhao, H.; Zhu, W.; Xie, P.; Li, H.; Zhang, X.; Sun, X.; Yu, J.; Xing, L. A Phase I Study of Concurrent Chemotherapy and Thoracic Radiotherapy with Oral Epigallocatechin-3-Gallate Protection in Patients with Locally Advanced Stage III Non-Small-Cell Lung Cancer. Radiother. Oncol. 2014, 110, 132–136. [Google Scholar] [CrossRef]
  310. Zhao, H.; Xie, P.; Li, X.; Zhu, W.; Sun, X.; Sun, X.; Chen, X.; Xing, L.; Yu, J. A Prospective Phase II Trial of EGCG in Treatment of Acute Radiation-Induced Esophagitis for Stage III Lung Cancer. Radiother. Oncol. 2015, 114, 351–356. [Google Scholar] [CrossRef]
  311. Zhao, H.; Jia, L.; Chen, G.; Li, X.; Meng, X.; Zhao, X.; Xing, L.; Zhu, W. A Prospective, Three-Arm, Randomized Trial of EGCG for Preventing Radiation-Induced Esophagitis in Lung Cancer Patients Receiving Radiotherapy. Radiother. Oncol. 2019, 137, 186–191. [Google Scholar] [CrossRef] [PubMed]
  312. Li, X.; Xing, L.; Zhang, Y.; Xie, P.; Zhu, W.; Meng, X.; Wang, Y.; Kong, L.; Zhao, H.; Yu, J. Phase II Trial of Epigallocatechin-3-Gallate in Acute Radiation-Induced Esophagitis for Esophagus Cancer. J. Med. Food 2020, 23, 43–49. [Google Scholar] [CrossRef] [PubMed]
  313. Dryden, G.W.; Lam, A.; Beatty, K.; Qazzaz, H.H.; McClain, C.J. A Pilot Study to Evaluate the Safety and Efficacy of an Oral Dose of (−)-Epigallocatechin-3-Gallate-Rich Polyphenon E in Patients with Mild to Moderate Ulcerative Colitis. Inflamm. Bowel Dis. 2013, 19, 1904–1912. [Google Scholar] [CrossRef]
  314. Kowalska, J.; Marzec, A.; Domian, E.; Galus, S.; Ciurzyńska, A.; Brzezińska, R.; Kowalska, H. Influence of Tea Brewing Parameters on the Antioxidant Potential of Infusions and Extracts Depending on the Degree of Processing of the Leaves of Camellia sinensis. Molecules 2021, 26, 4773. [Google Scholar] [CrossRef]
  315. Lakenbrink, C.; Lapczynski, S.; Maiwald, B.; Engelhardt, U.H. Flavonoids and Other Polyphenols in Consumer Brews of Tea and Other Caffeinated Beverages. J. Agric. Food Chem. 2000, 48, 2848–2852. [Google Scholar] [CrossRef] [PubMed]
  316. Hu, J.; Zhou, D.; Chen, Y. Preparation and Antioxidant Activity of Green Tea Extract Enriched in Epigallocatechin (EGC) and Epigallocatechin Gallate (EGCG). J. Agric. Food Chem. 2009, 57, 1349–1353. [Google Scholar] [CrossRef]
  317. Saklar, S.; Ertas, E.; Ozdemir, I.S.; Karadeniz, B. Effects of Different Brewing Conditions on Catechin Content and Sensory Acceptance in Turkish Green Tea Infusions. J. Food Sci. Technol. 2015, 52, 6639–6646. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  318. Liu, M.; Xu, Y.-F.; Feng, Y.; Yang, F.-Q.; Luo, J.; Zhai, W.; Che, J.-P.; Wang, G.-C.; Zheng, J.-H. Epigallocatechin Gallate Attenuates Interstitial Cystitis in Human Bladder Urothelium Cells by Modulating Purinergic Receptors. J. Surg. Res. 2013, 183, 397–404. [Google Scholar] [CrossRef] [PubMed]
  319. Trudel, D.; Labbé, D.P.; Araya-Farias, M.; Doyen, A.; Bazinet, L.; Duchesne, T.; Plante, M.; Grégoire, J.; Renaud, M.-C.; Bachvarov, D.; et al. A Two-Stage, Single-Arm, Phase II Study of EGCG-Enriched Green Tea Drink as a Maintenance Therapy in Women with Advanced Stage Ovarian Cancer. Gynecol. Oncol. 2013, 131, 357–361. [Google Scholar] [CrossRef]
  320. Kumar, N.B.; Pow-Sang, J.; Egan, K.M.; Spiess, P.E.; Dickinson, S.; Salup, R.; Helal, M.; McLarty, J.; Williams, C.R.; Schreiber, F.; et al. Randomized, Placebo-Controlled Trial of Green Tea Catechins for Prostate Cancer Prevention. Cancer Prev. Res. 2015, 8, 879–887. [Google Scholar] [CrossRef] [Green Version]
  321. Tsampoukas, G.; Manolas, V.; Brown, D.; Dellis, A.; Deliveliotis, K.; Moussa, M.; Papatsoris, A. Atypical Small Acinar Proliferation and Its Significance in Pathological Reports in Modern Urological Times. Asian J. Urol. 2022, 9, 12–17. [Google Scholar] [CrossRef] [PubMed]
  322. Andras, I.; Telecan, T.; Crisan, D.; Cata, E.; Kadula, P.; Andras, D.; Bungardean, M.; Coman, I.; Crisan, N. Different Clinical Significance of ASAP/HGPIN Pattern in Systematic vs. MRI-US Fusion Guided Prostate Biopsy. Exp. Ther. Med. 2020, 20, 195. [Google Scholar] [CrossRef]
  323. Bettuzzi, S.; Brausi, M.; Rizzi, F.; Castagnetti, G.; Peracchia, G.; Corti, A. Chemoprevention of Human Prostate Cancer by Oral Administration of Green Tea Catechins in Volunteers with High-Grade Prostate Intraepithelial Neoplasia: A Preliminary Report from a One-Year Proof-of-Principle Study. Cancer Res. 2006, 66, 1234–1240. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  324. Brausi, M.; Rizzi, F.; Bettuzzi, S. Chemoprevention of Human Prostate Cancer by Green Tea Catechins: Two Years Later. A Follow-up Update. Eur. Urol. 2008, 54, 472–473. [Google Scholar] [CrossRef] [PubMed]
  325. Kumar, N.B.; Pow-Sang, J.; Spiess, P.E.; Park, J.; Salup, R.; Williams, C.R.; Parnes, H.; Schell, M.J. Randomized, Placebo-Controlled Trial Evaluating the Safety of One-Year Administration of Green Tea Catechins. Oncotarget 2016, 7, 70794–70802. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  326. Perletti, G.; Magri, V.; Vral, A.; Stamatiou, K.; Trinchieri, A. Green Tea Catechins for Chemoprevention of Prostate Cancer in Patients with Histologically-Proven HG-PIN or ASAP. Concise Review and Meta-Analysis. Arch. Ital. Urol. Androl. 2019, 91. [Google Scholar] [CrossRef] [Green Version]
  327. Henning, S.M.; Aronson, W.; Niu, Y.; Conde, F.; Lee, N.H.; Seeram, N.P.; Lee, R.-P.; Lu, J.; Harris, D.M.; Moro, A.; et al. Tea Polyphenols and Theaflavins Are Present in Prostate Tissue of Humans and Mice after Green and Black Tea Consumption. J. Nutr. 2006, 136, 1839–1843. [Google Scholar] [CrossRef] [Green Version]
  328. Wang, P.; Aronson, W.J.; Huang, M.; Zhang, Y.; Lee, R.-P.; Heber, D.; Henning, S.M. Green Tea Polyphenols and Metabolites in Prostatectomy Tissue: Implications for Cancer Prevention. Cancer Prev. Res. 2010, 3, 985–993. [Google Scholar] [CrossRef] [Green Version]
  329. McLarty, J.; Bigelow, R.L.H.; Smith, M.; Elmajian, D.; Ankem, M.; Cardelli, J.A. Tea Polyphenols Decrease Serum Levels of Prostate-Specific Antigen, Hepatocyte Growth Factor, and Vascular Endothelial Growth Factor in Prostate Cancer Patients and Inhibit Production of Hepatocyte Growth Factor and Vascular Endothelial Growth Factor in Vitro. Cancer Prev. Res. 2009, 2, 673–682. [Google Scholar] [CrossRef] [Green Version]
  330. Fhu, C.W.; Ali, A. Fatty Acid Synthase: An Emerging Target in Cancer. Molecules 2020, 25, 3935. [Google Scholar] [CrossRef]
  331. Sun, X.; Kaufman, P.D. Ki-67: More than a Proliferation Marker. Chromosoma 2018, 127, 175–186. [Google Scholar] [CrossRef] [PubMed]
  332. Zhang, Z.; Garzotto, M.; Beer, T.M.; Thuillier, P.; Lieberman, S.; Mori, M.; Stoller, W.A.; Farris, P.E.; Shannon, J. Effects of ω-3 Fatty Acids and Catechins on Fatty Acid Synthase in the Prostate: A Randomized Controlled Trial. Nutr. Cancer 2016, 68, 1309–1319. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  333. Hu, Y.; McIntosh, G.H.; Le Leu, R.K.; Somashekar, R.; Meng, X.Q.; Gopalsamy, G.; Bambaca, L.; McKinnon, R.A.; Young, G.P. Supplementation with Brazil Nuts and Green Tea Extract Regulates Targeted Biomarkers Related to Colorectal Cancer Risk in Humans. Br. J. Nutr. 2016, 116, 1901–1911. [Google Scholar] [CrossRef]
  334. Zhang, X.; Zhang, H.; Tighiouart, M.; Lee, J.E.; Shin, H.J.; Khuri, F.R.; Yang, C.S.; Chen, Z.; Shin, D.M. Synergistic Inhibition of Head and Neck Tumor Growth by Green Tea (−)-Epigallocatechin-3-Gallate and EGFR Tyrosine Kinase Inhibitor. Int. J. Cancer 2008, 123, 1005–1014. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  335. Shin, D.M.; Nannapaneni, S.; Patel, M.R.; Shi, Q.; Liu, Y.; Chen, Z.; Chen, A.Y.; El-Deiry, M.W.; Beitler, J.J.; Steuer, C.E.; et al. Phase Ib Study of Chemoprevention with Green Tea Polyphenon E and Erlotinib in Patients with Advanced Premalignant Lesions (APL) of the Head and Neck. Clin. Cancer Res. 2020, 26, 5860–5868. [Google Scholar] [CrossRef] [PubMed]
  336. Thomford, N.E.; Senthebane, D.A.; Rowe, A.; Munro, D.; Seele, P.; Maroyi, A.; Dzobo, K. Natural Products for Drug Discovery in the 21st Century: Innovations for Novel Drug Discovery. Int. J. Mol. Sci. 2018, 19, 1578. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  337. Dube, A.; Nicolazzo, J.A.; Larson, I. Chitosan Nanoparticles Enhance the Plasma Exposure of (−)-Epigallocatechin Gallate in Mice through an Enhancement in Intestinal Stability. Eur. J. Pharm. Sci. Off. J. Eur. Fed. Pharm. Sci. 2011, 44, 422–426. [Google Scholar] [CrossRef]
  338. Son, Y.-R.; Chung, J.-H.; Ko, S.; Shim, S.-M. Combinational Enhancing Effects of Formulation and Encapsulation on Digestive Stability and Intestinal Transport of Green Tea Catechins. J. Microencapsul. 2016, 33, 183–190. [Google Scholar] [CrossRef]
  339. Chung, J.-H.; Lee, S.-J.; Chung, J.-O.; Oh, Y.-J.; Hwang, J.-A.; Kim, Y.-K.; Ko, S.; Shim, S.-M. Effect of Hydroxypropyl Methyl Cellulose Phthalate Coating on Digestive Stability and Intestinal Transport of Green Tea Catechins. Integr. Med. Res. 2014, 3, 34–37. [Google Scholar] [CrossRef] [Green Version]
  340. Huang, Y.B.; Tsai, M.J.; Wu, P.C.; Tsai, Y.H.; Wu, Y.H.; Fang, J.Y. Elastic Liposomes as Carriers for Oral Delivery and the Brain Distribution of (+)-Catechin. J. Drug Target. 2011, 19, 709–718. [Google Scholar] [CrossRef]
  341. Fang, J.Y.; Lee, W.R.; Shen, S.C.; Huang, Y.L. Effect of Liposome Encapsulation of Tea Catechins on Their Accumulation in Basal Cell Carcinomas. J. Dermatol. Sci. 2006, 42, 101–109. [Google Scholar] [CrossRef] [PubMed]
  342. Zou, L.Q.; Liu, W.; Liu, W.L.; Liang, R.H.; Li, T.; Liu, C.M.; Cao, Y.L.; Niu, J.; Liu, Z. Characterization and Bioavailability of Tea Polyphenol Nanoliposome Prepared by Combining an Ethanol Injection Method with Dynamic High-Pressure Microfluidization. J. Agric. Food Chem. 2014, 62, 934–941. [Google Scholar] [CrossRef] [PubMed]
  343. Shi, M.; Huang, L.-Y.; Nie, N.; Ye, J.-H.; Zheng, X.-Q.; Lu, J.-L.; Liang, Y.-R. Binding of Tea Catechins to Rice Bran Protein Isolate: Interaction and Protective Effect during in Vitro Digestion. Food Res. Int. 2017, 93, 1–7. [Google Scholar] [CrossRef]
  344. Li, Y.; Lim, L.-T.; Kakuda, Y. Electrospun Zein Fibers as Carriers to Stabilize (−)-Epigallocatechin Gallate. J. Food Sci. 2009, 74, C233–C240. [Google Scholar] [CrossRef] [PubMed]
  345. Li, Z.; Ha, J.; Zou, T.; Gu, L. Fabrication of Coated Bovine Serum Albumin (BSA)-Epigallocatechin Gallate (EGCG) Nanoparticles and Their Transport across Monolayers of Human Intestinal Epithelial Caco-2 Cells. Food Funct. 2014, 5, 1278–1285. [Google Scholar] [CrossRef] [PubMed]
  346. Li, Z.; Gu, L. Fabrication of Self-Assembled (−)-Epigallocatechin Gallate (EGCG) Ovalbumin-Dextran Conjugate Nanoparticles and Their Transport across Monolayers of Human Intestinal Epithelial Caco-2 Cells. J. Agric. Food Chem. 2014, 62, 1301–1309. [Google Scholar] [CrossRef]
  347. Sahadevan, R.; Singh, S.; Binoy, A.; Sadhukhan, S. Chemico-Biological Aspects of (−)-Epigallocatechin-3-Gallate (EGCG) to Improve Its Stability, Bioavailability and Membrane Permeability: Current Status and Future Prospects. Crit. Rev. Food Sci. Nutr. 2022, 1–30. [Google Scholar] [CrossRef] [PubMed]
  348. Cai, Z.-Y.; Li, X.-M.; Liang, J.-P.; Xiang, L.-P.; Wang, K.-R.; Shi, Y.-L.; Yang, R.; Shi, M.; Ye, J.-H.; Lu, J.-L.; et al. Bioavailability of Tea Catechins and Its Improvement. Molecules 2018, 23, 2346. [Google Scholar] [CrossRef] [Green Version]
  349. Janle, E.M.; Morré, D.M.; Morré, D.J.; Zhou, Q.; Zhu, Y. Pharmacokinetics of Green Tea Catechins in Extract and Sustained-Release Preparations. J. Diet. Suppl. 2008, 5, 248–263. [Google Scholar] [CrossRef] [Green Version]
  350. Andreu Fernández, V.; Almeida Toledano, L.; Pizarro Lozano, N.; Navarro Tapia, E.; Gómez Roig, M.D.; De la Torre Fornell, R.; García Algar, Ó. Bioavailability of Epigallocatechin Gallate Administered with Different Nutritional Strategies in Healthy Volunteers. Antioxidants 2020, 9, 440. [Google Scholar] [CrossRef]
  351. Mereles, D.; Hunstein, W. Epigallocatechin-3-Gallate (EGCG) for Clinical Trials: More Pitfalls than Promises? Int. J. Mol. Sci. 2011, 12, 5592–5603. [Google Scholar] [CrossRef] [Green Version]
  352. Chen, L.; Lee, M.J.; Li, H.; Yang, C.S. Absorption, Distribution, Elimination of Tea Polyphenols in Rats. Drug Metab. Dispos. 1997, 25, 1045–1050. [Google Scholar]
  353. Oketch-Rabah, H.A.; Roe, A.L.; Rider, C.V.; Bonkovsky, H.L.; Giancaspro, G.I.; Navarro, V.; Paine, M.F.; Betz, J.M.; Marles, R.J.; Casper, S.; et al. United States Pharmacopeia (USP) Comprehensive Review of the Hepatotoxicity of Green Tea Extracts. Toxicol. Rep. 2020, 7, 386–402. [Google Scholar] [CrossRef]
  354. Siblini, H.; Al-Hendy, A.; Segars, J.; González, F.; Taylor, H.S.; Singh, B.; Flaminia, A.; Flores, V.A.; Christman, G.M.; Huang, H.; et al. Assessing the Hepatic Safety of Epigallocatechin Gallate (EGCG) in Reproductive-Aged Women. Nutrients 2023, 15, 320. [Google Scholar] [CrossRef] [PubMed]
  355. 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] [PubMed] [Green Version]
  356. Isbrucker, R.A.; Edwards, J.A.; Wolz, E.; Davidovich, A.; Bausch, J. Safety Studies on Epigallocatechin Gallate (EGCG) Preparations. Part 2: Dermal, Acute and Short-Term Toxicity Studies. Food Chem. Toxicol. 2006, 44, 636–650. [Google Scholar] [CrossRef] [PubMed]
  357. Kapetanovic, I.M.; Crowell, J.A.; Krishnaraj, R.; Zakharov, A.; Lindeblad, M.; Lyubimov, A. Exposure and Toxicity of Green Tea Polyphenols in Fasted and Non-Fasted Dogs. Toxicology 2009, 260, 28–36. [Google Scholar] [CrossRef] [Green Version]
  358. Isbrucker, R.A.; Bausch, J.; Edwards, J.A.; Wolz, E. Safety Studies on Epigallocatechin Gallate (EGCG) Preparations. Part 1: Genotoxicity. Food Chem. Toxicol. 2006, 44, 626–635. [Google Scholar] [CrossRef]
  359. Isbrucker, R.A.; Edwards, J.A.; Wolz, E.; Davidovich, A.; Bausch, J. Safety Studies on Epigallocatechin Gallate (EGCG) Preparations. Part 3: Teratogenicity and Reproductive Toxicity Studies in Rats. Food Chem. Toxicol. 2006, 44, 651–661. [Google Scholar] [CrossRef] [PubMed]
  360. Hu, J.; Webster, D.; Cao, J.; Shao, A. The Safety of Green Tea and Green Tea Extract Consumption in Adults—Results of a Systematic Review. Regul. Toxicol. Pharmacol. 2018, 95, 412–433. [Google Scholar] [CrossRef] [PubMed]
  361. Hong, J.; Lambert, J.D.; Lee, S.H.; Sinko, P.J.; Yang, C.S. Involvement of Multidrug Resistance-Associated Proteins in Regulating Cellular Levels of (−)-Epigallocatechin-3-Gallate and Its Methyl Metabolites. Biochem. Biophys. Res. Commun. 2003, 310, 222–227. [Google Scholar] [CrossRef] [PubMed]
  362. Jodoin, J.; Demeule, M.; Beliveau, R. Inhibition of the Multidrug Resistance P-Glycoprotein Activity by Green Tea Polyphenols. Biochim. Biophys. Acta 2002, 1542, 149–159. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  363. Rein, M.J.; Renouf, M.; Cruz-Hernandez, C.; Actis-Goretta, L.; Thakkar, S.K.; da Silva Pinto, M. Bioavailability of Bioactive Food Compounds: A Challenging Journey to Bioefficacy. Br. J. Clin. Pharmacol. 2013, 75, 588–602. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  364. Scholl, C.; Lepper, A.; Lehr, T.; Hanke, N.; Schneider, K.L.; Brockmöller, J.; Seufferlein, T.; Stingl, J.C. Population Nutrikinetics of Green Tea Extract. PLoS ONE 2018, 13, e0193074. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  365. Misaka, S.; Abe, O.; Sato, H.; Ono, T.; Shikama, Y.; Onoue, S.; Yabe, H.; Kimura, J. Lack of Pharmacokinetic Interaction between Fluvastatin and Green Tea in Healthy Volunteers. Eur. J. Clin. Pharmacol. 2018, 74, 601–609. [Google Scholar] [CrossRef]
  366. Abe, O.; Ono, T.; Sato, H.; Müller, F.; Ogata, H.; Miura, I.; Shikama, Y.; Yabe, H.; Onoue, S.; Fromm, M.F.; et al. Role of (−)-Epigallocatechin Gallate in the Pharmacokinetic Interaction between Nadolol and Green Tea in Healthy Volunteers. Eur. J. Clin. Pharmacol. 2018, 74, 775–783. [Google Scholar] [CrossRef]
  367. Veerman, G.D.M.; van der Werff, S.C.; Koolen, S.L.W.; Miedema, J.R.; Oomen-de Hoop, E.; van der Mark, S.C.; Chandoesing, P.P.; de Bruijn, P.; Wijsenbeek, M.S.; Mathijssen, R.H.J. The Influence of Green Tea Extract on Nintedanib’s Bioavailability in Patients with Pulmonary Fibrosis. Biomed. Pharmacother. 2022, 151, 113101. [Google Scholar] [CrossRef]
  368. Misaka, S.; Ono, Y.; Taudte, R.V.; Hoier, E.; Ogata, H.; Ono, T.; König, J.; Watanabe, H.; Fromm, M.F.; Shimomura, K. Exposure of Fexofenadine, but Not Pseudoephedrine, Is Markedly Decreased by Green Tea Extract in Healthy Volunteers. Clin. Pharmacol. Ther. 2022, 112, 627–634. [Google Scholar] [CrossRef]
  369. Lambert, J.D.; Sang, S.; Yang, C.S. Biotransformation of Green Tea Polyphenols and the Biological Activities of Those Metabolites. Mol. Pharm. 2007, 4, 819–825. [Google Scholar] [CrossRef]
  370. Lee, M.-J.; Maliakal, P.; Chen, L.; Meng, X.; Bondoc, F.Y.; Prabhu, S.; Lambert, G.; Mohr, S.; Yang, C.S. Pharmacokinetics of Tea Catechins after Ingestion of Green Tea and (−)-Epigallocatechin-3-Gallate by Humans: Formation of Different Metabolites and Individual Variability. Cancer Epidemiol. Biomark. Prev. 2002, 11, 1025–1032. [Google Scholar]
  371. Nagle, D.G.; Ferreira, D.; Zhou, Y.-D. Epigallocatechin-3-Gallate (EGCG): Chemical and Biomedical Perspectives. Phytochemistry 2006, 67, 1849–1855. [Google Scholar] [CrossRef] [Green Version]
  372. Dai, W.; Ruan, C.; Zhang, Y.; Wang, J.; Han, J.; Shao, Z.; Sun, Y.; Liang, J. Bioavailability Enhancement of EGCG by Structural Modification and Nano-Delivery: A Review. J. Funct. Foods 2020, 65, 103732. [Google Scholar] [CrossRef]
  373. Hayashi, A.; Terasaka, S.; Nukada, Y.; Kameyama, A.; Yamane, M.; Shioi, R.; Iwashita, M.; Hashizume, K.; Morita, O. 4″-Sulfation Is the Major Metabolic Pathway of Epigallocatechin-3-Gallate in Humans: Characterization of Metabolites, Enzymatic Analysis, and Pharmacokinetic Profiling. J. Agric. Food Chem. 2022, 70, 8264–8273. [Google Scholar] [CrossRef] [PubMed]
  374. Li, Y.; Gao, X.; Lou, Y. Interactions of Tea Polyphenols with Intestinal Microbiota and Their Implication for Cellular Signal Conditioning Mechanism. J. Food Biochem. 2019, 43, e12953. [Google Scholar] [CrossRef] [PubMed]
  375. Naumovski, N.; Blades, B.L.; Roach, P.D. Food Inhibits the Oral Bioavailability of the Major Green Tea Antioxidant Epigallocatechin Gallate in Humans. Antioxidants 2015, 4, 373–393. [Google Scholar] [CrossRef] [Green Version]
  376. Wang, P.; Heber, D.; Henning, S.M. Quercetin Increased Bioavailability and Decreased Methylation of Green Tea Polyphenols in Vitro and in Vivo. Food Funct. 2012, 3, 635–642. [Google Scholar] [CrossRef] [Green Version]
  377. Peters, C.M.; Green, R.J.; Janle, E.M.; Ferruzzi, M.G. Formulation with Ascorbic Acid and Sucrose Modulates Catechin Bioavailability from Green Tea. Food Res. Int. 2010, 43, 95–102. [Google Scholar] [CrossRef] [Green Version]
  378. Lazzeroni, M.; Guerrieri-Gonzaga, A.; Gandini, S.; Johansson, H.; Serrano, D.; Cazzaniga, M.; Aristarco, V.; Macis, D.; Mora, S.; Caldarella, P.; et al. A Presurgical Study of Lecithin Formulation of Green Tea Extract in Women with Early Breast Cancer. Cancer Prev. Res. 2017, 10, 363–370. [Google Scholar] [CrossRef] [Green Version]
  379. Hayasaka, S.; Goto, Y.; Maeda-Yamamoto, M. The Effects of Bathing in Hot Springs on the Absorption of Green Tea Catechin: A Pilot Study. Complement. Ther. Clin. Pract. 2013, 19, 243–245. [Google Scholar] [CrossRef]
Figure 1. Chemical structure of epicatechin (EC), epicatechin-3-gallate (ECG), epigallocatechin (EGC), and epigallocatechin-3-gallate (EGCG). Created with BioRender.com, accessed on 29 March 2023.
Figure 1. Chemical structure of epicatechin (EC), epicatechin-3-gallate (ECG), epigallocatechin (EGC), and epigallocatechin-3-gallate (EGCG). Created with BioRender.com, accessed on 29 March 2023.
Molecules 28 05246 g001
Figure 2. The molecular basis of the anticancer mechanism of action of epigallocatechin-3-gallate (EGCG) related to the suppression of inflammation. Cyclooxygenase 2 enzyme (COX2) catalyzes the conversion of arachidonic acid to prostaglandins (PGEs), including PGE2. PGE2 stimulates the expression of anti-apoptotic BCL-2 protein and the nuclear factor NF-kappa-B (NF-κB)-mediated gene expression of vascular endothelial growth factor, interleukins (IL-6 and -8), matrix metalloproteinases (MMPs), and COX2. EGCG was shown to act as a COX2, PGE2, NFκB, and MMPs inhibitor, suppressing the angiogenesis and metastasis of cancer cells and enhancing apoptosis induction. Additionally, EGCG may work as phosphoinositide-3-kinase/RAC-alpha serine/threonine-protein kinase (PI3K/AKT), epithelial growth factor (EGFR), and vascular endothelia growth factor, which control NF-κB and MMPs activity and expression and confer enhanced proliferation, angiogenesis, and metastasis. Furthermore, EGCG treatment may down-regulate the expression of tumor necrosis factor α (TNF-α) and suppress the activation of tumor necrosis factor receptor (TNFR) signaling. For full protein names, see the abbreviations section. Red arrows indicate down-regulation of the protein expression or its inhibition. Created with BioRender.com, accessed on 29 March 2023.
Figure 2. The molecular basis of the anticancer mechanism of action of epigallocatechin-3-gallate (EGCG) related to the suppression of inflammation. Cyclooxygenase 2 enzyme (COX2) catalyzes the conversion of arachidonic acid to prostaglandins (PGEs), including PGE2. PGE2 stimulates the expression of anti-apoptotic BCL-2 protein and the nuclear factor NF-kappa-B (NF-κB)-mediated gene expression of vascular endothelial growth factor, interleukins (IL-6 and -8), matrix metalloproteinases (MMPs), and COX2. EGCG was shown to act as a COX2, PGE2, NFκB, and MMPs inhibitor, suppressing the angiogenesis and metastasis of cancer cells and enhancing apoptosis induction. Additionally, EGCG may work as phosphoinositide-3-kinase/RAC-alpha serine/threonine-protein kinase (PI3K/AKT), epithelial growth factor (EGFR), and vascular endothelia growth factor, which control NF-κB and MMPs activity and expression and confer enhanced proliferation, angiogenesis, and metastasis. Furthermore, EGCG treatment may down-regulate the expression of tumor necrosis factor α (TNF-α) and suppress the activation of tumor necrosis factor receptor (TNFR) signaling. For full protein names, see the abbreviations section. Red arrows indicate down-regulation of the protein expression or its inhibition. Created with BioRender.com, accessed on 29 March 2023.
Molecules 28 05246 g002
Figure 3. The modulation of epigenetic targets by epigallocatechin-3-gallate (EGCG). Methylation of CpG islands and histone proteins works as a mark that confers the closed chromatin formation and transcriptional suppression of genes encoding proteins involved in tumor suppression, invasion and metastasis inhibition, and apoptosis induction. EGCG down-regulates and inhibits the activity of DNA methyltransferases (DNMT1, DNMT3A, and 3B) that introduce methyl groups to DNA. Moreover, EGCG was shown to inhibit the activity of histone deacetylases (HDACs) that counteract the introduction of acetyl groups to histone proteins via histone acetyltransferases (HATs). The introduction of acetyl groups to histone proteins (H3K9, H3K14, H4K5, H4K12, and H4K16) contributes to the open chromatin state and elevated expression of tumor suppressor proteins (CDH1, DAPK1, P12, P16INK4a, P27, and RARβ), proteins involved in apoptosis induction and cell death control (DR5, GAD153, and TP53), and TIMP-3, which inhibits the proteolytic activity of MMP-2 and MMP-9. For full protein names, see the abbreviations section. Created with BioRender.com, accessed on 29 March 2023.
Figure 3. The modulation of epigenetic targets by epigallocatechin-3-gallate (EGCG). Methylation of CpG islands and histone proteins works as a mark that confers the closed chromatin formation and transcriptional suppression of genes encoding proteins involved in tumor suppression, invasion and metastasis inhibition, and apoptosis induction. EGCG down-regulates and inhibits the activity of DNA methyltransferases (DNMT1, DNMT3A, and 3B) that introduce methyl groups to DNA. Moreover, EGCG was shown to inhibit the activity of histone deacetylases (HDACs) that counteract the introduction of acetyl groups to histone proteins via histone acetyltransferases (HATs). The introduction of acetyl groups to histone proteins (H3K9, H3K14, H4K5, H4K12, and H4K16) contributes to the open chromatin state and elevated expression of tumor suppressor proteins (CDH1, DAPK1, P12, P16INK4a, P27, and RARβ), proteins involved in apoptosis induction and cell death control (DR5, GAD153, and TP53), and TIMP-3, which inhibits the proteolytic activity of MMP-2 and MMP-9. For full protein names, see the abbreviations section. Created with BioRender.com, accessed on 29 March 2023.
Molecules 28 05246 g003
Figure 4. Overview of inhibitory effects of EGCG on key signaling pathways associated with the development of cancer. The details can be found in the main text of the manuscript. For the full names of the proteins, see the abbreviations section. Created with BioRender.com, accessed on 29 March 2023.
Figure 4. Overview of inhibitory effects of EGCG on key signaling pathways associated with the development of cancer. The details can be found in the main text of the manuscript. For the full names of the proteins, see the abbreviations section. Created with BioRender.com, accessed on 29 March 2023.
Molecules 28 05246 g004
Figure 5. Overview of EGCG’s metabolism. Based on [347,348,372,373,374]. A detailed description can be found in the main text of the article. For full names, see the abbreviations section. Created with BioRender.com, accessed on 29 March 2023.
Figure 5. Overview of EGCG’s metabolism. Based on [347,348,372,373,374]. A detailed description can be found in the main text of the article. For full names, see the abbreviations section. Created with BioRender.com, accessed on 29 March 2023.
Molecules 28 05246 g005
Figure 6. Anticancer properties of EGCG. EGCG exhibits antioxidant, anti-inflammatory, anti-proliferative, and anti-metastatic properties through the inhibition or modulation of key signaling pathways in cells and targets of epigenetic machinery. Details can be found in the main text of this article. Based on [94]. Created with BioRender.com, accessed on 29 March 2023.
Figure 6. Anticancer properties of EGCG. EGCG exhibits antioxidant, anti-inflammatory, anti-proliferative, and anti-metastatic properties through the inhibition or modulation of key signaling pathways in cells and targets of epigenetic machinery. Details can be found in the main text of this article. Based on [94]. Created with BioRender.com, accessed on 29 March 2023.
Molecules 28 05246 g006
Table 1. Ongoing and future clinical studies on EGCG in cancer treatment.
Table 1. Ongoing and future clinical studies on EGCG in cancer treatment.
TitleNumberStatusEstimated EnrollmentObjective
A Pilot Study to Evaluate the Chemopreventive Effects of Epigallocatechin Gallate (EGCG) in Colorectal Cancer (CRC) Patients with Curative Resections.NCT02891538Recruiting50Evaluating cthe hemopreventive properties of EGCG in CRC patients with curative resections by comparing DNA methylation change during 1 year of EGCG treatment.
The Study of Quadruple Therapy Quercetin, Zinc, Metformin, and EGCG as Adjuvant Therapy for Early, Metastatic Breast Cancer and Triple-negative Breast Cancer, a Novel MechanismNCT05680662Not yet recruiting200Achieving the best efficacy for different stages breast cancer and triple-negative breast cancer treatment in female patients by applying various combinations of adjuvants.
Study of Epigallocatechin-3-gallate (EGCG) for Supportive and Symptomatic Management in Patients with Esophageal CancerNCT05039983Recruiting15Assessing the application of EGCG for alleviating esophageal stenosis, choking and pain while swallowing, weight change, and blood parameter changes in pathological esophageal squamous cell carcinoma patients.
Reducing Frailty for Older Cancer Survivors Using Supplements (ReFOCUS): A Phase 2 Randomized Controlled Trial of Epigallocatechin-3-Gallate (EGCG) on Frailty and Inflammation in Older Survivors of CancerNCT04553666Recruiting40Investigating the beneficial properties and safety of EGCG treatment in older cancer survivors with frailty and inflammation.
A Phase II Randomized Double Blinded Study of Green Tea Catechins (GTC) vs. Placebo in Men on Active Surveillance for Prostate Cancer: Modulation of Biological and Clinical Intermediate BiomarkersNCT04597359Recruiting360Assessing the abilities of tea catechins in preventing the progression of prostate cancer using various genetic, biochemical, and clinical markers.
Phase II Clinical Trial of Green Tea Catechins in Men on Active Surveillance (AS)NCT04300855Recruiting135An evaluation of the bioavailability, safety, effectiveness, and
validation of the acting mechanism of a drug containing EGCG applied for 24 months to men with adenocarcinoma of the prostate with cancer present.
Pioneering Pre- and Post-Operative Integrative Care to Improve Thoracic Cancer Quality of Care—The Thoracic Peri-Operative Integrative Surgical Care Evaluation (POISE) Trial—Stage IINCT04871412Recruiting20Improving health-related quality of life, decreasing surgical adverse events, prolonging overall survival, and pioneering integrative care delivery for thoracic cancer patients.
Fibroids and Unexplained Infertility Treatment With Epigallocatechin Gallate; A Natural CompounD in Green Tea (FRIEND)NCT05364008Recruiting200Determining the effect of low-caffeine green tea extract containing 45% EGCG on fibroids and subsequent pregnancy and live births in women seeking fertility treatment.
The Use of Vitamin D in Combination With Epigallocatechin Gallate, D-chiro-inositol and Vitamin B6 in the Treatment of Women With Uterine FibroidNCT05448365Recruiting60Evaluating the impact on uterine fibroid volume of a combination of natural molecules including epigallocatechin gallate, vitamin D, D-chiro-inositol, and vitamin B6.
Effects of Vitamin D, Epigallocatechin Gallate, Vitamin B6, and D-Chiro-inositol Combination on Uterine Fibroids: a Randomized Controlled TrialNCT05409872Recruiting108Evaluating the efficacy of a combination of epigallocatechin gallate, vitamin D3, D-chiro-inositol, and vitamin B6 as a treatment for uterine fibroids.
A Phase I Single-arm, Multicenter Pilot Study Aimed at Validating γ-OHPdG as a Biomarker and Testing the Effects of Polyphenon E on Its Levels in Patients With CirrhosisNCT03278925Active, not recruiting48Investigating the side effects and best dose of defined green tea catechin extract and verifying its ability to prevent liver cancer in participants with cirrhosis.
A Pilot Study of Gemcitabine, Abraxane, Metformin and a Standardized Dietary Supplement (DS) in Patients With Unresectable Pancreatic CancerNCT02336087Active, not recruiting21Investigating the side effects of gemcitabine hydrochloride, nab-paclitaxel, metformin hydrochloride, and a standardized dietary supplement including EGCG in treating patients with pancreatic cancer that cannot be removed by surgery.
Phase ⅠStudy of Oral Green Tea Extract as Maintenance Therapy for Extensive-stage Small Cell Lung CancerNCT01317953AvailableNo informationAssessing the safety of the application of EGCG for extensive-stage small lung cancer patients who have achieved an objective tumor response after first-line therapy.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kciuk, M.; Alam, M.; Ali, N.; Rashid, S.; Głowacka, P.; Sundaraj, R.; Celik, I.; Yahya, E.B.; Dubey, A.; Zerroug, E.; et al. Epigallocatechin-3-Gallate Therapeutic Potential in Cancer: Mechanism of Action and Clinical Implications. Molecules 2023, 28, 5246. https://doi.org/10.3390/molecules28135246

AMA Style

Kciuk M, Alam M, Ali N, Rashid S, Głowacka P, Sundaraj R, Celik I, Yahya EB, Dubey A, Zerroug E, et al. Epigallocatechin-3-Gallate Therapeutic Potential in Cancer: Mechanism of Action and Clinical Implications. Molecules. 2023; 28(13):5246. https://doi.org/10.3390/molecules28135246

Chicago/Turabian Style

Kciuk, Mateusz, Manzar Alam, Nemat Ali, Summya Rashid, Pola Głowacka, Rajamanikandan Sundaraj, Ismail Celik, Esam Bashir Yahya, Amit Dubey, Enfale Zerroug, and et al. 2023. "Epigallocatechin-3-Gallate Therapeutic Potential in Cancer: Mechanism of Action and Clinical Implications" Molecules 28, no. 13: 5246. https://doi.org/10.3390/molecules28135246

Article Metrics

Back to TopTop