Next Article in Journal
ACTN4 Mediates SEPT14 Mutation-Induced Sperm Head Defects
Next Article in Special Issue
Chronic Treatment with a Phytosomal Preparation Containing Centella asiatica L. and Curcuma longa L. Affects Local Protein Synthesis by Modulating the BDNF-mTOR-S6 Pathway
Previous Article in Journal
Macrophage Polarization in Chronic Lymphocytic Leukemia: Nurse-Like Cells Are the Caretakers of Leukemic Cells
Previous Article in Special Issue
Treatment with Luteolin Improves Lipopolysaccharide-Induced Periodontal Diseases in Rats
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Molecular Insights into the Multifunctional Role of Natural Compounds: Autophagy Modulation and Cancer Prevention

1
Center for Neuroscience, Korea Institute of Science and Technology (KIST), 5 Hwarang-ro 14-gil, Seoul 02792, Korea
2
Global Biotechnology & Biomedical Research Network (GBBRN), Department of Biotechnology and Genetic Engineering, Faculty of Biological Sciences, Islamic University, Kushtia 7003, Bangladesh
3
ABEx Bio-Research Center, East Azampur, Dhaka 1230, Bangladesh
4
Department of Biotechnology and Genetic Engineering, Bangabandhu Sheikh Mujibur Rahman Science and Technology University, Gopalganj 8100, Bangladesh
5
Department of Biotechnology and Genetic Engineering, Noakhali Science and Technology University, Noakhali 3814, Bangladesh
6
Department of Genetic Engineering and Biotechnology, Jashore University of Science and Technology, Jashore 7408, Bangladesh
7
Department of Biotechnology and Genetic Engineering, Faculty of Biological Sciences, Islamic University, Kushtia 7003, Bangladesh
8
Graduate School of Pharmaceutical Sciences, College of Pharmacy, Ewha Womans University, Seoul 03760, Korea
9
Department of Global Medical Science, Wonju College of Medicine, Yonsei University, Seoul 03722, Korea
10
Division of Bio-Medical Science and Technology, KIST School, Korea University of Science and Technology (UST), Seoul 02792, Korea
*
Authors to whom correspondence should be addressed.
Submission received: 15 October 2020 / Revised: 12 November 2020 / Accepted: 12 November 2020 / Published: 19 November 2020
(This article belongs to the Special Issue Natural Medicine in Therapy)

Abstract

:
Autophagy is a vacuolar, lysosomal degradation pathway for injured and damaged protein molecules and organelles in eukaryotic cells, which is controlled by nutrients and stress responses. Dysregulation of cellular autophagy may lead to various diseases such as neurodegenerative disease, obesity, cardiovascular disease, diabetes, and malignancies. Recently, natural compounds have come to attention for being able to modulate the autophagy pathway in cancer prevention, although the prospective role of autophagy in cancer treatment is very complex and not yet clearly elucidated. Numerous synthetic chemicals have been identified that modulate autophagy and are favorable candidates for cancer treatment, but they have adverse side effects. Therefore, different phytochemicals, which include natural compounds and their derivatives, have attracted significant attention for use as autophagy modulators in cancer treatment with minimal side effects. In the current review, we discuss the promising role of natural compounds in modulating the autophagy pathway to control and prevent cancer, and provide possible therapeutic options.

Graphical Abstract

1. Introduction

Autophagy is triggered by the entrapment of abnormal intracellular proteins, invading microorganisms, and damaged organelles with the formation of double-layer autophagosomes, in the presence of nutrient stress, injury, and fasting [1,2]. Along with cellular energy stability, autophagy contributes to the control of cellular quality, and destruction of abnormal proteins and damaged organelles [3]. As a result, the origin and development of many diseases, such as neurodegenerative disorders, cancer, and autoimmune diseases, can be explained by defective autophagy [4]. Multiple data states a connection between the variety of diet and feedback to cancer medication [5]. It has been established that sufficient consumption of dietary and medicinal plant-derived biochemical compounds lowers the frequency of cancer mortality. This occurs due to the activation or balancing of various cellular oncogenes. In different human disorders, including cancer, there is documented evidence of reduction of autophagic regulation. Several lines of evidence have shown that cancer plays a critical role in inhibiting or augmenting autophagic pathways [6]. In addition, this dual role of autophagy has a significant therapeutic benefit against cancer [7]. The present report has established that natural product-derived bioactive molecules, along with lysosomal inhibitors, including chloroquine (CQ) and hydroxychloroquine (HCQ) are important regulators of autophagy signaling [8]. In addition, these proactive molecules can regulate the process of autophagy both in vitro and in vivo by involving the enzymes, transcription factors, and various intracellular communication pathways [9]. The autophagy stimulation or prohibition process is vastly complicated and regulated, so it needs to be extensively explored. However, emerging evidence has implicated that extreme or damaged autophagy might lead to a unique type of cell death known as autophagic cell death [10]. Conversely, the proposal of using natural compounds as anti-cancer and autophagy-modulating agents requires a better understanding of their cellular and molecular mechanisms. The mechanism of action needs to be further addressed.
In this review, we compiled the cellular and molecular mechanisms involved in the dual role of autophagy, as a tumor-suppressing as well as a tumor-augmenting phenomenon in cancer. Current improvements in the use of natural bioactive compounds for cancer treatment by targeting autophagic action have been consistently analyzed [11]. Given the vital role of autophagy in cancer prevention and treatment, we closely reviewed different plant-derived biomolecules that may play a role in modulating the autophagy-linked signaling pathways as well as contribute towards competent treatment strategies against specific molecular aspects in cancer. In particular, administration of naturally derived compounds have been considered to regulate autophagy modulation, which might help improve our understanding of the mechanisms of natural compounds in the management and treatment of autophagy-linked diseases and cancers.

2. Mechanisms of Autophagy Signaling Pathway

Autophagy is a sensitive and tightly mannered process that has multiple steps [12]. Proteins crucial for autophagy were first reported in yeast and are termed as autophagy-related (Atg) proteins [1,13]. Multiple sequential molecular events take place to initiate the autophagy signaling cascade. First, an Atg1/unc51-like kinase (ULK) complex kinase regulates the induction, and the BECN1/class III PI3K complex initiates the nucleation of vesicles [14]. In addition, ubiquitin-like conjugation systems (Atg12 and Atg8/LC3B) control vesicle elongation as well as retrieval of lipids mediated by transmembrane Atg9 and associated proteins, which involve other Atg proteins [15]. Lysosome-associated membrane protein 2 (LAMP2) and RAS-related protein-7 (RAB7A) participate in the fusion between lysosomes and autophagosome formation [16]. Finally, vesicle breakdown and degradation by lysosomal hydrolases produce recyclable products [17] (Figure 1). Microtubule-associated protein 1 light chain 3-beta (MAP1LC3B) and Atg5-Atg12 complex pathways, which are ubiquitin-like pathways, are regulatory for vesicle elongation and are encoded by the yeast Atg8 pathway, which has a human homologue [18]. Following the autophagy response, human Atg4 cysteine protease cleaves LC3B protein and thereby produces the LC3B-I isoform [19]. Activated LC3B-I isoform causes the activation of Atg7 protein through LC3B-I and is transported to the Atg3 before formation of the LC3B-II isoform by the phosphatidylethanolamine (PE) conjugation with the carboxyl glycine of LC3B-I protein [20]. As a consequence, elevated synthesis and processing of LC3B shows promise and has been mentioned as a key marker of autophagy [20]. In the degradation of matured autophagosomes through the enzymatic action of lysosomal enzymes, small molecules are produced that are used for de novo protein synthesis, helping in survival and maintenance of homeostasis [21].

3. Molecular Mechanism of Autophagy Signaling in Cancer Pathogenesis

Under healthy conditions, cells have an innate ability for autophagy mechanisms to defend against malignant transformation [22]. Generally, autophagy is initiated by the induction of tumor-suppressing proteins, and oncoproteins are reported to inhibit this action [23]. Therefore, autophagy can act in both pro-survival and pro-death stages of tumor initiation and development [23]. The disturbed autophagy pathway can contribute to tumor development, as it leads to accumulation of damaged organelles and protein aggregates, which ultimately produce reactive oxygen species (ROS) and lead to genome instability [24]. Several researchers have reported that chemotherapeutics and avenues for cancer treatment can be accelerated by altering autophagic signaling that result in tumor cell death by inhibition of pro-survival and tissue specific apoptosis inducing factors [25]. Additionally, autophagy has been described as having a dual function in the promotion as well as inhibition of metastasis [26]. In the early stages, autophagy prevents metastasis, whereas in later stages it favors metastatic activity [26]. However, poorly vascularized tumor cells may survive at low nutrient and low oxygen levels through TGF-β, and other signals that trigger autophagy [27]. Furthermore, autophagy has directly regulated the metastatic cascade mainly through migration, invasion, and epithelial to mesenchymal transition (EMT) of cancer cells [26,28]. The importance of autophagy at diverse stages of tumorigenesis and metastasis is illustrated in Figure 2. It has been found that many cells within the tumors such as immune cells, fibroblasts, and endothelial cells are surrounded by the tumor microenvironment and have been acknowledged as an attractive target to reduce resistance to anticancer treatment. Natural compounds from vegetables, spices, marine organisms, herbs, and fruits have been reported to prevent or reverse multistage carcinogenesis as well as to prevent cancerous cell proliferation [29]. Additionally, microRNAs (miRNAs), which control the expression of genes, play an important role in diverse biological processes. It has been found that miRNA expression dysregulation is highly related to cancer development [30]. Recently, several studies have shown that natural compounds, such as paclitaxel (PTX), genistein, curcumin, epigallocatechin-3-gallate (EGCG), and resveratrol exhibit pro-apoptotic or anti-proliferative properties that are controlled by miRNAs, and lead to inhibition of cancer cell growth and proliferation, apoptosis induction, and improvement of conventional cancer treatment and therapeutic efficiency [30]. However, these naturally occurring compounds could be modulated by different signal transduction pathways through their connection via cancer cells, tumor microenvironment, and miRNA. These are described later in this study.

3.1. Autophagy Signaling Roles in Cancer Inhibition

Several studies have reported that autophagy inhibition via manipulated pharmacological or genetic tools can be a possible chemotherapeutic option. Recently, scientists have focused on the combined treatment of chemically synthesized autophagy inhibitors and traditional chemotherapeutics [31]. Cancer cell sensitization with 3-methyladenine (3-MA), wortmannin, CQ, LY294002, bafilomycin A1 (Baf A1), and HCQ-like inhibitors of autophagy to chemotherapeutic drugs have been reported to accelerate cell death pathway action [32,33]. In the abovementioned inhibitors, some of them are preliminary stage inhibitors of autophagy, like wortmannin, LY294002, and 3-MA, which block the autophagosomes formation, while others act on the basis of lysosomal action at late stages [34]. Table 1 shows several valuable effects of natural compounds combined with some autophagy inhibitors in different cellular cancer models. In breast cancer cells, the pro-survival nature of cancer cells can be downregulated by LC3 via shRNA blocking and sensitization of the carcinoma cells to apoptotic factors such as trastuzumab [35]. In leukemia, pro-survival function with pharmacological inhibitors or RNAi can also block accelerated induced cell death by imatinib mesylate [36]. Autophagy can sensitize and make cells susceptible to radiotherapy, such as HBL-100 cells, for radiotherapy treatment, and 3-MA leads to the inhibition of autophagy by imposing the pro-death effects of therapy [37]. In BIF-1 knockout mice, there is impairment of autophagy in addition to an increase in the rate of tumor formation, while mutations in UVRAG decrease autophagy along with increased proliferation of colorectal cancer cells [38]. These reports propose that Beclin-1 acts as an insufficient tumor-suppressor gene and fully incorporates the theory of tumor-suppressive autophagy induction at early stages [39]. In the livers of mice, Atg gene deletion can lead to the formation of multiple benign tumors. In systematic mosaic Atg5 and Atg7 deletion mice, these specific deletions magnify the chance of tumor induction and lead to tumor development [40]. At early stages of cancer formation, Atg4 suppresses tumor developing activity and Atg4c deficiency is proven to be more easily identifiable in chemical carcinogen-induced fibrosarcoma. These reports help to understand the suppressive roles of autophagy in cancer with genetic deletion of specific regions [41].

3.2. Autophagy Signaling Roles in Cancer Promotion

Autophagy induction is a possible anti-cancer mechanism in which cancer cells have a faulty or altered apoptotic cell death pathway [52]. Consequently, autophagy induction along with autophagy inhibition can induce apoptosis and exert anti-cancer effects [53]. Several inhibitors, such as Bcl-2, EGFR, and mTOR, are currently used in cancer treatment [54]. Inhibitory molecules are known to be effective against a variety of cancers, and some induce autophagy in breast cancers, gliomas, and lung cancers through rapamycin, an inhibitor of mTORC2 [54]. Another known inhibitor of the mTOR, deforolimus (AP23573, MK-8669), is proposed to treat patients with relapsed or refractory hematologic malignancies [55]. Another antibody against epidermal growth factor, cetuximab, can regulate autophagic cell death in cancer cells by disrupting the interaction between Bcl-2 and Beclin-1 [56]. Several reports explain that the expression of the viral oncogenes Kirsten and Harvey RAS virus is capable of autophagy induction in certain cells [22]. In pancreatic and colorectal cancer cells, elevated levels of RAS mutations are reported to be correlated with higher autophagy rates, where increased levels maintain cell proliferation in RAS-activated tumors and inhibition of autophagy results in reduced cell proliferation, ultimately leading to tumor regression [57]. Lung cancers are also dependent on autophagy, which is similar to the mechanism of RAS-driven cancers regulated by valine-to-glutamic acid substitution at BRAF position 600 (BRAFV600E) [58]. In vivo studies have reported that deletion of subunit FIP200 related to the ULK1 complex in autophagy initiation blocks the growth of breast cancer and extends the life cycle of mice [59]. In addition, activated AMPK generated by ATP can also trigger an autophagic event [60].

4. Natural Compound Triggers in Autophagy Modulation

A single therapeutic strategy and a combination modality can both be favorable for cancer treatment. The various strategies include radiation therapy, immune or genetic therapy, and chemotherapy, all of which create new diversions and avenues in the field of cancer treatment with adequate levels of safety and accuracy [15,61,62]. Natural compounds are effective in the positive regulation of anticancer activity, as they inhibit cellular proliferation, adjust the oxidative stress response, and manipulate autophagic signaling and response [15,63]. This entire response of natural compounds is dependent on the cell type. The role of several natural compounds that modulate different pathways of autophagy signaling is illustrated in Figure 3.

4.1. Inhibition of Autophagy by Natural Compounds for Cancer Therapy

Apigenin: Apigenin, a flavonoid found in vegetables and fruits, can initiate tumor cell death through autophagy and apoptosis [65]. In combination with retinamide, N-(4-hydroxyphenyl), apigenin suppresses autophagy and increases apoptosis in human neuroblastoma (NB) cells [66].
Genistein: Genistein, a nutritive natural compound, has been documented to repress autophagy and induce apoptosis of human colon cancer HT-29 cells alone with downregulation of the PI3K/Akt signaling pathway [67]. Another mechanism of action of genistein is that it also exerts an antiproliferative effect on ovarian cancer cells through autophagy as well as apoptosis [67]. Genistein augments miR-451 expression and improves isoproterenol-mediated cardiac hypertrophy [68]. Furthermore, genistein increases miR-1469 expression and prevents Mcl-1 expression in laryngeal cancer [69].
Indole-3-carbinol: Indole-3-carbinol, a natural compound collected from cruciferous vegetables, usually inhibits the PI3K/Akt pathway along with blockage of autophagy signaling in human colon cancer HT-29 cells [67].
Plumbagin: Plumbagin is obtained from Plumbago zeylandica L. root and has a prompt antiproliferative activity through autophagic cell death, mediated via inhibition of AKT/mTOR pathway in human cervical cancer cells [70].

4.2. Induction of Autophagy by Natural Compounds as Potential Cancer Therapy

Antroquinonol: Antroquinonol, produced from Antrodia Camphorata, shows anti-tumor activity in various cancer cells [71]. Antroquinonol has been demonstrated to exhibit inhibitory effects on non-small cell lung cancer (NSCLC), as well as human pancreatic cancer, PANC-1 and ASPC-1, via the PI3K/Akt/mTOR pathway [72]. The anti-proliferative effect of antroquinonol is related to apoptosis, autophagic cell death, and rapid senescence of the pancreatic cells [72]. Several studies have demonstrated that antroquinonol can be introduced for the treatment of neoplasms [73]. A proposed clinical trial in metastatic pancreatic carcinoma is aimed at evaluating the anticancer activity of antroquinonol in combination with nab-PTX and gemcitabine [74].
18α-Glycyrrhetinic acid: 18α-Glycyrrhetinic acid (18-GA), a known gap-junction inhibitor, has been shown to demonstrate anticancer properties in human NB cells [42]. 18-GA has been shown to induce autophagy as well as apoptosis. Autophagy increases Atg5, Atg7, and LC3II along with degradation of p62 (Figure 4). Furthermore, 18-GA, along with the autophagy inhibitor CQ, induces significant cell death in NB cells. Moreover, the Bcl-2/Beclin-1 interaction in addition to the cleavage of Beclin-1 has been revealed by 18-GA in autophagy and apoptosis induced cell death. Caspase-3 siRNA and pan-caspase inhibitor treatment have been demonstrated to prevent 18-GA-induced cellular cytotoxicity, indicating that caspase-mediated apoptosis induction was observed in human NB cells [42].
Celastrol: Celastrol, a popular natural medicinal compound triperine secreted from Tripterygium wilfordii Hook, is a polyubiquitinated aggregate that degrades the autophagy substrate p62 in the human glioblastoma (GBM) cancer cells [75]. Celastrol-mediated cytotoxicity was not sensitized by adjustment with autophagy inhibitors. Paraptosis-like cytoplasmic vacuolization can be induced by celastrol and has been associated with autophagy and the initiation of apoptosis in PC-3, A549, and HeLa cancer cells [76]. Celastrol can be used to treat fever, joint pain, and edema without causing any side effects [77]. Recently, many research studies have identified celastrol as a neuroprotective agent via a collaborative drug screen that can be used for the treatment of many neurodegenerative diseases.
Monanchocidin A: An active novel alkaloid, monanchocidin A, has been recently isolated from Monanchora pulchra, a marine sponge [78]. Monanchocidin can initiate autophagy and lysosomal membrane permeabilization in tumor cells [78]. Many studies have demonstrated that monanchocidin A can inhibit human urogenital cancers, including germ cell tumors, through autophagy signaling [79].
Paclitaxel: PTX, a natural medicinal compound isolated from Taxus brevifolia Nutt, shows a vast range of effect against various types of cancer such as breast, endometrial, bladder, and cervical cancer, by influencing the autophagy pathway [80]. The US Food and Drug Administration (FDA) has approved PTX for the treatment of ovarian cancer and early stage breast cancer. The anticancer activity of PTX is increased when PTX is combined with cisplatin or carboplatin, and the combined compound can be used for the early treatment of advanced ovarian cancer [81]. It has been shown that PTX can inhibit the initial stage of autophagy, along with apoptosis, under normoxic and hypoxic conditions in human breast cancer cells [82]. PTX-initiated apoptosis was combined with the initiation of autophagy in A549 lung cancer cells [47]. Treatment with PTX also initiated acidic vesicular organelle formation, Beclin-1, Atg5, and LC3 expression, and the prevention of autophagy proceeds via PTX-mediated cell death [83,84]. Autophagy can promote PTX-mediated cell death. In human breast cancer cells, PTX prevents autophagy via an individual mechanism based on the cell-cycle phase [84].
Gintonin: Gintonin (GT), a novel ginseng-derived exogenous ligand of lysophosphatidic acid (LPA) receptors, has shown to induce autophagy in cortical astrocytes [45]. GT intensely augmented the autophagy marker LC3 via the G protein-coupled LPA receptor-mediated pathway. However, GT-mediated autophagy was considerably decreased via autophagy inhibition 3-MA as well as Beclin-1, Atg5, and Atg7 gene knockdown [15,45]. Notably, pretreatment with a lysosomotropic agent, Baf A1 and E-64d/peps A, GT significantly improved LC3-II levels in addition to the formation of LC3 puncta (Figure 5). Furthermore, GT treatment improved autophagic flux, which influenced lysosome-associated membrane protein 1 (LAMP1) in addition to degradation of the autophagy substrate ubiquitinated p62/SQSTM1 protein in mouse cortical astrocytes [45].
β-elemene: β-elemene, derived from Rhizoma Curcumae, is a natural terpenoid bioactive compound that shows strong activity against various types of cancers [40]. In the human body, β-elemene works against the PI3K/Akt/mTOR/p70S6K pathway that induces autophagy and apoptosis of human NSCLCA549 cells [85]. As β-elemene increases the antitumor effect, autophagy is prevented by the interception with chlorochine [85]. In human renal-cell carcinoma 786-0 cells, β-elemene is used for the inhibition of the MAPK/ERK and P13K/Akt/mTOR signaling pathways to initiate defensive autophagy and apoptosis [86].
Caffeine: In HeLa cells, the Akt/mTOR/p70S6K pathway initiates autophagy and apoptotic cell death under the influence of caffeine that is routinely consumed in beverages [87].
Curcumin: Curcumin is a non-toxic toxic polyphenol with yellow pigment isolated from the rhizome of Curcuma longa L. [88]. Curcumin-associated autophagy is considered to be a signal for cellular death in various types of cancer cells in the human body. Nevertheless, in the human body, curcumin can induce cellular differentiation and cellular survival by regulating the activities of AKT-AMPK-mediated autophagy induction [89]. Thus, it prevents the growth of malignant glioma cells in vitro and in vivo. Curcumin is used to induce non-apoptotic autophagy cell death in U87-MG and U373-MG malignant glioma cells [90]. This activity is related to the G2/M phase cell cycle attack, the prevention of the ribosomal S6 protein kinase pathway with the ERK1/ERK2 activation pathway [90]. Curcumin is responsible for Atg apoptosis in mesothelioma and k562 long-standing myelogenous leukemia cells by AKT/mTOR and NF-κB signaling pathways [91]. Curcumin can be used as an effective therapeutic compound in cancer management and is well endured, as confirmed by preclinical data in clinical experiments [88]. Curcumin has been found to suppress TGF-β expression, such as p-SMAD2, TGF-β3, MMP-13, and NF-κB tumor-promoting factors as well as metastatic active adhesion molecules, such as intercellular adhesion molecule 1 (ICAM-1) and β1-integrin in stromal fibroblasts and colorectal cancer cells [92]. In addition, MiR-1246 has been shown to be involved in curcumin-induced radiosensitizing actions on bladder cancer cells by targeting p53 translation [93].
Tetrahydrocurcumin: Tetrahydrocurcumin (THC) is one of the strongest bioactive natural metabolite derived from curcumin. THC effectively shows strong antioxidant, anticancer, and cardioprotective properties [94]. The action of THC on autophagy was assessed by the decrease in the PI3K/Akt-mTOR and MAPK signaling pathways and initiation of caspase-7 mediated apoptosis [95].
Oxyresveratrol: Oxyresveratrol (Oxy R), 4-[(E)-2-(3,5-dihydroxyphenyl) ethenyl] benzene-1,3-diol/ 2,3,4,5-tetrahydroxy-trans-stilbene found in Morus alba, has been shown to activate autophagy and apoptotic cell death in NB cells [51] (Figure 6). Oxy R-mediated autophagic cells were independent of apoptotic cell death. The PI3K/AKT/mTOR and p38 MAPK pathways control Oxy R-induced cell death in SH-SY5Y human NB cells [51].
Epigallocatechin-3-gallate: The activity of cisplatin and oxaliplatin-initiated autophagy was promoted by EGCG in HT-29 and DLD-1 colorectal cancer cells, as identified by the multiplication of LC3B-II protein and autophagosome organization [96]. EGCG initiated autophagy as well as apoptosis in SSC-4 squamous cell carcinoma in combination with the upregulation of FAS, BAK, BAD, IGF-IR, WNTll, and ZEBI proteins along with downregulation of MYC, TP53, and CASP8 proteins [97]. The formation of autophagosomes is increased by EGCG. In hepatic cells, the processes of lysosomal acidification and autophagic flux are also increased by EGCG in vitro, as well as in vivo [98]. Remarkably, luteolin and EGCG combination repressed TGF-β-mediated demonstration of myofibroblast phenotypes by reducing RhoA as well as ERK activation [99]. EGCG suggestively repressed rat hepatic stellate cells proliferation through hindering expression of PDGF-β receptor and tyrosine phosphorylation [100]. Furthermore, it has been found that EGCG attenuates uric acid-mediated NRK-49 F cell injury by up-regulating miR-9, and successively by JAK-STAT as well as NF-κB signaling pathway activation [101]. EGCG reduces carcinoma cell growth probably by regulating miRNA expression, and may be a potential beneficial target for the prevention of cancers.
Fisetin: Fisetin (3,7,3′,4′-tetrahydroxyflavone), a flavonol and a member of flavonoid polyphenols [102], is usually isolated from many vegetables and fruits [102]. Fisetin can function as an inhibitor of the expression of PI3K/Akt/mTOR pathways and regulate autophagy in prostate cancer and human NSCLC cells [103,104]. Fisetin has been found to inhibit the mTOR pathway and induce autophagy in human prostate cancer cells [105]. In addition, the effects of fisetin on autophagy are cell-type specific because fisetin inhibits autophagy in MCF-7 breast cancer cells and induces caspase-7-mediated apoptosis [106].
γ-Tocotrienol: Tocotrienols and the isoforms of vitamin E are differentiated on the basis of their antioxidant, anti-inflammatory, and anticancer properties. γ-tocotrienol induces endoplasmic inflammation and autophagy-associated cell death [107]. Pretreatment with the autophagy inhibitors 3-MA or Baf1 prevented the γ-tocotrienol-initiated cytotoxicity [108]. Many other studies have demonstrated that in mouse mammary cancer cells, the attachment of γ-tocotrienol with oridonin effectively initiated both autophagic and apoptotic effects [109]. The potency of autophagy cellular markers was significantly enhanced by the attachment of γ-tocotrienol with oridonin [110]. Simultaneously, this potency also improved the transformation of LC3-I to LC3-II, Atg3, Atg7, Beclin-1, Atg5-Atg12, LAMP-1, and cathepsin-D.
Geraniol: Fruits, vegetables, and cereal grains contain geraniol, a secondary metabolite, that in human PC-3 prostate cancer cells was shown to induce apoptosis and Atg cell death [111].
Seriniquinone: Seriniquinone is a medicinal compound derived from a marine bacterium of the genus Serinicoccus. It exhibits effective anti-proliferative activity against melanoma cell lines by activation of autophagocytosis [112].
Thymoquinone: Thymoquinone (TQ) is a key bioactive natural product isolated from black cumin, Nigella sativa L. TQ is related to increased volumes of autophagic vacuoles, as well as LC3 proteins, and enhances the combination of autophagosomes in the head and neck squamous carcinoma cells [113]. TQ effectively reduced the growth rate of GBM cells in association with the lysosomal inhibitor CQ, mostly causing apoptosis, and initiating autophagy [114]. LC3-II and p62 proteins are enhanced by the expression of TQ [114]. It was also shown that TQ prevented the growth of irinotecan-inhibited LoVo colon cancer cells by primarily activating apoptosis before autophagy [115]. Activation of the p38 and JNK MAP kinase pathways is related to TQ, which causes autophagic cell death [115]. TQ-induced caspase and autophagic cell death were also related to mitochondrial outward membrane penetrability [114].
Magnolol: Magnolol, isolated from Magnolia officinalis, exhibited anti-cancer and anti-tumor activity by directing autophagy and apoptosis signaling pathways [116]. However, magnolol promotes autophagy-mediated cell death in human non-small lung cancer H460 cells at high concentrations [117]. Additionally, in SGC-7901 human gastric adenocarcinoma cells, magnolol has been shown to repress PI3K/Akt pathway and encourage death by inducing autophagy [118].
Polyphenol mulberry extract: Mulberry is derived from Morus Alba leaf, which modulates the autophagy AMPK/PI3K/Akt pathway [119]. In Hep3B human hepatocellular carcinoma cells, mulberry extract modulates autophagic or apoptotic cell death by the activation of the p53-dependent pathway [120]. It has been found that in NB cells, Morus alba root extract initiates apoptosis through FOXO-Caspase-3 dependent pathway [121].
Oblongifolin C: Oblongifolin C (OC), isolated from Garcinia yunnanensis Hu, is a strong autophagy inhibitor [122]. Treatment with OC resulted in an augmented number of autophagosomes as well as reduced SQSTM1/p62 and degradation [123]. OC showed anticancer effectiveness by improved staining of LC3 puncta, SQSTM1, and cleaved CASP-3, along with decreased expression levels of lysosomal cathepsins in an in vivo xenograft mouse model [122].
Naphthazarin: Naphthazarin repressed the Akt/PI3K pathway by activating autophagy and apoptosis pathways and by inducing A549 lung cancer cell death [124].
Sulforaphane: Sulforaphane is derived from cruciferous vegetables, such as cauliflower, cabbage, broccoli, and hoary weed [125]. In human prostate cancer PC-3 cells, sulforaphane has been shown to prompt autophagosome formation as well as acidic vesicular organelle formation [126]. Sulforaphane augmented the protein levels of LC3B-I in addition to encouraging LC3B-II processing. Exposure to sulforaphane in tumor cells displayed LC3B-II puncta related to autophagosome formation [127]. It also disrupted BECN1/BCL-2 interaction, causing autophagy initiation through the release of BECN1. Phosphorylated AKT-Ser473 levels decreased by sulforaphane in addition to simultaneous treatment with autophagy inhibitors, 3-MA or CQ, and inhibited tumor cell growth and proliferation [128]. Sulforaphane initiation decreases oxidative stress in addition to misfolded protein accumulation [129].
Mollugin: A bioactive phytochemical, mollugin, sequestered from Rubia cordifolia L., showed anticancer potential against numerous cancer cells [130]. Additional studies confirmed that the mTOR and ERK pathways are involved in mollugin-prompted autophagy in addition to apoptosis [131].
Jujuboside B: Jujuboside B is a saponin found in the seeds of Zizyphus jujubavar Spinose. It increased autophagy and apoptosis in human HCT-116 and AGS gastric adenocarcinoma cells in vitro and successfully repressed tumor growth in a xenograft model of nude mice in vivo [132]. Furthermore, jujuboside B activated p38/JNK-mediated autophagy induction through cytoplasmic vacuole formation in addition to LC3-I/II conversion [132].
Cyclovirobuxine D: Cyclovirobuxine D (CVB-D) has been shown to activate autophagy in the human MCF-7 breast cancer cell line through the addition of autophagosomes as well as raised LC3 puncta, along with augmented conversion of LC3-I to LC3-II [133]. However, CVB-D-mediated autophagy induction and cell viability were blocked by 3-MA [133].
Polygonatum odoratum lectin: Polygonatum odoratum lectin (POL) has been shown to reveal apoptosis-inducing and anti-proliferative properties in a wide range of cancer cells [134]. In human breast cancer NSCLC A549 and MCF-7 cells, POL prompted both autophagy and apoptosis [135]. POL stimulated apoptosis by preventing the AKT/NF-κB pathway, whereas augmented autophagy through AKT-mTOR pathway suppression [135].
Resveratrol: Resveratrol, a polyphenol compound derived from berries, red grapes, and peanuts, has been shown to facilitate cell death in a diverse variety of cancer cells through autophagy, apoptosis, and necrosis [136,137]. In ovarian cancer cells, resveratrol has been found to activate autophagic cell death [138]. However, resveratrol inhibited NF-κB stimulation in association with augmented permeability of lysosomal in cervical cancer cells, demonstrating autophagic cell death [139]. Additionally, in HL-60 promyelocytic leukemia cells, it has been described that resveratrol-prompted apoptosis increased LC3-II levels in addition to dependency on mTOR pathway [140]. Meanwhile, resveratrol-mediated apoptosis was related to reduced p53 and AMPK/mTOR autophagy signaling pathways in renal carcinoma cells [141]. The antiproliferative action of resveratrol has been found in liver myofibroblasts by preventing PDGF signaling and decreasing EGF-dependent DNA synthesis [142]. Additionally, in retinal endothelial cells, resveratrol has been found to increase miR15a expression under conditions of high glucose levels and decreased insulin signaling [143].
Quercetin: Quercetin, a bioflavonoid, is generally found in fruits, beverages, and vegetables. In vitro and in vivo studies have shown that it exhibits antiproliferative activities in numerous tumors conventionally related to its antioxidant properties [144].
Rottlerin: Rottlerin, a natural product sequestered from Mallotus philippinensis, repressed PI3K/mTOR signaling in human pancreatic cancer stem cells, and activated autophagy-mediated apoptotic cell death [145].
Angelicin: Angelicin, a psoralen, is derived from Angelica polymorpha and has been shown to induce apoptosis [146,147] and autophagy [148]. Angelicin increased autophagy-related proteins Atg3, Atg7 and Atg12-5 with phosphorylation of mTOR [148,149].
Ursolic acid: A triterpenoid phytochemical, ursolic acid (UA), has been found to activate autophagy by inducing LC3 in addition to p62 protein accumulation [150]. In HCT15 cells, UA repressed growth by modulating autophagy in the JNK pathway [151]. However, UA induced cytotoxicity in addition to repressing concentration-dependent TC-1 cervical cancer cells by controlling Atg5 and LC3-II, which demonstrated its anticancer activity [150]. Furthermore, in MCF7 breast cancer cells, UA activated autophagy under ER stress [152]. In PC3 prostate cancer cells, UA-induced autophagy was facilitated by the Akt/mTOR and Beclin-1 pathways [153].
Polygonatum Cyrtonema Lectin: Polygonatum Cyrtonema Lectin (PCL) inhibits the PI3K-Akt pathway in addition to prompting autophagy and apoptosis in cancer cells [154]. Although PCL significantly prevented the growth of A375 human melanoma cells, normal melanocytes were not affected [155]. Additionally, PCL has been found to stimulate both autophagy and apoptosis in A375 cells [156].
Silibinin: Silibinin, derived from Silybum marianum, has been shown to induce autophagic HT1080 cell death in human fibrosarcoma through diverse mechanisms, including inhibition of MEK/ERK and PI3K/Akt pathways [157].
Plant extract induces autophagy: Several plant-derived extracts from Dioscorea nipponica Makino, Melandrium firmum, marine algae, and natural flavonoids have been found to induce apoptosis and autophagic properties [158,159,160,161]. Saussurea lappa ethanol extract has been investigated to activate apoptosis in NB [162] and autophagy in LNCaP prostate cancer cells [163].
Ginsenoside Rk1: Ginsenoside Rk1, an NMDA receptor inhibitor [164], exhibited antitumor and autophagy modulation activities in HepG2 cells. Rk1-prompted autophagy was recognized through LC3-I to LC3-II conversion, which incorporated lysosomes into the autolysosome process [165]. However, the autophagy inhibitor, Beclin 1 siRNA or bafilomycin A1, and Rk1 combination boosted autophagic activities in HepG2 [165].
Climacostol: In mouse B16-F10 melanoma cells, autophagosomes accumulate with dysfunctional autophagic degradation. However, climacostol provides mechanistic insights and promotes autophagosome turnover through the p53-AMPK pathway and activated p53 protein levels [166].

5. Role of Natural Compounds in Autophagy Modulation of Neurodegenerative Diseases

Natural plant-derived compounds are able to modulate autophagy and can be used to develop treatments for neurodegenerative diseases such as Alzheimer’s disease (AD), Parkinson’s disease (PD), Huntington’s disease (HD), Amyotrophic lateral sclerosis (ALS), Spinocerebellar ataxia (SCA), and NB [167]. Oleuropein aglycone (OLE), a natural phenol, can protect the Aβ-driven cytotoxicity of neuronal cells by stimulating autophagy found in in vitro and in vivo examinations, thereby easing Aβ toxicity clearance of AD. [168,169]. Arctigenin, derived from Arctium lappa (L.), inhibits Aβ production by decreasing BACE1 and boosts Aβ clearance by increasing autophagy via AKT/mTOR inhibition [170]. Another polyphenol, resveratrol, significantly represses Aβ aggregation via autophagy and shows antioxidant effects in AD [171,172,173]. Kaempferol, a flavone, increases autophagy and encourages mitochondrial damage repair, protecting neuronal cells against death via rotenone-induced toxicity in PD models [174]. Uncaria rhynchophylla contains alkaloids of oxindole and was found to prompt Beclin-1-dependent as well as mTOR-independent autophagy, which stimulated α-synuclein clearance in the Drosophila model of A53T, both wild-type and mutant [175]. Conophylline, Tavertaemontana divaricate-derived vinca alkaloid, has been shown to promote α-synuclein degradation via the mTOR-independent pathway in MPP+-mediated neurotoxin [176]. Curcumin, a polyphenol from Curcuma longa turmeric plant, has been shown to exhibit neuroprotective properties against PD models via mTOR-dependent autophagy and preventing oxidative stress, inflammation, and α-synuclein aggregation. [177]. Trehalose, resveratrol, and onjisaponin B improve autophagic degradation of α-synuclein when treating MPTP-induced mice, in addition to stimulating AMPK and sirtuin 1 protein [178,179]. A natural alkaloid, harmine, has been found to stimulate α-synuclein removal by PKA-mediated PD pathogenesis of UPS activation [180]. Distinct AD, PD, and HD neurodegenerative syndromes might be treated via natural compounds through modulation of autophagy [1]. Nelumbo nucifera contains neferine, which has been shown to protect Htt mutant proteins via autophagy triggering with the AMPK/mTOR pathway [181]. Aggregate-prone mutant Htt protein has been removed by trehalose through an mTOR-independent autophagy pathway [182]. Berberine, an isoquinoline alkaloid, has been shown to prevent mutant Htt protein accumulation and activate autophagy in mouse and cell models of HD [183]. In particular, trehalose has been found to stimulate mTOR-independent autophagy in superoxide dismutase mutant mice in the management of ALS in vivo [184]. Trehalose and its analogs, such as lactulose and melibiose, have been determined to substantially reduce the aggregation of abnormal ataxin-3 by stimulation of autophagy, which reduces free radical production [185], indicating that natural compounds might be used as a therapeutic approach to control SCA.

6. Effect of Natural Compounds on Solid Tumors and Lymphomas

Neuroblastoma, the most common extracranial solid malignant tumor of childhood, is the third most common form of pediatric malignancy worldwide. Didymin, a dietary flavonoid glycoside, has been found to be effective in NB treatment [186]. A Morus alba root extract compound, Oxy R, has been shown to accumulate ROS as well as inducing autophagic and apoptotic cell death in NB cells via the FOXO and caspase-3 induction pathways [51,121]. A psoralen compound angelicin derived from Angelica polymorpha increases cytotoxicity in addition to inducing apoptosis through anti-apoptotic proteins (Bcl-xL, Bcl-2, and Mcl-1) downregulation in SH-SY5Y human NB cells [146,147]. However, Saussurea lappa Clarke, Dioscorea nipponica Makino, and Melandrium firmum extracts have been shown to have apoptotic and anti-proliferative effects in NB cells [158,162,187]. Additionally, curcumin, propolis, quercetin, icariin, withaferin A, tetrandrine, and resveratrol have been used for the treatment of aggressive GBM [188]. In addition, a combination therapy with temozolomide and resveratrol substantially diminished the growth of GBM cells in a xenograft model [189]. Natural compounds such as PTX, wortmannin, and beta-lapachone have been found to be involved in the ARF/MDM2-MDMX/p53 signaling pathway in the human retinoblastoma (RB) Y79 cell line [190].
Lymphoma is the most common type of blood cancer, and various natural compounds have been used to treat lymphomas. Several natural compounds have shown anticancer effects on lymphomas (Table 2). Nigella sativa Linn. containing TQ has been shown to induce apoptosis and ROS production at 5 and 10 mM concentrations of 24 h treatment in ABC-DLBCL activated B cell lymphoma cell lines [191]. It has been stated that fuxocanthinol induces apoptosis in addition to arrest in cell cycle progression in lymphoma cells [192].
Curcumin has been reported to induce antitumor effects on CH12F3 cell lines through DNA breaks and apoptosis [193]. An extract from Peperomia tetraphylla, Peperobtusin A, has been found to induce S phase cell cycle arrest and p38 MAPK-dependent apoptosis in U937 cells [194]. However, induction of apoptosis, necrosis, and cell cycle arrest by psilostachyin C has been found in lymphoma BW5147 cell lines [196]. In a xenograft mouse model, 11(13)-dehydroivaxillin (DHI) has shown antitumor activity via the apoptosis pathway [195].

7. Therapeutic View of Autophagy in Heart/Cardiovascular Diseases

Autophagy has been shown to play a major regulatory role in cellular longevity [197] and in heart/cardiovascular diseases towards maintaining homeostasis by conserving cardiac function and structure [198] and determining a potential connection between ventricular fibrillation and autophagy [199]. There is evidence suggesting that mitochondrial autophagy downregulation plays a vital role in mitochondrial dysfunction, as well as heart failure, while re-establishment of mitochondrial autophagy mitigates this heart dysfunction [200]. It has been revealed that cardiac myocytes are reduced through autophagy inhibitor, 3-MA, with AMP-activated protein kinase (AMPK) activation via inactivation of mTOR under glucose deprivation [201]. In addition, the autophagy-related gene, Beclin-1, protected the heart in LPS-induced sepsis in a mouse model and targeted autophagy induction, and therefore has an important role with therapeutic potential for treating cardiovascular diseases [202]. Under high glucose conditions, it has been found that heme-oxygenase-1 (HO-1) overexpression via human HO-1 recombinant plasmid inhibited dysfunction of cardiac arrest, improving autophagy levels. Numerous hemin concentrations induce HO-1, which affects endothelial cell mitochondrial dysfunction. Likewise, exposure to hemin further prompts mitophagy, although it is not adequate to inhibit cell death [203]. It has been found that regulating Sirt3 concentration in myocytes via proper autophagy expression in diverse stages of myocardial ischemia-reperfusion might effectively decrease the morbidity of patients with myocardial infarction in the future [204]. Metformin has been used to treat cardiotoxicity induced by doxorubicin via the autophagy pathway [205]. In contrast, a non-selective β-adrenergic agonist, isoproterenol, has been investigated for the treatment of autophagy as well as cell death pathway induced cardiac injury such as necrosis or apoptosis induced cardiac injury [206]. Currently, it is accepted that pharmacological approaches targeting autophagy and apoptosis are useful in the treatment of coronary heart disease, therefore providing new favorable therapeutic interventions in the future [207]. Hence, consideration of emerging molecular mechanisms towards the understanding of autophagy has recently gained importance for regulating cellular systems that can be used to prevent and treat cardiovascular disease [208].

8. Perspectives of Naturally Occurring Autophagy Modulators in Cancer Therapy

Despite anti-cancer drug development, natural compounds contribute to cancer and cancer stem cell [209] suppressive action without any severe complications, and thus it is an emerging topic with respect to autophagy. Natural compounds have also appeared as unique therapeutic agents for cancer and drug repositioning via their impact on autophagy [210]. It is remarkable that the study of cell fate by natural stimulators should be the focal point, although technological dissimilarities in autophagy detection may affect the outcome. In addition, in vivo examination of natural stimulators of autophagy in cancer therapy or prevention will be an emerging field of research in the future. Moreover, studies on the effects of natural product-derived molecules on autophagy should be highlighted at the translational level.

9. Conclusions and Future Directions

Cancer cells can convert and reuse necessary amino acids by imposing autophagic signaling pathways, thereby ensuring their consistent growth and longevity. Accordingly, autophagy serves a dual role in cancer—augmentation or hindrance—and can either boost or down-regulate cancer cell proliferation. As a result, autophagy moderators can play the role of a promising novel therapeutic approach towards cancer inhibition by helping overcome resistance against chemotherapy and radiotherapy. Potential targets of autophagic progress and autophagosome formation in the initial phase and lysosomal deterioration in the late phase have been considered. However, mTOR and AMPK are the leading regulatory molecules of autophagy in upstream signaling pathways. These are also known targets for natural compound derived modulators of autophagy. Therefore, using plant-derived and semisynthetic components to target the regulation of the entire pathway can provide an opportunity to design a powerful remedy for cancer patients.

Author Contributions

Conceptualization by M.A.R. Writing and original draft preparation by M.S.H., P.B., R.I., M.H.R. (MD. Hasanur Rahman) Figures are drawing by M.H.R. (Md. Habibur Rahman). Review and editing by M.J.U. Visualization and supervision by H.R. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by NRF Research Program (2016M3C7A1913845) and the Korea Research Fellowship (KRF) Program (M.A.R. 2016H1D3A1908615; R.I. 2020H1D3A1A04104782) funded by the Ministry of Science and ICT, Republic of Korea, and also supported by (M.J.U. 2020R1I1A1A01072879 and 2020H1D3A2A02110924), Republic of Korea.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Rahman, M.A.; Rhim, H. Therapeutic implication of autophagy in neurodegenerative diseases. BMB Rep. 2017, 50, 345–354. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Corti, O.; Blomgren, K.; Poletti, A.; Beart, P.M. Autophagy in neurodegeneration: New insights underpinning therapy for neurological diseases. J. Neurochem. 2020, e15002. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Uddin, M.S.; Stachowiak, A.; Al Mamun, A.; Tzvetkov, N.T.; Takeda, S.; Atanasov, A.G.; Bergantin, L.B.; Abdel-Daim, M.M.; Stankiewicz, A.M. Autophagy and Alzheimer’s Disease: From Molecular Mechanisms to Therapeutic Implications. Front. Aging Neurosci. 2018, 10. [Google Scholar] [CrossRef] [PubMed]
  4. Arroyo, D.S.; Gaviglio, E.A.; Ramos, J.M.P.; Bussi, C.; Rodriguez-Galan, M.C.; Iribarren, P. Autophagy in inflammation, infection, neurodegeneration and cancer. Int. Immunopharmacol. 2014, 18, 55–65. [Google Scholar] [CrossRef] [Green Version]
  5. Perry, R.J.; Shulman, G.I. Mechanistic Links between Obesity, Insulin, and Cancer. Trends Cancer 2020, 6, 75–78. [Google Scholar] [CrossRef]
  6. Ling, Y.; Perez-Soler, R. Disruption of autophagic and autolysosomal signaling pathways leads to synergistic augmentation of erlotinib-induced apoptosis in wild type EGFR human non-small cell lung cancer cell lines. EJC Suppl. 2010, 8, 183. [Google Scholar] [CrossRef]
  7. Towers, C.G.; Wodetzki, D.; Thorburn, A. Autophagy and cancer: Modulation of cell death pathways and cancer cell adaptations. J. Cell Biol. 2020, 219. [Google Scholar] [CrossRef]
  8. Manic, G.; Obrist, F.; Kroemer, G.; Vitale, I.; Galluzzi, L. Chloroquine and hydroxychloroquine for cancer therapy. Mol. Cell. Oncol. 2014, 1, e29911. [Google Scholar] [CrossRef]
  9. Khandia, R.; Dadar, M.; Munjal, A.; Dhama, K.; Karthik, K.; Tiwari, R.; Yatoo, M.I.; Iqbal, H.M.N.; Singh, K.P.; Joshi, S.K.; et al. A Comprehensive Review of Autophagy and Its Various Roles in Infectious, Non-Infectious, and Lifestyle Diseases: Current Knowledge and Prospects for Disease Prevention, Novel Drug Design, and Therapy. Cells 2019, 8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Bialik, S.; Dasari, S.K.; Kimchi, A. Autophagy-dependent cell death—Where, how and why a cell eats itself to death. J. Cell Sci. 2018, 131. [Google Scholar] [CrossRef] [Green Version]
  11. Lagoa, R.; Silva, J.; Rodrigues, J.R.; Bishayee, A. Advances in phytochemical delivery systems for improved anticancer activity. Biotechnol. Adv. 2020, 38. [Google Scholar] [CrossRef]
  12. Tavakol, S.; Ashrafizadeh, M.; Deng, S.; Azarian, M.; Abdoli, A.; Motavaf, M.; Poormoghadam, D.; Khanbabaei, H.; Afshar, E.G.; Mandegary, A.; et al. Autophagy Modulators: Mechanistic Aspects and Drug Delivery Systems. Biomolecules 2019, 9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Hirata, E.; Ohya, Y.; Suzuki, K. Atg4 plays an important role in efficient expansion of autophagic isolation membranes by cleaving lipidated Atg8 in Saccharomyces cerevisiae. PLoS ONE 2017, 12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Chen, Y.; He, J.; Tian, M.; Zhang, S.Y.; Guo, M.R.; Kasimu, R.; Wang, J.H.; Ouyang, L. UNC51-like kinase 1, autophagic regulator and cancer therapeutic target. Cell Proliferat. 2014, 47, 494–505. [Google Scholar] [CrossRef] [PubMed]
  15. Rahman, M.A.; Rahman, M.R.; Zaman, T.; Uddin, M.S.; Islam, R.; Abdel-Daim, M.M.; Rhim, H. Emerging Potential of Naturally Occurring Autophagy Modulators Against Neurodegeneration. Curr. Pharm. Des. 2020, 26, 772–779. [Google Scholar] [CrossRef]
  16. Nakamura, S.; Yoshimori, T. New insights into autophagosome-lysosome fusion. J. Cell Sci. 2017, 130, 1209–1216. [Google Scholar] [CrossRef] [Green Version]
  17. Schulze, H.; Kolter, T.; Sandhoff, K. Principles of lysosomal membrane degradation Cellular topology and biochemistry of lysosomal lipid degradation. BBA-Mol. Cell Res. 2009, 1793, 674–683. [Google Scholar] [CrossRef] [Green Version]
  18. Shpilka, T.; Weidberg, H.; Pietrokovski, S.; Elazar, Z. Atg8: An autophagy-related ubiquitin-like protein family. Genome Biol. 2011, 12. [Google Scholar] [CrossRef]
  19. Agrotis, A.; von Chamier, L.; Oliver, H.; Kiso, K.; Singh, T.; Ketteler, R. Human ATG4 autophagy proteases counteract attachment of ubiquitin-like LC3/GABARAP proteins to other cellular proteins. J. Biol. Chem. 2019, 294, 12610–12621. [Google Scholar] [CrossRef] [Green Version]
  20. Lystad, A.H.; Simonsen, A. Mechanisms and Pathophysiological Roles of the ATG8 Conjugation Machinery. Cells 2019, 8. [Google Scholar] [CrossRef] [Green Version]
  21. Ward, C.; Martinez-Lopez, N.; Otten, E.G.; Carroll, B.; Maetzel, D.; Singh, R.; Sarkar, S.; Korolchuk, V.I. Autophagy, lipophagy and lysosomal lipid storage disorders. BBA-Mol. Cell Biol. Lipids 2016, 1861, 269–284. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Galluzzi, L.; Pietrocola, F.; Bravo-San Pedro, J.M.; Amaravadi, R.K.; Baehrecke, E.H.; Cecconi, F.; Codogno, P.; Debnath, J.; Gewirtz, D.A.; Karantza, V.; et al. Autophagy in malignant transformation and cancer progression. Embo J. 2015, 34, 856–880. [Google Scholar] [CrossRef]
  23. Desantis, V.; Saltarella, I.; Lamanuzzi, A.; Mariggio, M.A.; Racanelli, V.; Vacca, A.; Frassanito, M.A. Autophagy: A New Mechanism of Prosurvival and Drug Resistance in Multiple Myeloma. Transl. Oncol. 2018, 11, 1350–1357. [Google Scholar] [CrossRef] [PubMed]
  24. Yang, Z.N.J.; Chee, C.E.; Huang, S.B.; Sinicrope, F.A. The Role of Autophagy in Cancer: Therapeutic Implications. Mol. Cancer Ther. 2011, 10, 1533–1541. [Google Scholar] [CrossRef] [Green Version]
  25. Ricci, M.S.; Zong, W.X. Chemotherapeutic approaches for targeting cell death pathways. Oncologist 2006, 11, 342–357. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Mowers, E.E.; Sharifi, M.N.; Macleod, K.F. Autophagy in cancer metastasis. Oncogene 2017, 36, 1619–1630. [Google Scholar] [CrossRef] [PubMed]
  27. Altman, B.J.; Rathmell, J.C. Metabolic Stress in Autophagy and Cell Death Pathways. CSH Perspect. Biol. 2012, 4. [Google Scholar] [CrossRef] [PubMed]
  28. Mowers, E.E.; Sharifi, M.N.; Macleod, K.F. Functions of autophagy in the tumor microenvironment and cancer metastasis. FEBS J. 2018, 285, 1751–1766. [Google Scholar] [CrossRef] [Green Version]
  29. Park, S.A.; Surh, Y.J. Modulation of tumor microenvironment by chemopreventive natural products. Ann. N. Y. Acad. Sci. 2017, 1401, 65–74. [Google Scholar] [CrossRef]
  30. Zhang, B.; Tian, L.; Xie, J.; Chen, G.; Wang, F. Targeting miRNAs by natural products: A new way for cancer therapy. Biomed. Pharmacother. 2020, 130, 110546. [Google Scholar] [CrossRef]
  31. Jiang, J.W.; Zhang, L.; Chen, H.N.; Lei, Y.L.; Zhang, T.; Wang, Y.L.; Jin, P.; Lan, J.; Zhou, L.; Huang, Z.; et al. Regorafenib induces lethal autophagy arrest by stabilizing PSAT1 in glioblastoma. Autophagy 2020, 16, 106–122. [Google Scholar] [CrossRef] [PubMed]
  32. Cuomo, F.; Altucci, L.; Cobellis, G. Autophagy Function and Dysfunction: Potential Drugs as Anti-Cancer Therapy. Cancers 2019, 11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Livesey, K.M.; Tang, D.L.; Zeh, H.J.; Lotze, M.T. Autophagy inhibition in combination cancer treatment. Curr. Opin. Investig. Drugs 2009, 10, 1269–1279. [Google Scholar] [PubMed]
  34. Wu, Y.T.; Tan, H.L.; Shui, G.H.; Bauvy, C.; Huang, Q.; Wenk, M.R.; Ong, C.N.; Codogno, P.; Shen, H.M. Dual Role of 3-Methyladenine in Modulation of Autophagy via Different Temporal Patterns of Inhibition on Class I and III Phosphoinositide 3-Kinase. J. Biol. Chem. 2010, 285, 10850–10861. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Vazquez-Martin, A.; Oliveras-Ferraros, C.; Menendez, J.A. Autophagy Facilitates the Development of Breast Cancer Resistance to the Anti-HER2 Monoclonal Antibody Trastuzumab. PLoS ONE 2009, 4. [Google Scholar] [CrossRef] [Green Version]
  36. Rothe, K.; Porter, V.; Jiang, X.Y. Current Outlook on Autophagy in Human Leukemia: Foe in Cancer Stem Cells and Drug Resistance, Friend in New Therapeutic Interventions. Int. J. Mol. Sci. 2019, 20. [Google Scholar] [CrossRef] [Green Version]
  37. Jo, G.H.; Bogler, O.; Chwae, Y.J.; Yoo, H.; Lee, S.H.; Park, J.B.; Kim, Y.J.; Kim, J.H.; Gwak, H.S. Radiation-Induced Autophagy Contributes to Cell Death and Induces Apoptosis Partly in Malignant Glioma Cells. Cancer Res. Treat. 2015, 47, 221–241. [Google Scholar] [CrossRef]
  38. He, S.S.; Zhao, Z.; Yang, Y.F.; O’Connell, D.; Zhang, X.W.; Oh, S.; Ma, B.Y.; Lee, J.H.; Zhang, T.; Varghese, B.; et al. Truncating mutation in the autophagy gene UVRAG confers oncogenic properties and chemosensitivity in colorectal cancers. Nat. Commun. 2015, 6. [Google Scholar] [CrossRef]
  39. Yue, Z.Y.; Jin, S.K.; Yang, C.W.; Levine, A.J.; Heintz, N. Beclin 1, an autophagy gene essential for early embryonic development, is a haploinsufficient tumor suppressor. Proc. Natl. Acad. Sci. USA 2003, 100, 15077–15082. [Google Scholar] [CrossRef] [Green Version]
  40. Takamura, A.; Komatsu, M.; Hara, T.; Sakamoto, A.; Kishi, C.; Waguri, S.; Eishi, Y.; Hino, O.; Tanaka, K.; Mizushima, N. Autophagy-deficient mice develop multiple liver tumors. Genes Dev. 2011, 25, 795–800. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Yang, X.; Yu, D.D.; Yan, F.; Jing, Y.Y.; Han, Z.P.; Sun, K.; Liang, L.; Hou, J.; Wei, L.X. The role of autophagy induced by tumor microenvironment in different cells and stages of cancer. Cell Biosci. 2015, 5. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Rahman, M.A.; Bishayee, K.; Habib, K.; Sadra, A.; Huh, S.O. 18alpha-Glycyrrhetinic acid lethality for neuroblastoma cells via de-regulating the Beclin-1/Bcl-2 complex and inducing apoptosis. Biochem. Pharmacol. 2016, 117, 97–112. [Google Scholar] [CrossRef] [PubMed]
  43. Lin, C.J.; Lee, C.C.; Shih, Y.L.; Lin, T.Y.; Wang, S.H.; Lin, Y.F.; Shih, C.M. Resveratrol enhances the therapeutic effect of temozolomide against malignant glioma in vitro and in vivo by inhibiting autophagy. Free Radi.c Biol. Med. 2012, 52, 377–391. [Google Scholar] [CrossRef] [PubMed]
  44. Zanotto-Filho, A.; Braganhol, E.; Klafke, K.; Figueiro, F.; Terra, S.R.; Paludo, F.J.; Morrone, M.; Bristot, I.J.; Battastini, A.M.; Forcelini, C.M.; et al. Autophagy inhibition improves the efficacy of curcumin/temozolomide combination therapy in glioblastomas. Cancer Lett. 2015, 358, 220–231. [Google Scholar] [CrossRef] [Green Version]
  45. Rahman, M.A.; Hwang, H.; Nah, S.Y.; Rhim, H. Gintonin stimulates autophagic flux in primary cortical astrocytes. J. Ginseng Res. 2020, 44, 67–78. [Google Scholar] [CrossRef]
  46. Kim, H.; Moon, J.Y.; Ahn, K.S.; Cho, S.K. Quercetin induces mitochondrial mediated apoptosis and protective autophagy in human glioblastoma U373MG cells. Oxid. Med. Cell. Longev. 2013, 2013, 596496. [Google Scholar] [CrossRef] [Green Version]
  47. Xi, G.; Hu, X.; Wu, B.; Jiang, H.; Young, C.Y.; Pang, Y.; Yuan, H. Autophagy inhibition promotes paclitaxel-induced apoptosis in cancer cells. Cancer Lett. 2011, 307, 141–148. [Google Scholar] [CrossRef]
  48. Suzuki, R.; Kang, Y.; Li, X.; Roife, D.; Zhang, R.; Fleming, J.B. Genistein potentiates the antitumor effect of 5-Fluorouracil by inducing apoptosis and autophagy in human pancreatic cancer cells. Anticancer Res. 2014, 34, 4685–4692. [Google Scholar]
  49. Lv, X.Q.; Liu, F.; Shang, Y.; Chen, S.Z. Honokiol exhibits enhanced antitumor effects with chloroquine by inducing cell death and inhibiting autophagy in human non-small cell lung cancer cells. Oncol. Rep. 2015, 34, 1289–1300. [Google Scholar] [CrossRef] [Green Version]
  50. Oh, J.M.; Kim, E.; Chun, S. Ginsenoside Compound K Induces Ros-Mediated Apoptosis and Autophagic Inhibition in Human Neuroblastoma Cells In Vitro and In Vivo. Int. J. Mol. Sci. 2019, 20. [Google Scholar] [CrossRef] [Green Version]
  51. Rahman, M.A.; Bishayee, K.; Sadra, A.; Huh, S.O. Oxyresveratrol activates parallel apoptotic and autophagic cell death pathways in neuroblastoma cells. Biochim. Biophys. Acta Gen Subj. 2017, 1861, 23–36. [Google Scholar] [CrossRef] [PubMed]
  52. Russo, M.; Russo, G.L. Autophagy inducers in cancer. Biochem. Pharmacol. 2018, 153, 51–61. [Google Scholar] [CrossRef] [PubMed]
  53. Nunez-Olvera, S.I.; Gallardo-Rincon, D.; Puente-Rivera, J.; Salinas-Vera, Y.M.; Marchat, L.A.; Morales-Villegas, R.; Lopez-Camarillo, C. Autophagy Machinery as a Promising Therapeutic Target in Endometrial Cancer. Front. Oncol. 2019, 9. [Google Scholar] [CrossRef] [PubMed]
  54. Wu, Y.Y.; Wu, H.C.; Wu, J.E.; Huang, K.Y.; Yang, S.C.; Chen, S.X.; Tsao, C.J.; Hsu, K.F.; Chen, Y.L.; Hong, T.M. The dual PI3K/mTOR inhibitor BEZ235 restricts the growth of lung cancer tumors regardless of EGFR status, as a potent accompanist in combined therapeutic regimens. J. Exp. Clin. Cancer Res. 2019, 38. [Google Scholar] [CrossRef] [PubMed]
  55. Rizzieri, D.A.; Feldman, E.; DiPersio, J.F.; Gabrail, N.; Stock, W.; Strair, R.; Rivera, V.M.; Albitar, M.; Bedrosian, C.L.; Giles, F.J. A phase 2 clinical trial of deforolimus (AP23573, MK-8669), a novel mammalian target of rapamycin inhibitor, in patients with relapsed or refractory hematologic malignancies. Clin. Cancer Res. 2008, 14, 2756–2762. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Li, X.; Lu, Y.; Pan, T.; Fan, Z. Roles of autophagy in cetuximab-mediated cancer therapy against EGFR. Autophagy 2010, 6, 1066–1077. [Google Scholar] [CrossRef] [Green Version]
  57. Lauzier, A.; Normandeau-Guimond, J.; Vaillancourt-Lavigueur, V.; Boivin, V.; Charbonneau, M.; Rivard, N.; Scott, M.S.; Dubois, C.M.; Jean, S. Colorectal cancer cells respond differentially to autophagy inhibition in vivo. Sci. Rep. 2019, 9, 11316. [Google Scholar] [CrossRef] [Green Version]
  58. Strohecker, A.M.; White, E. Autophagy promotes BrafV600E-driven lung tumorigenesis by preserving mitochondrial metabolism. Autophagy 2014, 10, 384–385. [Google Scholar] [CrossRef] [Green Version]
  59. Avalos, Y.; Canales, J.; Bravo-Sagua, R.; Criollo, A.; Lavandero, S.; Quest, A.F. Tumor suppression and promotion by autophagy. BioMed Res. Int. 2014, 2014, 603980. [Google Scholar] [CrossRef]
  60. Shteingauz, A.; Porat, Y.; Voloshin, T.; Schneiderman, R.S.; Munster, M.; Zeevi, E.; Kaynan, N.; Gotlib, K.; Giladi, M.; Kirson, E.D.; et al. AMPK-dependent autophagy upregulation serves as a survival mechanism in response to Tumor Treating Fields (TTFields). Cell. Death Dis. 2018, 9. [Google Scholar] [CrossRef] [Green Version]
  61. Rahman, M.A.; Rahman, M.S.; Uddin, M.J.; Mamum-Or-Rashid, A.N.M.; Pang, M.G.; Rhim, H. Emerging risk of environmental factors: Insight mechanisms of Alzheimer’s diseases. Environ. Sci. Pollut. Res. 2020. [Google Scholar] [CrossRef] [PubMed]
  62. Chen, H.H.W.; Kuo, M.T. Improving radiotherapy in cancer treatment: Promises and challenges. Oncotarget 2017, 8, 62742–62758. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Luo, H.; Vong, C.T.; Chen, H.B.; Gao, Y.; Lyu, P.; Qiu, L.; Zhao, M.M.; Liu, Q.; Cheng, Z.H.; Zou, J.; et al. Naturally occurring anti-cancer compounds: Shining from Chinese herbal medicine. Chin. Med. 2019, 14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Wang, P.Q.; Zhu, L.J.; Sun, D.J.; Gan, F.H.; Gao, S.Y.; Yin, Y.Y.; Chen, L.X. Natural products as modulator of autophagy with potential clinical prospects. Apoptosis 2017, 22, 325–356. [Google Scholar] [CrossRef]
  65. Choi, E.J.; Kim, G.H. Apigenin Induces Apoptosis through a Mitochondria/Caspase-Pathway in Human Breast Cancer MDA-MB-453 Cells. J. Clin. Biochem. Nutr. 2009, 44, 260–265. [Google Scholar] [CrossRef] [Green Version]
  66. Mohan, N.; Banik, N.L.; Ray, S.K. Combination of N-(4-hydroxyphenyl) retinamide and apigenin suppressed starvation-induced autophagy and promoted apoptosis in malignant neuroblastoma cells. Neurosci. Lett. 2011, 502, 24–29. [Google Scholar] [CrossRef] [Green Version]
  67. Nakamura, Y.; Yogosawa, S.; Izutani, Y.; Watanabe, H.; Otsuji, E.; Sakai, T. A combination of indol-3-carbinol and genistein synergistically induces apoptosis in human colon cancer HT-29 cells by inhibiting Akt phosphorylation and progression of autophagy. Mol. Cancer 2009, 8, 100. [Google Scholar] [CrossRef] [Green Version]
  68. Gan, M.L.; Zheng, T.; Shen, L.Y.; Tan, Y.; Fan, Y.; Shuai, S.R.; Bai, L.; Li, X.W.; Wang, J.Y.; Zhang, S.H.; et al. Genistein reverses isoproterenol-induced cardiac hypertrophy by regulating miR-451/TIMP2. Biomed. Pharmacother. 2019, 112. [Google Scholar] [CrossRef]
  69. Ma, C.H.; Zhang, Y.X.; Tang, L.H.; Yang, X.J.; Cui, W.M.; Han, C.C.; Ji, W.Y. MicroRNA-1469, a p53-responsive microRNA promotes Genistein induced apoptosis by targeting Mcl1 in human laryngeal cancer cells. Biomed. Pharmacother. 2018, 106, 665–671. [Google Scholar] [CrossRef]
  70. Kuo, P.L.; Hsu, Y.L.; Cho, C.Y. Plumbagin induces G2-M arrest and autophagy by inhibiting the AKT/mammalian target of rapamycin pathway in breast cancer cells. Mol. Cancer Ther. 2006, 5, 3209–3221. [Google Scholar] [CrossRef] [Green Version]
  71. Thiyagarajan, V.; Tsai, M.J.; Weng, C.F. Antroquinonol Targets FAK-Signaling Pathway Suppressed Cell Migration, Invasion, and Tumor Growth of C6 Glioma. PLoS ONE 2015, 10, e0141285. [Google Scholar] [CrossRef] [PubMed]
  72. Yu, C.C.; Chiang, P.C.; Lu, P.H.; Kuo, M.T.; Wen, W.C.; Chen, P.N.; Guh, J.H. Antroquinonol, a natural ubiquinone derivative, induces a cross talk between apoptosis, autophagy and senescence in human pancreatic carcinoma cells. J. Nutr. Biochem. 2012, 23, 900–907. [Google Scholar] [CrossRef] [PubMed]
  73. Chiang, P.C.; Lin, S.C.; Pan, S.L.; Kuo, C.H.; Tsai, I.L.; Kuo, M.T.; Wen, W.C.; Chen, P.; Guh, J.H. Antroquinonol displays anticancer potential against human hepatocellular carcinoma cells: A crucial role of AMPK and mTOR pathways. Biochem. Pharmacol. 2010, 79, 162–171. [Google Scholar] [CrossRef] [PubMed]
  74. Chiou, J.F.; Wu, A.T.H.; Wang, W.T.; Kuo, T.H.; Gelovani, J.G.; Lin, I.H.; Wu, C.H.; Chiu, W.T.; Deng, W.P. A Preclinical Evaluation of Antrodia camphorata Alcohol Extracts in the Treatment of Non-Small Cell Lung Cancer Using Non-Invasive Molecular Imaging. Evid.-Based Complement. Altern. Med. 2011, 1–12. [Google Scholar] [CrossRef]
  75. Kannaiyan, R.; Manu, K.A.; Chen, L.X.; Li, F.; Rajendran, P.; Subramaniam, A.; Lam, P.; Kumar, A.P.; Sethi, G. Celastrol inhibits tumor cell proliferation and promotes apoptosis through the activation of c-Jun N-terminal kinase and suppression of PI3 K/Akt signaling pathways. Apoptosis 2011, 16, 1028–1041. [Google Scholar] [CrossRef]
  76. Wang, W.B.; Feng, L.X.; Yue, Q.X.; Wu, W.Y.; Guan, S.H.; Jiang, B.H.; Yang, M.; Liu, X.; Guo, D.A. Paraptosis accompanied by autophagy and apoptosis was induced by celastrol, a natural compound with influence on proteasome, ER stress and Hsp90. J. Cell. Physiol. 2012, 227, 2196–2206. [Google Scholar] [CrossRef]
  77. Li, H.; Zhang, Y.Y.; Tan, H.W.; Jia, Y.F.; Li, D. Therapeutic effect of tripterine on adjuvant arthritis in rats. J. Ethnopharmacol. 2008, 118, 479–484. [Google Scholar] [CrossRef] [Green Version]
  78. Dyshlovoy, S.A.; Hauschild, J.; Amann, K.; Tabakmakher, K.M.; Venz, S.; Walther, R.; Guzii, A.G.; Makarieva, T.N.; Shubina, L.K.; Fedorov, S.N.; et al. Marine alkaloid Monanchocidin a overcomes drug resistance by induction of autophagy and lysosomal membrane permeabilization. Oncotarget 2015, 6, 17328–17341. [Google Scholar] [CrossRef]
  79. Dyshlovoy, S.A.; Venz, S.; Hauschild, J.; Tabakmakher, K.M.; Otte, K.; Madanchi, R.; Walther, R.; Guzii, A.G.; Makarieva, T.N.; Shubina, L.K.; et al. Anti-migratory activity of marine alkaloid monanchocidin A—Proteomics-based discovery and confirmation. Proteomics 2016, 16, 1590–1603. [Google Scholar] [CrossRef]
  80. McGuire, W.P.; Blessing, J.A.; Moore, D.; Lentz, S.S.; Photopulos, G. Paclitaxel has moderate activity in squamous cervix cancer: A gynecologic oncology group study. J. Clin. Oncol. 1996, 14, 792–795. [Google Scholar] [CrossRef]
  81. Kampan, N.C.; Madondo, M.T.; McNally, O.M.; Quinn, M.; Plebanski, M. Paclitaxel and Its Evolving Role in the Management of Ovarian Cancer. Biomed. Res. Int. 2015, 2015, 413076. [Google Scholar] [CrossRef] [PubMed]
  82. Qiu, Y.; Li, P.; Ji, C.Y. Cell Death Conversion under Hypoxic Condition in Tumor Development and Therapy. Int. J. Mol. Sci. 2015, 16, 25536–25551. [Google Scholar] [CrossRef] [PubMed]
  83. Lee, Y.; Na, J.; Lee, M.S.; Cha, E.Y.; Sul, J.Y.; Park, J.B.; Lee, J.S. Combination of pristimerin and paclitaxel additively induces autophagy in human breast cancer cells via ERK1/2 regulation. Mol. Med. Rep. 2018, 18, 4281–4288. [Google Scholar] [CrossRef] [PubMed]
  84. Veldhoen, R.A.; Banman, S.L.; Hemmerling, D.R.; Odsen, R.; Simmen, T.; Simmonds, A.J.; Underhill, D.A.; Goping, I.S. The chemotherapeutic agent paclitaxel inhibits autophagy through two distinct mechanisms that regulate apoptosis. Oncogene 2013, 32, 736–746. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Liu, J.; Hu, X.J.; Jin, B.; Qu, X.J.; Hou, K.Z.; Liu, Y.P. ss-Elemene induces apoptosis as well as protective autophagy in human non-small-cell lung cancer A549 cells. J. Pharm. Pharmacol. 2012, 64, 146–153. [Google Scholar] [CrossRef]
  86. Zhan, Y.H.; Liu, J.; Qu, X.J.; Hou, K.Z.; Wang, K.F.; Liu, Y.P.; Wu, B. beta-Elemene Induces Apoptosis in Human Renal-cell Carcinoma 786-0 Cells through Inhibition of MAPK/ERK and PI3K/Akt/mTOR Signalling Pathways. Asian Pac. J. Cancer P 2012, 13, 2739–2744. [Google Scholar] [CrossRef] [Green Version]
  87. Saiki, S.; Sasazawa, Y.; Imamichi, Y.; Kawajiri, S.; Fujimaki, T.; Tanida, I.; Kobayashi, H.; Sato, F.; Sato, S.; Ishikawa, K.I.; et al. Caffeine induces apoptosis by enhancement of autophagy via PI3K/Akt/mTOR/p70S6K inhibition. Autophagy 2011, 7, 176–187. [Google Scholar] [CrossRef] [Green Version]
  88. Lopez-Lazaro, M. Anticancer and carcinogenic properties of curcumin: Considerations for its clinical development as a cancer chemopreventive and chemotherapeutic agent. Mol. Nutr. Food Res. 2008, 52, S103–S127. [Google Scholar] [CrossRef]
  89. Aoki, H.; Takada, Y.; Kondo, S.; Sawaya, R.; Aggarwal, B.B.; Kondo, Y. Evidence that curcumin suppresses the growth of malignant gliomas in vitro and in vivo through induction of autophagy: Role of Akt and extracellular signal-regulated kinase signaling pathways. Mol. Pharmacol. 2007, 72, 29–39. [Google Scholar] [CrossRef] [Green Version]
  90. Klinger, N.V.; Mittal, S. Therapeutic Potential of Curcumin for the Treatment of Brain Tumors. Oxid. Med. Cell. Longev. 2016, 2016, 9324085. [Google Scholar] [CrossRef]
  91. Masuelli, L.; Benvenuto, M.; Di Stefano, E.; Mattera, R.; Fantini, M.; De Feudis, G.; De Smaele, E.; Tresoldi, I.; Giganti, M.G.; Modesti, A.; et al. Curcumin blocks autophagy and activates apoptosis of malignant mesothelioma cell lines and increases the survival of mice intraperitoneally transplanted with a malignant mesothelioma cell line. Oncotarget 2017, 8, 34405–34422. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Buhrmann, C.; Kraehe, P.; Lueders, C.; Shayan, P.; Goel, A.; Shakibaei, M. Curcumin Suppresses Crosstalk between Colon Cancer Stem Cells and Stromal Fibroblasts in the Tumor Microenvironment: Potential Role of EMT. PLoS ONE 2014, 9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Xu, R.; Li, H.B.; Wu, S.Q.; Qu, J.; Yuan, H.Y.; Zhou, Y.G.; Lu, Q. MicroRNA-1246 regulates the radio-sensitizing effect of curcumin in bladder cancer cells via activating P53. Int. Urol. Nephrol. 2019, 51, 1771–1779. [Google Scholar] [CrossRef] [PubMed]
  94. Aggarwal, B.B.; Deb, L.; Prasad, S. Curcumin Differs from Tetrahydrocurcumin for Molecular Targets, Signaling Pathways and Cellular Responses. Molecules 2015, 20, 185–205. [Google Scholar] [CrossRef] [Green Version]
  95. Wu, J.C.; Lai, C.S.; Badmaev, V.; Nagabhushanam, K.; Ho, C.T.; Pan, M.H. Tetrahydrocurcumin, a major metabolite of curcumin, induced autophagic cell death through coordinative modulation of PI3K/Akt-mTOR and MAPK signaling pathways in human leukemia HL-60 cells. Mol. Nutr. Food Res. 2011, 55, 1646–1654. [Google Scholar] [CrossRef] [PubMed]
  96. Hu, F.; Wei, F.; Wang, Y.L.; Wu, B.B.; Fang, Y.; Xiong, B. EGCG synergizes the therapeutic effect of cisplatin and oxaliplatin through autophagic pathway in human colorectal cancer cells. J. Pharmacol. Sci. 2015, 128, 27–34. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Irimie, A.I.; Braicu, C.; Zanoaga, O.; Pileczki, V.; Gherman, C.; Berindan-Neagoe, I.; Campian, R.S. Epigallocatechin-3-gallate suppresses cell proliferation and promotes apoptosis and autophagy in oral cancer SSC-4 cells. Oncotargets Ther. 2015, 8, 461–470. [Google Scholar] [CrossRef] [Green Version]
  98. Zhou, J.; Farah, B.L.; Sinha, R.A.; Wu, Y.; Singh, B.K. Epigallocatechin-3-Gallate (EGCG), a Green Tea Polyphenol, Stimulates Hepatic Autophagy and Lipid Clearance (vol 9, e87161, 2014). PLoS ONE 2014, 9. [Google Scholar] [CrossRef]
  99. Gray, A.L.; Stephens, C.A.; Bigelow, R.L.H.; Coleman, D.T.; Cardelli, J.A. The Polyphenols (-)-Epigallocatechin-3-Gallate and Luteolin Synergistically Inhibit TGF-beta-Induced Myofibroblast Phenotypes through RhoA and ERK Inhibition. PLoS ONE 2014, 9, e109208. [Google Scholar] [CrossRef] [Green Version]
  100. Chen, A.P.; Zhang, L. The antioxidant (-)-epigallocatechin-3-gallate inhibits rat hepatic stellate cell proliferation in vitro by blocking the tyrosine phosphorylation and reducing the gene expression of platelet-derived growth factor-beta receptor. J. Biol. Chem. 2003, 278, 23381–23389. [Google Scholar] [CrossRef] [Green Version]
  101. Chen, L.L.; Xu, Y. Epigallocatechin gallate attenuates uric acid-induced injury in rat renal interstitial fibroblasts NRK-49F by up-regulation of miR-9. Eur. Rev. Med. Pharmacol. 2018, 22, 7458–7469. [Google Scholar]
  102. Adhami, V.M.; Syed, D.N.; Khan, N.; Mukhtar, H. Dietary flavonoid fisetin: A novel dual inhibitor of PI3K/Akt and mTOR for prostate cancer management. Biochem. Pharmacol. 2012, 84, 1277–1281. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Sun, X.; Ma, X.M.; Li, Q.W.; Yang, Y.; Xu, X.L.; Sun, J.Q.; Yu, M.W.; Cao, K.X.; Yang, L.; Yang, G.W.; et al. Anti-cancer effects of fisetin on mammary carcinoma cells via regulation of the PI3K/Akt/mTOR pathway: In vitro and in vivo studies. Int. J. Mol. Med. 2018, 42, 811–820. [Google Scholar] [CrossRef] [PubMed]
  104. Khan, N.; Afaq, F.; Khusro, F.H.; Adhami, V.M.; Suh, Y.; Mukhtar, H. Dual inhibition of phosphatidylinositol 3-kinase/Akt and mammalian target of rapamycin signaling in human nonsmall cell lung cancer cells by a dietary flavonoid fisetin. Int. J. Cancer 2012, 130, 1695–1705. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Suh, Y.; Afaq, F.; Khan, N.; Johnson, J.J.; Khusro, F.H.; Mukhtar, H. Fisetin induces autophagic cell death through suppression of mTOR signaling pathway in prostate cancer cells. Carcinogenesis 2010, 31, 1424–1433. [Google Scholar] [CrossRef] [Green Version]
  106. Yang, P.M.; Tseng, H.H.; Peng, C.W.; Chen, W.S.; Chiu, S.J. Dietary flavonoid fisetin targets caspase-3-deficient human breast cancer MCF-7 cells by induction of caspase-7-associated apoptosis and inhibition of autophagy. Int. J. Oncol. 2012, 40, 469–478. [Google Scholar] [CrossRef] [Green Version]
  107. Tiwari, R.V.; Parajuli, P.; Sylvester, P.W. gamma-Tocotrienol-induced endoplasmic reticulum stress and autophagy act concurrently to promote breast cancer cell death. Biochem. Cell. Biol. 2015, 93, 306–320. [Google Scholar] [CrossRef]
  108. Deng, S.; Shanmugam, M.K.; Kumar, A.P.; Yap, C.T.; Sethi, G.; Bishayee, A. Targeting autophagy using natural compounds for cancer prevention and therapy. Cancer-Am. Cancer Soc. 2019, 125, 1228–1246. [Google Scholar] [CrossRef]
  109. Jiang, Q.; Rao, X.Y.; Kim, C.Y.; Freiser, H.; Zhang, Q.B.; Jiang, Z.Y.; Li, G.L. Gamma-tocotrienol induces apoptosis and autophagy in prostate cancer cells by increasing intracellular dihydrosphingosine and dihydroceramide. Int. J. Cancer 2012, 130, 685–693. [Google Scholar] [CrossRef] [Green Version]
  110. Tiwari, R.V.; Parajuli, P.; Sylvester, P.W. gamma-Tocotrienol-induced autophagy in malignant mammary cancer cells. Exp. Biol. Med. 2014, 239, 33–44. [Google Scholar] [CrossRef]
  111. Kim, S.H.; Park, E.J.; Lee, C.R.; Chun, J.N.; Cho, N.H.; Kim, I.G.; Lee, S.; Kim, T.W.; Park, H.H.; So, I.; et al. Geraniol induces cooperative interaction of apoptosis and autophagy to elicit cell death in PC-3 prostate cancer cells. Int. J. Oncol. 2012, 40, 1683–1690. [Google Scholar] [CrossRef] [PubMed]
  112. Trzoss, L.; Fukuda, T.; Costa-Lotufo, L.V.; Jimenez, P.; La Clair, J.J.; Fenical, W. Seriniquinone, a selective anticancer agent, induces cell death by autophagocytosis, targeting the cancer-protective protein dermcidin. Proc. Natl. Acad. Sci. USA 2014, 111, 14687–14692. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Chu, S.C.; Hsieh, Y.S.; Yu, C.C.; Lai, Y.Y.; Chen, P.N. Thymoquinone Induces Cell Death in Human Squamous Carcinoma Cells via Caspase Activation-Dependent Apoptosis and LC3-II Activation-Dependent Autophagy. PLoS ONE 2014, 9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Racoma, I.O.; Meisen, W.H.; Wang, Q.E.; Kaur, B.; Wani, A.A. Thymoquinone Inhibits Autophagy and Induces Cathepsin-Mediated, Caspase-Independent Cell Death in Glioblastoma Cells. PLoS ONE 2013, 8. [Google Scholar] [CrossRef] [PubMed]
  115. Chen, M.C.; Lee, N.H.; Hsu, H.H.; Ho, T.J.; Tu, C.C.; Hsieh, D.J.Y.; Lin, Y.M.; Chen, L.M.; Kuo, W.W.; Huang, C.Y. Thymoquinone Induces Caspase-Independent, Autophagic Cell Death in CPT-11-Resistant LoVo Colon Cancer via Mitochondrial Dysfunction and Activation of JNK and p38. J. Agric. Food Chem. 2015, 63, 1540–1546. [Google Scholar] [CrossRef] [PubMed]
  116. Xu, H.L.; Tang, W.; Du, G.H.; Kokudo, N. Targeting apoptosis pathways in cancer with magnolol and honokiol, bioactive constituents of the bark of Magnolia officinalis. Drug Discov. Ther. 2011, 5, 202–210. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Li, H.B.; Yi, X.; Gao, J.M.; Ying, X.X.; Guan, H.Q.; Li, J.C. Magnolol-induced H460 cells death via autophagy but not apoptosis. Arch. Pharm. Res. 2007, 30, 1566–1574. [Google Scholar] [CrossRef]
  118. Rasul, A.; Yu, B.; Khan, M.; Zhang, K.; Iqbal, F.; Ma, T.H.; Yang, H. Magnolol, a natural compound, induces apoptosis of SGC-7901 human gastric adenocarcinoma cells via the mitochondrial and PI3K/Akt signaling pathways. Int. J. Oncol. 2012, 40, 1153–1161. [Google Scholar] [CrossRef] [Green Version]
  119. Bae, U.J.; Jung, E.S.; Jung, S.J.; Chae, S.W.; Park, B.H. Mulberry leaf extract displays antidiabetic activity in db/db mice via Akt and AMP-activated protein kinase phosphorylation. Food Nutr. Res. 2018, 62. [Google Scholar] [CrossRef] [Green Version]
  120. Cheng, K.C.; Wang, C.J.; Chang, Y.C.; Hung, T.W.; Lai, C.J.; Kuo, C.W.; Huang, H.P. Mulberry fruits extracts induce apoptosis and autophagy of liver cancer cell and prevent hepatocarcinogenesis in vivo. J. Food Drug Anal. 2020, 28, 84–93. [Google Scholar] [CrossRef] [Green Version]
  121. Kwon, Y.H.; Bishayee, K.; Rahman, M.A.; Hong, J.S.; Lim, S.S.; Huh, S.O. Morus alba Accumulates Reactive Oxygen Species to Initiate Apoptosis via FOXO-Caspase 3-Dependent Pathway in Neuroblastoma Cells. Mol. Cells 2015, 38, 630–637. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Lao, Y.Z.; Wan, G.; Liu, Z.Y.; Wang, X.Y.; Ruan, P.; Xu, W.; Xu, D.Q.; Xie, W.D.; Zhang, Y.; Xu, H.X.; et al. The natural compound oblongifolin C inhibits autophagic flux and enhances antitumor efficacy of nutrient deprivation. Autophagy 2014, 10, 736–749. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Wu, M.; Lao, Y.Z.; Tan, H.S.; Lu, G.; Ren, Y.; Zheng, Z.Q.; Yi, J.; Fu, W.W.; Shen, H.M.; Xu, H.X. Oblongifolin C suppresses lysosomal function independently of TFEB nuclear translocation. Acta Pharmacol. Sin. 2019, 40, 929–937. [Google Scholar] [CrossRef] [PubMed]
  124. Acharya, B.R.; Bhattacharyya, S.; Choudhury, D.; Chakrabarti, G. The microtubule depolymerizing agent naphthazarin induces both apoptosis and autophagy in A549 lung cancer cells. Apoptosis 2011, 16, 924–939. [Google Scholar] [CrossRef]
  125. Herr, I.; Lozanovski, V.; Houben, P.; Schemmer, P.; Buchler, M.W. Sulforaphane and related mustard oils in focus of cancer prevention and therapy. Wien. Med. Wochenschr. 2013, 163, 80–88. [Google Scholar] [CrossRef]
  126. Herman-Antosiewicz, A.; Johnson, D.E.; Singh, S.V. Sulforaphane causes autophagy to inhibit release of cytochrome C and apoptosis in human prostate cancer cells. Cancer Res. 2006, 66, 5828–5835. [Google Scholar] [CrossRef] [Green Version]
  127. Zheng, Z.; Lin, K.; Hu, Y.; Zhou, Y.; Ding, X.; Wang, Y.; Wu, W. Sulforaphane metabolites inhibit migration and invasion via microtubule-mediated Claudins dysfunction or inhibition of autolysosome formation in human non-small cell lung cancer cells. Cell Death Dis. 2019, 10, 259. [Google Scholar] [CrossRef] [Green Version]
  128. Chaudhuri, D.; Orsulic, S.; Ashok, B.T. Antiproliferative activity of sulforaphane in Akt-overexpressing ovarian cancer cells. Mol. Cancer Ther. 2007, 6, 334–345. [Google Scholar] [CrossRef] [Green Version]
  129. Uddin, M.S.; Mamun, A.A.; Jakaria, M.; Thangapandiyan, S.; Ahmad, J.; Rahman, M.A.; Mathew, B.; Abdel-Daim, M.M.; Aleya, L. Emerging promise of sulforaphane-mediated Nrf2 signaling cascade against neurological disorders. Sci. Total Environ. 2020, 707, 135624. [Google Scholar] [CrossRef]
  130. Kim, S.M.; Park, H.S.; Jun, D.Y.; Woo, H.J.; Woo, M.H.; Yang, C.H.; Kim, Y.H. Mollugin induces apoptosis in human Jurkat T cells through endoplasmic reticulum stress-mediated activation of JNK and caspase-12 and subsequent activation of mitochondria-dependent caspase cascade regulated by Bcl-xL. Toxicol. Appl. Pharm. 2009, 241, 210–220. [Google Scholar] [CrossRef]
  131. Zhang, L.; Wang, H.D.; Zhu, J.H.; Xu, J.G.; Ding, K. Mollugin induces tumor cell apoptosis and autophagy via the PI3K/AKT/mTOR/p70S6K and ERK signaling pathways. Biochem. Biophys Res. Commun. 2014, 450, 247–254. [Google Scholar] [CrossRef] [PubMed]
  132. Xu, M.Y.; Lee, S.Y.; Kang, S.S.; Kim, Y.S. Antitumor Activity of Jujuboside B and the Underlying Mechanism via Induction of Apoptosis and Autophagy. J. Nat. Prod. 2014, 77, 370–376. [Google Scholar] [CrossRef] [PubMed]
  133. Lu, J.; Sun, D.P.; Gao, S.; Gao, Y.; Ye, J.T.; Liu, P.Q. Cyclovirobuxine D Induces Autophagy-Associated Cell Death via the Akt/mTOR Pathway in MCF-7 Human Breast Cancer Cells. J. Pharmacol. Sci. 2014, 125, 74–82. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Ouyang, L.; Chen, Y.; Wang, X.Y.; Lu, R.F.; Zhang, S.Y.; Tian, M.; Xie, T.; Liu, B.; He, G. Polygonatum odoratum lectin induces apoptosis and autophagy via targeting EGFR-mediated Ras-Raf-MEK-ERK pathway in human MCF-7 breast cancer cells. Phytomedicine 2014, 21, 1658–1665. [Google Scholar] [CrossRef]
  135. Li, C.Y.; Chen, J.; Lu, B.M.; Shi, Z.; Wang, H.L.; Zhang, B.; Zhao, K.L.; Qi, W.; Bao, J.K.; Wang, Y. Molecular Switch Role of Akt in Polygonatum odoratum Lectin-Induced Apoptosis and Autophagy in Human Non-Small Cell Lung Cancer A549 Cells. PLoS ONE 2014, 9, e101526. [Google Scholar] [CrossRef]
  136. Jiang, H.; Zhang, L.; Kuo, J.; Kuo, K.; Gautam, S.C.; Groc, L.; Rodriguez, A.I.; Koubi, D.; Hunter, T.J.; Corcoran, G.B.; et al. Resveratrol-induced apoptotic death in human U251 glioma cells. Mol. Cancer Ther. 2005, 4, 554–561. [Google Scholar] [CrossRef] [Green Version]
  137. Rahman, M.A.; Kim, N.H.; Kim, S.H.; Oh, S.M.; Huh, S.O. Antiproliferative and Cytotoxic Effects of Resveratrol in Mitochondria-Mediated Apoptosis in Rat B103 Neuroblastoma Cells. Korean J. Physiol. Pharm. 2012, 16, 321–326. [Google Scholar] [CrossRef]
  138. Lang, F.F.; Qin, Z.Y.; Li, F.; Zhang, H.L.; Fang, Z.H.; Hao, E.K. Apoptotic Cell Death Induced by Resveratrol Is Partially Mediated by the Autophagy Pathway in Human Ovarian Cancer Cells. PLoS ONE 2015, 10. [Google Scholar] [CrossRef]
  139. Garg, T.; Yadav, V.K. Effects of Resveratrol as an anticancer agent. A Systematic Review and Meta-Analysis. Indian J. Pharmacol. 2013, 45, S206. [Google Scholar]
  140. Gong, C.H.; Xia, H.L. Resveratrol suppresses melanoma growth by promoting autophagy through inhibiting the PI3K/AKT/mTOR signaling pathway. Exp. Ther. Med. 2020, 19, 1878–1886. [Google Scholar] [CrossRef] [Green Version]
  141. Liu, Q.J.; Fang, Q.; Ji, S.Q.; Han, Z.X.; Cheng, W.L.; Zhang, H.J. Resveratrol-mediated apoptosis in renal cell carcinoma via the p53/AMP-activated protein kinase/mammalian target of rapamycin autophagy signaling pathway. Mol. Med. Rep. 2018, 17, 502–508. [Google Scholar] [CrossRef] [PubMed]
  142. Godichaud, S.; Si-Tayeb, K.; Auge, N.; Desmouliere, A.; Balabaud, C.; Payrastre, B.; Negre-Salvayre, A.; Rosenbaum, J. The grape-derived polyphenol resveratrol differentially affects epidermal and platelet-derived growth factor signaling in human liver myofibroblasts. Int. J. Biochem. Cell Biol. 2006, 38, 629–637. [Google Scholar] [CrossRef] [PubMed]
  143. Jiang, Y.D.; Liu, L.; Steinle, J.J. miRNA15a regulates insulin signal transduction in the retinal vasculature. Cell. Signal. 2018, 44, 28–32. [Google Scholar] [CrossRef]
  144. Anand David, A.V.; Arulmoli, R.; Parasuraman, S. Overviews of Biological Importance of Quercetin: A Bioactive Flavonoid. Pharmacogn. Rev. 2016, 10, 84–89. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Singh, B.N.; Kumar, D.; Shankar, S.; Srivastava, R.K. Rottlerin induces autophagy which leads to apoptotic cell death through inhibition of PI3K/Akt/mTOR pathway in human pancreatic cancer stem cells. Biochem. Pharmacol. 2012, 84, 1154–1163. [Google Scholar] [CrossRef]
  146. Rahman, M.A.; Kim, N.H.; Yang, H.; Huh, S.O. Angelicin induces apoptosis through intrinsic caspase-dependent pathway in human SH-SY5Y neuroblastoma cells. Mol. Cell. Biochem. 2012, 369, 95–104. [Google Scholar] [CrossRef]
  147. Rahman, M.A.; Bishayee, K.; Huh, S.O. Angelica polymorpha Maxim Induces Apoptosis of Human SH-SY5Y Neuroblastoma Cells by Regulating an Intrinsic Caspase Pathway. Mol. Cells 2016, 39, 119–128. [Google Scholar] [CrossRef] [Green Version]
  148. Wang, Y.R.; Chen, Y.Q.; Chen, X.D.; Liang, Y.; Yang, D.P.; Dong, J.; Yang, N.; Liang, Z.Q. Angelicin inhibits the malignant behaviours of human cervical cancer potentially via inhibiting autophagy. Exp. Ther. Med. 2019, 18, 3365–3374. [Google Scholar] [CrossRef] [Green Version]
  149. Uddin, M.S.; Rahman, M.A.; Kabir, M.T.; Behl, T.; Mathew, B.; Perveen, A.; Barreto, G.E.; Bin-Jumah, M.N.; Abdel-Daim, M.M.; Ashraf, G.M. Multifarious roles of mTOR signaling in cognitive aging and cerebrovascular dysfunction of Alzheimer’s disease. IUBMB Life 2020. [Google Scholar] [CrossRef]
  150. Leng, S.; Hao, Y.; Du, D.; Xie, S.; Hong, L.; Gu, H.; Zhu, X.; Zhang, J.; Fan, D.; Kung, H.F. Ursolic acid promotes cancer cell death by inducing Atg5-dependent autophagy. Int. J. Cancer 2013, 133, 2781–2790. [Google Scholar] [CrossRef]
  151. Xavier, C.P.R.; Lima, C.F.; Pedro, D.F.N.; Wilson, J.M.; Kristiansen, K.; Pereira-Wilson, C. Ursolic acid induces cell death and modulates autophagy through JNK pathway in apoptosis-resistant colorectal cancer cells. J. Nutr. Biochem. 2013, 24, 706–712. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Zhao, C.; Yin, S.T.; Dong, Y.H.; Guo, X.; Fan, L.H.; Ye, M.; Hu, H.B. Autophagy-dependent EIF2AK3 activation compromises ursolic acid-induced apoptosis through upregulation of MCL1 in MCF-7 human breast cancer cells. Autophagy 2013, 9, 196–207. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Shin, S.W.; Park, J.W. Autophagy inhibition enhances ursolic acid-induced apoptosis in PC3 cells. Free Radic. Biol. Med. 2012, 53, S116. [Google Scholar] [CrossRef]
  154. Wang, S.Y.; Yu, Q.J.; Bao, J.K.; Liu, B. Polygonatum cyrtonema lectin, a potential antineoplastic drug targeting programmed cell death pathways. Biochem. Biophys Res. Commun. 2011, 406, 497–500. [Google Scholar] [CrossRef]
  155. Liu, B.; Cheng, Y.; Bian, H.J.; Bao, J.K. Molecular mechanisms of Polygonatum cyrtonema lectin-induced apoptosis and autophagy in cancer cells. Autophagy 2009, 5, 253–255. [Google Scholar] [CrossRef] [Green Version]
  156. Liu, B.; Cheng, Y.; Zhang, B.; Bian, H.J.; Bao, J.K. Polygonatum cyrtonema lectin induces apoptosis and autophagy in human melanoma A375 cells through a mitochondria-mediated ROS-p38-p53 pathway. Cancer Lett. 2009, 275, 54–60. [Google Scholar] [CrossRef]
  157. Duan, W.J.; Li, Q.S.; Xia, M.Y.; Tashiro, S.I.; Onodera, S.; Ikejima, T. Silibinin Activated p53 and Induced Autophagic Death in Human Fibrosarcoma HT1080 Cells via Reactive Oxygen Species-p38 and c-Jun N-Terminal Kinase Pathways. Biol. Pharm. Bull. 2011, 34, 47–53. [Google Scholar] [CrossRef] [Green Version]
  158. Rahman, M.A.; Yang, H.; Kim, N.H.; Huh, S.O. Induction of apoptosis by Dioscorea nipponica Makino extracts in human SH-SY5Y neuroblastoma cells via mitochondria-mediated pathway. Anim. Cells Syst. 2014, 18, 41–51. [Google Scholar] [CrossRef] [Green Version]
  159. Rahman, M.A.; Yang, H.; Lim, S.S.; Huh, S.O. Apoptotic Effects of Melandryum firmum Root Extracts in Human SH-SY5Y Neuroblastoma Cells. Exp. Neurobiol. 2013, 22, 208–213. [Google Scholar] [CrossRef] [Green Version]
  160. Bassham, D.C.; Crespo, J.L. Autophagy in plants and algae. Front. Plant Sci. 2014, 5. [Google Scholar] [CrossRef] [Green Version]
  161. Uddin, M.S.; Mamun, A.A.; Rahman, M.A.; Kabir, M.T.; Alkahtani, S.; Alanazi, I.S.; Perveen, A.; Ashraf, G.M.; Bin-Jumah, M.N.; Abdel-Daim, M.M. Exploring the Promise of Flavonoids to Combat Neuropathic Pain: From Molecular Mechanisms to Therapeutic Implications. Front. Neurosci. 2020, 14, 478. [Google Scholar] [CrossRef]
  162. Rahman, M.A.; Hong, J.S.; Huh, S.O. Antiproliferative properties of Saussurea lappa Clarke root extract in SH-SY5Y neuroblastoma cells via intrinsic apoptotic pathway. Anim. Cells Syst. 2015, 19, 119–126. [Google Scholar] [CrossRef] [Green Version]
  163. Tian, X.; Song, H.S.; Cho, Y.M.; Park, B.; Song, Y.J.; Jang, S.; Kang, S.C. Anticancer effect of Saussurea lappa extract via dual control of apoptosis and autophagy in prostate cancer cells. Medicine 2017, 96. [Google Scholar] [CrossRef]
  164. Ryoo, N.; Rahman, M.A.; Hwang, H.; Ko, S.K.; Nah, S.Y.; Kim, H.C.; Rhim, H. Ginsenoside Rk1 is a novel inhibitor of NMDA receptors in cultured rat hippocampal neurons. J. Ginseng Res. 2020, 44, 490–495. [Google Scholar] [CrossRef]
  165. Ko, H.; Kim, Y.J.; Park, J.S.; Park, J.H.; Yang, H.O. Autophagy Inhibition Enhances Apoptosis Induced by Ginsenoside Rk1 in Hepatocellular Carcinoma Cells. Biosci. Biotechnol. Biochem. 2009, 73, 2183–2189. [Google Scholar] [CrossRef] [Green Version]
  166. Zecchini, S.; Serafini, F.P.; Catalani, E.; Giovarelli, M.; Coazzoli, M.; Di Renzo, I.; De Palma, C.; Perrotta, C.; Clementi, E.; Buonanno, F.; et al. Dysfunctional autophagy induced by the pro-apoptotic natural compound climacostol in tumour cells. Cell Death Dis. 2018, 10. [Google Scholar] [CrossRef] [Green Version]
  167. Saha, S.; Panigrahi, D.P.; Patil, S.; Bhutia, S.K. Autophagy in health and disease: A comprehensive review. Biomed. Pharmacother. 2018, 104, 485–495. [Google Scholar] [CrossRef]
  168. Grossi, C.; Rigacci, S.; Ambrosini, S.; Ed Dami, T.; Luccarini, I.; Traini, C.; Failli, P.; Berti, A.; Casamenti, F.; Stefani, M. The polyphenol oleuropein aglycone protects TgCRND8 mice against Ass plaque pathology. PLoS ONE 2013, 8, e71702. [Google Scholar] [CrossRef]
  169. Luccarini, I.; Grossi, C.; Rigacci, S.; Coppi, E.; Pugliese, A.M.; Pantano, D.; la Marca, G.; Ed Dami, T.; Berti, A.; Stefani, M.; et al. Oleuropein aglycone protects against pyroglutamylated-3 amyloid-ss toxicity: Biochemical, epigenetic and functional correlates. Neurobiol. Aging 2015, 36, 648–663. [Google Scholar] [CrossRef]
  170. Zhu, Z.; Yan, J.; Jiang, W.; Yao, X.G.; Chen, J.; Chen, L.; Li, C.; Hu, L.; Jiang, H.; Shen, X. Arctigenin effectively ameliorates memory impairment in Alzheimer’s disease model mice targeting both beta-amyloid production and clearance. J. Neurosci. 2013, 33, 13138–13149. [Google Scholar] [CrossRef]
  171. Lu, C.; Guo, Y.; Yan, J.; Luo, Z.; Luo, H.B.; Yan, M.; Huang, L.; Li, X. Design, synthesis, and evaluation of multitarget-directed resveratrol derivatives for the treatment of Alzheimer’s disease. J. Med. Chem. 2013, 56, 5843–5859. [Google Scholar] [CrossRef] [PubMed]
  172. de Oliveira, M.R.; Nabavi, S.F.; Manayi, A.; Daglia, M.; Hajheydari, Z.; Nabavi, S.M. Resveratrol and the mitochondria: From triggering the intrinsic apoptotic pathway to inducing mitochondrial biogenesis, a mechanistic view. BBA-Gen. Subj. 2016, 1860, 727–745. [Google Scholar] [CrossRef] [PubMed]
  173. Juhasz, B.; Varga, B.; Gesztelyi, R.; Kemeny-Beke, A.; Zsuga, J.; Tosaki, A. Resveratrol: A Multifunctional Cytoprotective Molecule. Curr. Pharm. Biotechnol. 2010, 11, 810–818. [Google Scholar] [CrossRef] [PubMed]
  174. Filomeni, G.; Graziani, I.; De Zio, D.; Dini, L.; Centonze, D.; Rotilio, G.; Ciriolo, M.R. Neuroprotection of kaempferol by autophagy in models of rotenone-mediated acute toxicity: Possible implications for Parkinson’s disease. Neurobiol. Aging 2012, 33, 767–785. [Google Scholar] [CrossRef]
  175. Lu, J.H.; Tan, J.Q.; Durairajan, S.S.; Liu, L.F.; Zhang, Z.H.; Ma, L.; Shen, H.M.; Chan, H.Y.; Li, M. Isorhynchophylline, a natural alkaloid, promotes the degradation of alpha-synuclein in neuronal cells via inducing autophagy. Autophagy 2012, 8, 98–108. [Google Scholar] [CrossRef] [Green Version]
  176. Sasazawa, Y.; Sato, N.; Umezawa, K.; Simizu, S. Conophylline protects cells in cellular models of neurodegenerative diseases by inducing mammalian target of rapamycin (mTOR)-independent autophagy. J. Biol. Chem. 2015, 290, 6168–6178. [Google Scholar] [CrossRef] [Green Version]
  177. Jiang, T.F.; Zhang, Y.J.; Zhou, H.Y.; Wang, H.M.; Tian, L.P.; Liu, J.; Ding, J.Q.; Chen, S.D. Curcumin ameliorates the neurodegenerative pathology in A53T alpha-synuclein cell model of Parkinson’s disease through the downregulation of mTOR/p70S6K signaling and the recovery of macroautophagy. J. Neuroimmune Pharmacol. 2013, 8, 356–369. [Google Scholar] [CrossRef]
  178. Guo, Y.J.; Dong, S.Y.; Cui, X.X.; Feng, Y.; Liu, T.; Yin, M.; Kuo, S.H.; Tan, E.K.; Zhao, W.J.; Wu, Y.C. Resveratrol alleviates MPTP-induced motor impairments and pathological changes by autophagic degradation of alpha-synuclein via SIRT1-deacetylated LC3. Mol. Nutr. Food Res. 2016, 60, 2161–2175. [Google Scholar] [CrossRef]
  179. Ferretta, A.; Gaballo, A.; Tanzarella, P.; Piccoli, C.; Capitanio, N.; Nico, B.; Annese, T.; Di Paola, M.; Dell’aquila, C.; De Mari, M.; et al. Effect of resveratrol on mitochondrial function: Implications in parkin-associated familiar Parkinson’s disease. Biochim. Biophys. Acta 2014, 1842, 902–915. [Google Scholar] [CrossRef] [Green Version]
  180. Cai, C.Z.; Zhou, H.F.; Yuan, N.N.; Wu, M.Y.; Lee, S.M.; Ren, J.Y.; Su, H.X.; Lu, J.J.; Chen, X.P.; Li, M.; et al. Natural alkaloid harmine promotes degradation of alpha-synuclein via PKA-mediated ubiquitin-proteasome system activation. Phytomedicine 2019, 61, 152842. [Google Scholar] [CrossRef]
  181. Wong, V.K.; Wu, A.G.; Wang, J.R.; Liu, L.; Law, B.Y. Neferine attenuates the protein level and toxicity of mutant huntingtin in PC-12 cells via induction of autophagy. Molecules 2015, 20, 3496–3514. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  182. Sarkar, S.; Davies, J.E.; Huang, Z.; Tunnacliffe, A.; Rubinsztein, D.C. Trehalose, a novel mTOR-independent autophagy enhancer, accelerates the clearance of mutant huntingtin and alpha-synuclein. J. Biol. Chem. 2007, 282, 5641–5652. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Jiang, W.; Wei, W.; Gaertig, M.A.; Li, S.; Li, X.J. Therapeutic Effect of Berberine on Huntington’s Disease Transgenic Mouse Model. PLoS ONE 2015, 10, e0134142. [Google Scholar] [CrossRef] [PubMed]
  184. Castillo, K.; Nassif, M.; Valenzuela, V.; Rojas, F.; Matus, S.; Mercado, G.; Court, F.A.; van Zundert, B.; Hetz, C. Trehalose delays the progression of amyotrophic lateral sclerosis by enhancing autophagy in motoneurons. Autophagy 2013, 9, 1308–1320. [Google Scholar] [CrossRef] [Green Version]
  185. Menzies, F.M.; Huebener, J.; Renna, M.; Bonin, M.; Riess, O.; Rubinsztein, D.C. Autophagy induction reduces mutant ataxin-3 levels and toxicity in a mouse model of spinocerebellar ataxia type 3. Brain 2010, 133, 93–104. [Google Scholar] [CrossRef] [Green Version]
  186. Singhal, J.; Nagaprashantha, L.D.; Vatsyayan, R.; Awasthi, S.; Singhal, S.S. Didymin induces apoptosis by inhibiting N-Myc and upregulating RKIP in neuroblastoma. Cancer Prev. Res. 2012, 5, 473–483. [Google Scholar] [CrossRef] [Green Version]
  187. Rahman, M.A.; Kim, N.H.; Huh, S.O. Cytotoxic effect of gambogic acid on SH-SY5Y neuroblastoma cells is mediated by intrinsic caspase-dependent signaling pathway. Mol. Cell. Biochem. 2013, 377, 187–196. [Google Scholar] [CrossRef]
  188. Vengoji, R.; Macha, M.A.; Batra, S.K.; Shonka, N.A. Natural products: A hope for glioblastoma patients. Oncotarget 2018, 9, 22194–22219. [Google Scholar] [CrossRef] [Green Version]
  189. Yuan, Y.; Xue, X.; Guo, R.B.; Sun, X.L.; Hu, G. Resveratrol Enhances the Antitumor Effects of Temozolomide in Glioblastoma via ROS-dependent AMPK-TSC-mTOR Signaling Pathway. CNS. Neurosci. Ther. 2012, 18, 536–546. [Google Scholar] [CrossRef]
  190. Shah, S.H. Carbonic anhydrase, net photosynthetic rate and yield of black cumin (Nigella sativa) plants sprayed with kinetin. Acta Bot. Croat. 2008, 67, 63–68. [Google Scholar]
  191. Hussain, A.R.; Uddin, S.; Ahmed, M.; Al-Dayel, F.; Bavi, P.P.; Al-Kuraya, K.S. Phosphorylated IkappaBalpha predicts poor prognosis in activated B-cell lymphoma and its inhibition with thymoquinone induces apoptosis via ROS release. PLoS ONE 2013, 8, e60540. [Google Scholar] [CrossRef] [PubMed]
  192. Yamamoto, K.; Ishikawa, C.; Katano, H.; Yasumoto, T.; Mori, N. Fucoxanthin and its deacetylated product, fucoxanthinol, induce apoptosis of primary effusion lymphomas. Cancer Lett. 2011, 300, 225–234. [Google Scholar] [CrossRef] [PubMed]
  193. Zhao, Q.; Guan, J.W.; Qin, Y.H.; Ren, P.; Zhang, Z.W.; Lv, J.; Sun, S.J.; Zhang, C.L.; Mao, W.F. Curcumin sensitizes lymphoma cells to DNA damage agents through regulating Rad51-dependent homologous recombination. Biomed. Pharmacother. 2018, 97, 115–119. [Google Scholar] [CrossRef] [PubMed]
  194. Shi, L.Y.; Qin, H.H.; Jin, X.D.; Yang, X.X.; Lu, X.; Wang, H.G.; Wang, R.Y.; Yu, D.Y.; Feng, B.M. The natural phenolic peperobtusin A induces apoptosis of lymphoma U937 cells via the Caspase dependent and p38 MAPK signaling pathways. Biomed. Pharmacother. 2018, 102, 772–781. [Google Scholar] [CrossRef] [PubMed]
  195. Xiao, X.H.; Li, H.L.; Jin, H.Z.; Jin, J.; Yu, M.; Ma, C.M.; Tong, Y.; Zhou, L.; Lei, H.; Xu, H.Z.; et al. Identification of 11(13)-dehydroivaxillin as a potent therapeutic agent against non-Hodgkin’s lymphoma. Cell Death Dis. 2017, 8. [Google Scholar] [CrossRef] [PubMed]
  196. Martino, R.; Beer, M.F.; Elso, O.; Donadel, O.; Sulsen, V.; Anesini, C. Sesquiterpene lactones from Ambrosia spp. are active against a murine lymphoma cell line by inducing apoptosis and cell cycle arrest. Toxicol. In Vitro 2015, 29, 1529–1536. [Google Scholar] [CrossRef]
  197. Leidal, A.M.; Levine, B.; Debnath, J. Autophagy and the cell biology of age-related disease. Nat. Cell Biol. 2018, 20, 1338–1348. [Google Scholar] [CrossRef]
  198. Sciarretta, S.; Maejima, Y.; Zablocki, D.; Sadoshima, J. The Role of Autophagy in the Heart. Annu. Rev. Physiol. 2018, 80, 1–26. [Google Scholar] [CrossRef]
  199. Meyer, G.; Czompa, A.; Reboul, C.; Csepanyi, E.; Czegledi, A.; Bak, I.; Balla, G.; Balla, J.; Tosaki, A.; Lekli, I. The cellular autophagy markers Beclin-1 and LC3B-II are increased during reperfusion in fibrillated mouse hearts. Curr. Pharm. Des. 2013, 19, 6912–6918. [Google Scholar] [CrossRef]
  200. Shirakabe, A.; Zhai, P.; Ikeda, Y.; Saito, T.; Maejima, Y.; Hsu, C.P.; Nomura, M.; Egashira, K.; Levine, B.; Sadoshima, J. Drp1-Dependent Mitochondrial Autophagy Plays a Protective Role Against Pressure Overload-Induced Mitochondrial Dysfunction and Heart Failure. Circulation 2016, 133, 1249–1263. [Google Scholar] [CrossRef] [Green Version]
  201. Matsui, Y.; Takagi, H.; Qu, X.P.; Abdellatif, M.; Sakoda, H.; Asano, T.; Levine, B.; Sadoshima, J. Distinct roles of autophagy in the heart during ischemia and reperfusion—Roles of AMP-activated protein kinase and Beclin 1 in mediating autophagy. Circ. Res. 2007, 100, 914–922. [Google Scholar] [CrossRef] [PubMed]
  202. Sun, Y.; Yao, X.; Zhang, Q.J.; Zhu, M.; Liu, Z.P.; Ci, B.; Xie, Y.; Carlson, D.; Rothermel, B.A.; Sun, Y.; et al. Beclin-1-Dependent Autophagy Protects the Heart During Sepsis. Circulation 2018, 138, 2247–2262. [Google Scholar] [CrossRef] [PubMed]
  203. Gyongyosi, A.; Szoke, K.; Fenyvesi, F.; Fejes, Z.; Debreceni, I.B.; Nagy, B.; Tosaki, A.; Lekli, I. Inhibited autophagy may contribute to heme toxicity in cardiomyoblast cells. Biochem. Biophys Res. Commun. 2019, 511, 732–738. [Google Scholar] [CrossRef] [PubMed]
  204. Zheng, Y.T.; Shi, B.H.; Ma, M.Q.; Wu, X.Q.; Lin, X.H. The novel relationship between Sirt3 and autophagy in myocardial ischemia-reperfusion. J. Cell Physiol. 2019, 234, 5488–5495. [Google Scholar] [CrossRef] [PubMed]
  205. Zilinyi, R.; Czompa, A.; Czegledi, A.; Gajtko, A.; Pituk, D.; Lekli, I.; Tosaki, A. The Cardioprotective Effect of Metformin in Doxorubicin-Induced Cardiotoxicity: The Role of Autophagy. Molecules 2018, 23. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  206. Gyongyosi, A.; Zilinyi, R.; Czegledi, A.; Tosaki, A.; Tosaki, A.; Lekli, I. The Role of Autophagy and Death Pathways in Dose-dependent Isoproterenolinduced Cardiotoxicity. Curr. Pharm. Des. 2019, 25, 2192–2198. [Google Scholar] [CrossRef]
  207. Dong, Y.; Chen, H.W.; Gao, J.L.; Liu, Y.M.; Li, J.; Wang, J. Molecular machinery and interplay of apoptosis and autophagy in coronary heart disease. J. Mol. Cell. Cardiol. 2019, 136, 27–41. [Google Scholar] [CrossRef] [Green Version]
  208. Lekli, I.; Haines, D.D.; Balla, G.; Tosaki, A. Autophagy: An adaptive physiological countermeasure to cellular senescence and ischaemia/reperfusion-associated cardiac arrhythmias. J. Cell. Mol. Med. 2017, 21, 1058–1072. [Google Scholar] [CrossRef]
  209. Rahman, M.A.; Saha, S.K.; Rahman, M.S.; Uddin, M.J.; Uddin, M.S.; Pang, M.G.; Rhim, H.; Cho, S.G. Molecular Insights Into Therapeutic Potential of Autophagy Modulation by Natural Products for Cancer Stem Cells. Front. Cell Dev. Biol. 2020, 8. [Google Scholar] [CrossRef]
  210. Sohn, E.J.; Park, H.T. Natural agents mediated autophagic signal networks in cancer. Cancer Cell Int. 2017, 17. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Molecular mechanism of autophagy signaling. Autophagy is generally initiated by a deficiency of growth factors or nutrients that trigger AMPK or mTOR inhibition. This stimulates FIP200 and Atg13 associated ULK complex. After that phosphorylation of Beclin-1, it leads to the activation of VPS34, which further initiates the formation of phagophore. Conjugation of Atg5-Atg12 encompasses Atg7 as well as Atg10 to form an Atg12-Atg5-Atg16 complex that also stimulates phagophores formation. Atg5, in addition to Atg12, makes an Atg16 complex, which acts as an E3-function concerning LC3-PE assembly (LC3-II). This complex likewise initiates the formation of a phagophore. Specifically, LC3-II is an autophagy marker that is ultimately interrupted via autolysosomes. Maturation of autophagosome leads to fusion with lysosomes in association with several lysosomal proteins, finally leading to degradation of cargo in addition to recycling of metabolites and nutrients.
Figure 1. Molecular mechanism of autophagy signaling. Autophagy is generally initiated by a deficiency of growth factors or nutrients that trigger AMPK or mTOR inhibition. This stimulates FIP200 and Atg13 associated ULK complex. After that phosphorylation of Beclin-1, it leads to the activation of VPS34, which further initiates the formation of phagophore. Conjugation of Atg5-Atg12 encompasses Atg7 as well as Atg10 to form an Atg12-Atg5-Atg16 complex that also stimulates phagophores formation. Atg5, in addition to Atg12, makes an Atg16 complex, which acts as an E3-function concerning LC3-PE assembly (LC3-II). This complex likewise initiates the formation of a phagophore. Specifically, LC3-II is an autophagy marker that is ultimately interrupted via autolysosomes. Maturation of autophagosome leads to fusion with lysosomes in association with several lysosomal proteins, finally leading to degradation of cargo in addition to recycling of metabolites and nutrients.
Biomedicines 08 00517 g001
Figure 2. In cancer cells, autophagy exhibits a dual role in metastasis and invasion. Depending on the stage and type of tumor cell, autophagy can prevent or stimulate cancer cell maintenance. In the early stages of autophagy in primary tumors, metastasis has been suppressed via anti-metastatic elements in addition to stimulation of tumor conservation along with progression by inducing cellular resistance against TRAIL, a tumor necrosis factor-related apoptosis-inducing ligand, to induce apoptosis. Later, in the course of invasion, tumor cells spread and enter lymphatic or blood vessels. Here, autophagy shows a pro-metastatic function through cell protection from anoikis triggered via extracellular matrix detachment.
Figure 2. In cancer cells, autophagy exhibits a dual role in metastasis and invasion. Depending on the stage and type of tumor cell, autophagy can prevent or stimulate cancer cell maintenance. In the early stages of autophagy in primary tumors, metastasis has been suppressed via anti-metastatic elements in addition to stimulation of tumor conservation along with progression by inducing cellular resistance against TRAIL, a tumor necrosis factor-related apoptosis-inducing ligand, to induce apoptosis. Later, in the course of invasion, tumor cells spread and enter lymphatic or blood vessels. Here, autophagy shows a pro-metastatic function through cell protection from anoikis triggered via extracellular matrix detachment.
Biomedicines 08 00517 g002
Figure 3. Key autophagy pathway and its regulators (natural compounds) for treating diverse cancers. This figure is modified from Wang et al. [64]. Natural compounds activate autophagy by several autophagic signaling targets and show a close relationship with cancer development and control. Multiple signaling pathways are activated or inhibited by natural compounds to modulate cancer cells.
Figure 3. Key autophagy pathway and its regulators (natural compounds) for treating diverse cancers. This figure is modified from Wang et al. [64]. Natural compounds activate autophagy by several autophagic signaling targets and show a close relationship with cancer development and control. Multiple signaling pathways are activated or inhibited by natural compounds to modulate cancer cells.
Biomedicines 08 00517 g003
Figure 4. 18α-Glycyrrhetinic acid-mediated cell death in human neuroblastoma cells. 18-GA induces mitochondria-mediated apoptosis to alter mitochondrial membrane potential in a significant change in with Bax/Bcl-2 ratio. Autophagy-related proteins Atg5/7, Beclin-1, LC3, and p62 were activated during 18-GA-mediated autophagic cell death. Beclin-1 contributes to autophagy in addition to apoptosis induction in 18-GA-mediated cell death. Additionally, ROS and ER stress showed no changes during 18-GA treatment in SH-SY5Y and B103.
Figure 4. 18α-Glycyrrhetinic acid-mediated cell death in human neuroblastoma cells. 18-GA induces mitochondria-mediated apoptosis to alter mitochondrial membrane potential in a significant change in with Bax/Bcl-2 ratio. Autophagy-related proteins Atg5/7, Beclin-1, LC3, and p62 were activated during 18-GA-mediated autophagic cell death. Beclin-1 contributes to autophagy in addition to apoptosis induction in 18-GA-mediated cell death. Additionally, ROS and ER stress showed no changes during 18-GA treatment in SH-SY5Y and B103.
Biomedicines 08 00517 g004
Figure 5. Gintonin (GT) activates autophagy in mouse cortical astrocytes. GT-mediated autophagy was dependent on GPCR-LPA receptors. Pretreatment with an LPA receptor antagonist, Ki16425, considerably decreased GT-mediated autophagy, while LPA agonist pretreatment, LPA 18:1, augmented autophagy in astrocytes. Autophagy initiation proteins Atg5, Atg7, and Beclin-1 were found to be activated in GT-mediated autophagy in astrocytes. Inhibition of autophagy by 3-MA, GT-mediated autophagy was prevented. Pretreated with E-64d/pepstatin A and bafilomycin A1, GT additionally improved LC3 puncta formation indicating enhanced autophagic flux in cortical astrocytes. GT treatment increased acidic vacuole formation along with p62 protein accumulation during this autophagy process.
Figure 5. Gintonin (GT) activates autophagy in mouse cortical astrocytes. GT-mediated autophagy was dependent on GPCR-LPA receptors. Pretreatment with an LPA receptor antagonist, Ki16425, considerably decreased GT-mediated autophagy, while LPA agonist pretreatment, LPA 18:1, augmented autophagy in astrocytes. Autophagy initiation proteins Atg5, Atg7, and Beclin-1 were found to be activated in GT-mediated autophagy in astrocytes. Inhibition of autophagy by 3-MA, GT-mediated autophagy was prevented. Pretreated with E-64d/pepstatin A and bafilomycin A1, GT additionally improved LC3 puncta formation indicating enhanced autophagic flux in cortical astrocytes. GT treatment increased acidic vacuole formation along with p62 protein accumulation during this autophagy process.
Biomedicines 08 00517 g005
Figure 6. Natural compound oxyresveratrol (Oxy R) stimulates apoptosis and autophagy-dependent cell death. Oxy R activates autophagy through autophagy initiation Atg7 and Beclin 1 dependent pathways. Atg7 knockdown prevents autophagy, but activates apoptotic cell death in SH-SY5Y cells. Blockage of autophagy via 3-MA, Oxy R inhibits autophagic cell death. When apoptosis signaling is inhibited by Z-DEVD-FMK caspase-3 inhibitor, Oxy R exhibits autophagic cell death.
Figure 6. Natural compound oxyresveratrol (Oxy R) stimulates apoptosis and autophagy-dependent cell death. Oxy R activates autophagy through autophagy initiation Atg7 and Beclin 1 dependent pathways. Atg7 knockdown prevents autophagy, but activates apoptotic cell death in SH-SY5Y cells. Blockage of autophagy via 3-MA, Oxy R inhibits autophagic cell death. When apoptosis signaling is inhibited by Z-DEVD-FMK caspase-3 inhibitor, Oxy R exhibits autophagic cell death.
Biomedicines 08 00517 g006
Table 1. Mechanism of action of combination of natural compounds and autophagy blockers, 3-methyladenine (3-MA), chloroquine (CQ), and bafilomycin A1 (Baf A1) in modulating autophagy and apoptosis in various cancer cell models.
Table 1. Mechanism of action of combination of natural compounds and autophagy blockers, 3-methyladenine (3-MA), chloroquine (CQ), and bafilomycin A1 (Baf A1) in modulating autophagy and apoptosis in various cancer cell models.
Natural Compound/ChemicalCell ModelMolecular MechanismsCombination with Autophagy InhibitorsReferences
18α-Glycyrrhetinic acidNeuroblastomaAutophagy and apoptosis inductionCQ combination induces cell death[42]
ResveratrolGlioblastomasInduction of apoptosis and autophagy via reduction of ROS/MAPK pathwayBaf A1 combination augments apoptotic cell death [43]
CurcuminGlioblastomasInduction of autophagy and apoptosisCQ combination rises apoptotic cell death[44]
GintoninCortical astrocytes and glioblastoma cells Induction of autophagyCQ and Baf A1 combination stimulates autophagy[45]
QuercetinU373MG cellsPI3K/Akt inhibition and causes apoptosis and autophagyCQ combination increases apoptotic cell death[46]
PaclitaxelHuman lung carcinoma A549 cellsInduction of autophagy and apoptosis3-MA combination enhances apoptotic cell death[47]
GenisteinMIA PaCa-2 human pancreatic cancer cellAutophagy and apoptosis inductionCQ combination stimulates apoptotic cell death[48]
HonokiolNon-small cell lung cancer A549 and H460 cellsp62 and LC3 induces autophagy inductionCombination of CQ triggered apoptosis[49]
Ginsenoside Compound KNeuroblastoma SK-N-BE(2) and SH-SY5Y cellsROS-mediated autophagy inhibition and apoptosis inductionCombination of CQ inhibit autophagy and induces apoptosis[50]
OxyresveratrolNeuroblastoma SK-N-BE(2) and SH-SY5Y cellsPI3K/AKT/mTOR pathway independent autophagy and apoptosis3-MA combination enhances apoptotic cell death[51]
Table 2. Anticancer effects of natural cancers in the management of lymphomas.
Table 2. Anticancer effects of natural cancers in the management of lymphomas.
Natural CompoundsCellular/Animal ModelsConcentrations/DosesEffects and MechanismsReferences
Thymoquinone from Nigella sativaActivated B cell lymphoma cell lines5–10 mMROS production and apoptosis[191]
ResveratrolNeuroblastoma30 µMApoptosis induction and antiproliferation[137]
FuxocanthinolPrimary effusion lymphomas BCBL-1 and TY-11.3–5 µMApoptosis induction and cell cycle arrest[192]
CurcuminCH12F3 lymphoma cells5 µMCaspase-3 dependent apoptosis and DNA damage[193]
Peperobtusin ALymphoma U937 cells25, 50, 100 μMCaspase-3, 8, 9 dependent apoptosis and p38 MAPK activation[194]
11(13)-dehydroivaxillin (DHI)Lymphoid malignancies of NHL cells xenografts5, 7, 10 μMInduction of NF-κB and apoptosis[195]
Psilostachyin CMurine lymphoma cell line BW51470.01–50 μg/mLInduction of apoptosis, necrosis, and ROS generation[196]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Rahman, M.A.; Rahman, M.H.; Hossain, M.S.; Biswas, P.; Islam, R.; Uddin, M.J.; Rahman, M.H.; Rhim, H. Molecular Insights into the Multifunctional Role of Natural Compounds: Autophagy Modulation and Cancer Prevention. Biomedicines 2020, 8, 517. https://0-doi-org.brum.beds.ac.uk/10.3390/biomedicines8110517

AMA Style

Rahman MA, Rahman MH, Hossain MS, Biswas P, Islam R, Uddin MJ, Rahman MH, Rhim H. Molecular Insights into the Multifunctional Role of Natural Compounds: Autophagy Modulation and Cancer Prevention. Biomedicines. 2020; 8(11):517. https://0-doi-org.brum.beds.ac.uk/10.3390/biomedicines8110517

Chicago/Turabian Style

Rahman, Md. Ataur, MD. Hasanur Rahman, Md. Shahadat Hossain, Partha Biswas, Rokibul Islam, Md Jamal Uddin, Md. Habibur Rahman, and Hyewhon Rhim. 2020. "Molecular Insights into the Multifunctional Role of Natural Compounds: Autophagy Modulation and Cancer Prevention" Biomedicines 8, no. 11: 517. https://0-doi-org.brum.beds.ac.uk/10.3390/biomedicines8110517

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

Article Metrics

Back to TopTop