Next Article in Journal
Design, Synthesis, and Anti-Fungal Evaluation of Heterocyclic Benzoxazole Derivatives
Next Article in Special Issue
Natural Polyphenols for Treatment of Colorectal Cancer
Previous Article in Journal
Molecular Cloning, Expression, and Functional Analysis of Glycosyltransferase (TbUGGT) Gene from Trapa bispinosa Roxb.
Previous Article in Special Issue
Pestalotiopsis Diversity: Species, Dispositions, Secondary Metabolites, and Bioactivities
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Natural Products as Anticancer Agents: Current Status and Future Perspectives

1
Key Laboratory of Modern Preparation of Chinese Medicine, Ministry of Education, Jiangxi University of Chinese Medicine, Nanchang 330004, China
2
Key Laboratory of Pharmacodynamics and Safety Evaluation, Health Commission of Jiangxi Province, Nanchang Medical College, Nanchang 330006, China
3
Key Laboratory of Pharmacodynamics and Quality Evaluation on Anti-Inflammatory Chinese Herbs, Jiangxi Administration of Traditional Chinese Medicine, Nanchang Medical College, Nanchang 330006, China
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(23), 8367; https://doi.org/10.3390/molecules27238367
Submission received: 9 November 2022 / Revised: 22 November 2022 / Accepted: 28 November 2022 / Published: 30 November 2022
(This article belongs to the Special Issue Bioactive Compounds in Food Bioscience and Pharmacology)

Abstract

:
Natural products have been an invaluable and useful source of anticancer agents over the years. Several compounds have been synthesized from natural products by modifying their structures or by using naturally occurring compounds as building blocks in the synthesis of these compounds for various purposes in different fields, such as biology, medicine, and engineering. Multiple modern and costly treatments have been applied to combat cancer and limit its lethality, but the results are not significantly refreshing. Natural products, which are a significant source of new therapeutic drugs, are currently being investigated as potential cytotoxic agents and have shown a positive trend in preclinical research and have prompted numerous innovative strategies in order to combat cancer and expedite the clinical research. Natural products are becoming increasingly important for drug discovery due to their high molecular diversity and novel biofunctionality. Furthermore, natural products can provide superior efficacy and safety due to their unique molecular properties. The objective of the current review is to provide an overview of the emergence of natural products for the treatment and prevention of cancer, such as chemosensitizers, immunotherapeutics, combinatorial therapies with other anticancer drugs, novel formulations of natural products, and the molecular mechanisms underlying their anticancer properties.

Graphical Abstract

1. Introduction

Noncommunicable diseases (NCDs) refer to diseases that are not transmitted directly from one individual to another, and account for more than 70% of all deaths worldwide (41 million deaths per year). Cancer is the second leading cause of death among NCDs, after cardiovascular disease [1]. The incidence and mortality of cancer are rapidly increasing throughout the world due to an increase in the aging population. The reasons are complex, including environmental pollution, chemical toxins pollution, ionizing radiation, free radical toxin, microorganisms (bacteria, fungi, viruses, etc.) and their metabolic toxins, genetic characteristics, endocrine imbalance, and immune dysfunction, etc. [2]. Global Cancer Observatory (GLOBOCAN) estimates that 19 million new cancer cases will be reported and 10 million deaths will be caused by this disease in 2020. Cancers in both sexes are most commonly diagnosed in the breast, lung, colon, prostate, and stomach (excluding non-melanoma skin cancer). It is estimated that by 2040, the incidence of cancer in the world population will increase to 30.2 million cases and the mortality rate will increase to 16.3 million cases, respectively [3].
The term “cancer” is used to refer to a large group of diseases (over 277) which are characterized by the uncontrolled growth of abnormal cells and which are caused by a wide range of factors [4]. In fact, any cell in the body is capable of becoming a cancer cell (carcinogenesis) once it has been subjected to a series of successive gene mutations. Specifically, this process occurs in three stages, namely: (a) the initiation stage, when the alteration has already been identified, (b) the promotion stage, where the altered cells are considered totally mutated and then the malignancy of the cells commences, and (c) the progression stage, once the tumor has already grown, and as the cells begin to divide in an accelerated and irreversible manner. Cancer cells may also spread to other parts of the body (metastasis) in addition to growing locally, which is an often cited cause of death from the disease [5].
Cancer research has always been a challenge because of its complexity. Different types of cancer may exhibit substantial differences in relation to genetic alterations, organ involvement, prognosis, and therapeutic management [6]. Even though various treatment options are available, their success relies upon the type and phase of the disease. Among various treatment alternatives, careful surgical removal of malignant tissues/tumor, radiation treatment, chemotherapy, and immunotherapy are generally employed. The effects of surgery and radiotherapy are local, while those of chemotherapy and targeted therapy are systemic. The type and stage of cancer determine whether such therapies are used alone or in combination with other treatments (e.g., radiotherapy and chemotherapy) [7]. Small molecule targeted therapy and chemotherapy are two therapeutic approaches used to treat cancer using chemical compounds. In general, chemotherapy drugs act as cytotoxic agents that interfere with various phases of the cell cycle. The rationale for their application is that cancer cells generally have a faster division rate than normal cells, making them more susceptible to chemotherapeutic agents [1]. Generally, these drugs can be classified into five categories according to their biochemical properties: alkylating agents (for example, cisplatin), antimetabolites (5-fluoroacil), antitumor antibiotics (doxorubicin), topoisomerase inhibitors (topotecan), and tubulin-binding drugs (paclitaxel) [8].
Despite the effectiveness of chemotherapeutic drugs, they can also cause adverse reactions in normal cells, including nausea, vomiting, mucositis, alopecia, neuropathy, alopecia, and myelosuppression. Furthermore, they have been found to be associated with multidrug resistance (MDR), an undesirable phenomenon responsible for more than 90% of the deaths of cancer patients undergoing chemotherapy [9]. Small molecule targeted therapy (SMTT) differs from chemotherapy in that it uses chemicals that target specific molecular targets within cancer cells. It is thought that these targets are modified genetically in cancer and are essential to the development and survival of tumors. The majority of the time, they are implicated in signaling pathways that are dysregulated during cancerous development [10]. The use of target-oriented compounds in clinics includes tyrosine kinases, proteasomes, and poly ADP-ribose polymerase inhibitors such as imatinib, carfilzomib, and ribociclib. The specificity of SMTT drugs is expected to result in a less toxic effect on healthy cells. However, side effects (such as a rash, diarrhea, or hypertension) have been reported [11]. Moreover, they may also trigger mechanisms that lead to the development of drug resistance. Even though these restorative choices are effective in treating different sorts of malignant growths, they have impediments; for example, recurrence of cancer, noncompliance because of extremely unfavorable implications such as fatigue, pain, nausea, anemia, emesis, and baldness, among the symptoms endured by patients. It is important to note that the vast majority of synthetic chemotherapeutic drugs developed before failed to satisfy the needs during clinical trials, despite the higher cost of expenditure. Thus, efforts continue to be made to find better alternatives that balance efficacy and toxicity and comply with drug resistance prevention measures.
The heterogeneous nature of cancer has limited the efficacy of conventional therapies such as radiation and standard chemotherapy in treating and preventing it, as they are likely to kill both normal and cancerous cells in the process, resulting in serious hematological toxicities and damage to the tissues involved [12]. A growing number of patients acquire or develop multiple drug resistance, making chemotherapy a treatment with limited therapeutic benefits. A number of anticancer drugs have also been associated with significant side effects, including cardiotoxicity caused by doxorubicin [13], ototoxicity caused by cisplatin, and cognitive impairment caused by 5-fluorouracil [14]. The adverse effects of chemotherapy on patients, such as kidney damage, gastrointestinal problems, hair loss, and fatigue, compromise adherence to treatment. A negative perception of treatment is also a result of these factors [15].
The majority of the chemotherapeutic drugs used in the clinic today are directed at only one particular target such as specific nucleic acids, particular proteins, or tumorigenic pathways. Platinum drugs, such as oxaliplatin, carboplatin, and cisplatin, are known to inhibit nucleotide synthesis and metabolism along with damaging DNA. Tyrosine kinase inhibitors such as gefitinib, erlotinib, and icotinib are directed at tyrosine kinase. Angiogenesis is inhibited by bevacizumab, sunitinib, and sorafenib [16]. The treatment of cancer has been greatly improved by the development of many drugs, but the disease is still an evolutionary process. Cancer cells are capable of adapting to a given drug treatment. Natural products have been observed to influence multiple oncogenic signaling pathways simultaneously by modulating the activity or expression of their molecular targets. Various natural products affect multiple pathways, including apoptotic cell death, cell proliferation, migration/invasion, angiogenesis, and metastasis. Natural products are capable of generating intracellular signals that trigger events that lead to the death of cancer cells. Chemotherapy can be enhanced by natural products in cases where cancer is resistant to treatment. The high cost of conventional drugs and the rising incidence of cancer have challenged researchers to find more cost-effective and eco-friendly alternatives. In this context, natural products are highly advantageous due to their chemical diversity, low toxicity, safety, and availability, which make them an attractive and affordable alternative to synthetic products [17,18].
In this review, we summarize novel studies and viewpoints on cancer therapy and natural products mostly from 2012 to 2022 (older sources were also cited if necessary) using Web of Science (http://www.webofknowledge.com, accessed on 1 July 2022), Google Scholar (https://scholar.google.com, accessed on 1 April 2022), PubMed (https://www.ncbi.nlm.nih.gov/pubmed, accessed on 1 June 2022), ScienceDirect (https://www.sciencedirect.com, accessed on 1 May 2022), Chinese National Knowledge Infrastructure (CNKI, https://www.cnki.net, accessed on 1 March 2022), Scopus (https://www.scopus.com, accessed on 1 April 2022), and Clinical Trials (https://clinicaltrials.gov, accessed on 1 November 2021). This study provides important insights into the efficacy and mechanism of action of natural products in the treatment of cancer, potentially improving clinical outcomes.
The review provides a comprehensive perspective on the anti-tumor potential of natural products by describing the outcomes of the main in vitro and in vivo experiments that have demonstrated some effects on various forms of cancer. Furthermore, this study summarizes the six directions of their anticancer activity by summarizing the mechanisms by which these agents affect cellular proliferation, differentiation, apoptosis, angiogenesis, and metastasis. In particular, it includes (I) compounds that possess an innate antitumor effect; (II) reverse chemoresistance; (III) inhibit metastasis; (IV) act as cancer immunotherapeutic; and (VI) the development of novel drug delivery systems for natural products. These natural products were chosen due to their potential anticancer properties as well as their substantial evaluation of chemotherapeutic effectiveness. It will provide the reader with an updated perspective on natural therapies and their potential application in cancer treatment in the future. As a result, this article offers a broader range of options for researchers interested in developing new alternatives for treating and preventing cancer, a topic that continues to grow in importance throughout the world.

2. Natural Products (NPs) against Cancer

Natural medicine refers to substances that are produced naturally by living organisms, such as plants, insects, animals, aquatic organisms, and microbes, and possess pharmacological or biological properties [19]. Natural products are precious gifts from nature that can be used for the prevention and treatment of illnesses in humans. Therefore, they are of vital importance and play an irreplaceable role in the development and design of drugs. Since ancient times, natural products have been used to treat human diseases. Recent research indicates that natural products still have the potential to be applied in drug development. The latest review by David J. Newman and Gordon M. Cragg revealed that 32% of all small-molecule drugs approved between January 1981 and September 2019 were natural products and their derivatives [20,21]. Moreover, from 1981 to 2014, 51% of all the 1211 new small molecule drugs approved worldwide were compounds derived from natural products [21]. In addition, a report by Eric Patridge and colleagues, published in 2016, revealed that the US Food and Drug Administration (FDA) approved 547 natural products and their derivatives for use as medications across the years (1827 to 2013) [22]. They are used to treat a number of diseases, primarily cancer, bacterial infections, and hypertension. The same study found that 68% of all 136 small-molecule anticancer drugs available from 1940 to 2014 were natural products based.
One of the most important sources of biologically active compounds is the plant kingdom. Currently, there are more than 350,000 vascular plant species registered in the world, and new species are being added every year [1]. It remains a vast and unexplored field of study that offers many opportunities for drug discovery. Plants can be used therapeutically in a variety of forms, such as tea, extracts, and dyes. In addition, their active compounds may be isolated and used as medicines or as precursors for synthetic and semi-synthetic drugs. There is a large list of phytochemicals (i.e., chemical compounds produced by plants) with therapeutic activity, including terpenes, alkaloids, essential oils, flavonoids, gums, and a variety of primary and secondary metabolic compounds [23]. In the past few decades, phytochemicals have been extensively studied for their potential anticancer properties, with the objective of using them in cancer treatment modalities such as chemotherapy and targeted therapy. For example, berberine is a secondary metabolite produced by plants, commonly found in the roots, rhizomes, and stem barks of Chinese herbs and plants of the Berberis genus. Numerous preclinical and limited human studies have demonstrated that berberine exerts beneficial biological activities against a variety of human diseases, including inflammation, metabolic dysfunction, depression, cardiovascular diseases, neurodegenerative diseases, and different types of cancers [24,25].
Since natural products exhibit unique properties as compared to conventional synthetic molecules, they are both advantageous and challenging in drug discovery. NPs are characterized by a wide range of scaffolds and a high degree of structural complexity [26]. These molecules generally have a higher molecular mass, a higher number of sp3 carbon and oxygen atoms, fewer nitrogen and halogen atoms, higher numbers of H-bond acceptors and donors, a lower calculated octanol–water partition coefficient (cLogP values, which indicate a higher degree of hydrophilicity), and a significantly higher molecular rigidity than synthetic compounds. A number of these differences can be advantageous; for example, the higher rigidity of NPs may be helpful when tackling protein–protein interactions in the drug discovery process. In fact, NPs are one of the most important sources of oral drugs beyond Lipinski’s rule of five [6]. Over the past 20 years, the increasing molecular mass of approved oral drugs illustrates the growing importance of drugs that do not comply with this rule. The structural optimization of NPs has been influenced by evolution to serve specific biological functions, such as regulating endogenous defense mechanisms and interacting (often in competition) with other organisms, which is why they are important for treating diseases such as cancer and infectious diseases. Additionally, the traditional use of these substances may provide insight into their efficacy as well as safety. Furthermore, the NPs pool contains a wide range of ‘bioactive’ compounds compared to typical synthetic small molecule libraries that cover a narrower chemical space [27].
Natural products account for approximately 50% of anti-tumor drugs. They may be derived from plants or can be semi-synthetic chemicals. Currently, many plant-based antitumor drugs are in clinical use, such as taxanes (e.g., Taxol), vinblastine, vincristine, and podophyllotoxin analogs. Robert Noble and Charles Beer discovered vincristine and vinblastine vinca alkaloids in the 1950s from the leaves of Catharanthus roseus (Madagascar periwinkle) [28]. Furthermore, the Taxus genus produces taxane-derived drugs. In 1971, paclitaxel, which is known commercially as Taxol®, was first isolated from Taxus brevifolia Nutt (Pacific yew). This is one of several diterpene taxanes that are used as the main chemotherapeutic agent against several forms of cancer, including ovarian, lung, and breast cancers. During the 1980s, another semi-synthetic taxane drug called docetaxel (Taxotere®) was isolated from Taxus baccata (European yew). Recent developments in chemical modification have focused on podophyllotoxin analogs, particularly teniposide and etoposide, two semi-synthetic antitumor agents. In 1880, Podwyssotzki discovered podophyllotoxin in the North American mayapple Podophyllum peltatum L. Furthermore, podophyllotoxin has also been found in Podophyllum emodi (Indian podophyllum) [29]. In 1966, etoposide was synthesized for the first time and was approved by the United States Food and Drug Administration (FDA) in 1983 as an anticancer agent [30].
Natural products can be used for chemoprevention and chemotherapy since it effectively suppresses cell proliferation, regulates the cell cycle, and interferes with several tumorigenic signaling pathways, such as phosphoinositide 3-kinase (PI3K), matrix metalloproteinase (MMP), MAPK/ERK known as the Ras-Raf-MEK-ERK pathway, toll-like receptor (TLR) pathway, and AKT pathway. Additionally, natural products may stimulate DNA repairing mechanisms through the action of p21, p27, p51, and/or p53 gene products, such as Bax, Bak, and Bid proteins, which cause the synthesis of protective enzymes such as caspases 3, 7, 8, 9, 10, 12, and modulate antioxidant enzymes such as GST, GSH, and GPx. The chemo-preventive effects occur by inhibiting important events involved in tumor initiation such as ROS removal, enhancing DNA repair, and eradicating transformed cells through immunosurveillance. Furthermore, these effects are exerted through inhibition of proliferation and immunotargeting of altered cells, as well as hindering tumor progression by promoting differentiation, inhibiting angiogenesis, and inducing apoptosis [31]. Curcumin has been one of the most studied phytochemical and its anticancer mechanism is shown in Figure 1.

2.1. Natural Products against Chemoresistance

Cancer develops from the abnormality of different molecular paths; therefore, treatment with current standard chemotherapeutic agents still remains less effective due to targeting of one or a few pathways. As a result, higher doses of conventional drugs are required to eliminate the tumor, which evokes harmful effects even worse than the condition itself. Although these agents are understood to kill tumor cells, they also trigger many survival pathways, which eventually induce cancer drug resistance, generally termed chemoresistance, which develops when cancer cells establish the resistance capability against chemotherapy and are categorized into two types: intrinsic (preexistent) and acquired (drugs induced). Chemoresistance development in tumor cells occurs via different molecular paths, e.g., regulation of drug influx and efflux via ATP binding cassette (ABC) transporter, epigenetic factors, hindering cell death, altering metabolism and degradation of drugs, and deactivation of chemotherapeutic drugs, mutations of drug targets, improved DNA repair and modifying growth factors signaling. These factors may act alone or in combination with different signaling pathways (Figure 2) [33].
Over the decades, the Food and Drug Administration (FDA) has approved about 300 chemotherapeutic drugs for the treatment of cancer, including taxol, vinca and its derivatives, Platinium analogues, 5-fluorouracil, bevacizumab, erlotinib, nivolumab, ipilimumab, sunitinib, and olaparib, with other different chemotherapeutic drug combinations. However, most of them are highly toxic and induce resistance, which ultimately results in tumor recurrence and metastasis, limiting their anticancer activities due to the development of adaptive resistance by targeted tumor cells and non-specific toxicities towards normal cells [34]. Therefore, better alternatives which are safer and more effective are required to overcome chemoresistance and subsequently sensitize tumor cells to chemotherapeutic agents. Chemosensitization is a widely used approach in which a drug’s activity is augmented using other drugs to suppress chemoresistance. However, the chemosensitizing agent needs to be safe and possess multitargeted inhibition potential of various chemoresistance pathways. The combinatorial chemotherapy produces synergistic effects by suppressing proteins, genes, and pathways involved in chemoresistance and concurrently modulates molecular targets of cancer to improve conventional drug therapy. Intriguingly, in vitro, ex vivo, in vivo, and clinical results suggest that phytochemicals are less toxic, safer, easily available, and highly effective with multi-targeting potential against different malignancies and also reduce the toxicity of conventional chemotherapy by increasing sensitivity of tumor cells [35].
Lupeol has been shown to sensitize cancer therapy-resistant cells by modulating various inflammatory cytokines such as IL-6, TNF-α, IFN-γ and also BAX, BCL-2, FAS, Caspases, Survivin, and PI3K-AKT-mTOR pathways and Cyclins, CDKs, P21, P53, and PCNA molecules, which are involved in cell cycle regulation [36]. Shikonin, a natural naphthoquinone, induces necroptosis in various cancer cell lines such as glioma, osteosarcoma, and hematopoietic cells by enhancing RIP1 and RIP3 and mROS, and also effectively overcomes P-gp, Bcl-2, Bcl-xl, MRP1, and BCRP1 drug resistance [37]. Similarly, Pterostilbene induces apoptosis, autophagy, S-phase cell cycle arrest, and inhibiting MDR1 expression via downregulation of RAGE/P13k/Akt pathways [38]. Epigallocatechin gallate induces G2/M cell cycle arrest and also suppresses MDR levels [39]. Plumbagin inhibits multiple molecular pathways such as cell cycle arrest, antiangiogenic, apoptotic, and autophagic pathways; regulates associated genes such as Akt, STAT3, and Nf-kβ; induces ROS; suppresses glutathione; and causes breakage of DNA strands [40]. Quercetin, a natural flavonoid, has been shown to facilitate cell death and chemosensitivity in pancreatic cells by RAGE/P13/AKT/mTOR axis [41]. Muthusamy et al. showed that ferulic acid reverses the P-gp mediated MDR by inhibiting P13K/AKT/Nf-kβ pathway in a tumor xenograft model [42].

2.2. Natural Products against Metastasis

Metastasis is a highly complicated process that involves primary tumor mass, circulating tumor cells (CTCs), and abnormal tumor microenvironments and is a leading contributor to cancer-linked deaths. The spread of metastasis to a specific organ is determined by various factors: for example, observing patterns of circulation, such as receiving direct blood from the primary site, e.g., hepatic metastasis happens in patients with colorectal cancer. Vascular permeability promotes extravasation at tumor sites, which further increases the probability of metastasis because instead of being autonomous, tumor cells become involved in bidirectional communication with metastatic microenvironments and modify the antitumoral immunity, genomic stability, extracellular milieu, survival signaling, chemoresistance, and proliferation activities. Despite improvements in advanced technologies, metastasis remains a major problem to resolve. The main therapeutic objectives are to prevent metastasis in high-risk patients, reducing advanced lesions, and avoiding additional metastasis with the limited disease [43].
Unlike conventional anticancer therapeutics, phytochemicals modulate multiple pathways during tumor initiation, progression, and metastasis development. Phytochemicals modulate primary epigenetic mechanisms regulating genes in metastasis such as DNA methylation, modification of histones, and/or non-coding RNA (ncRNA) associated multigene silencing. Regarding the complexity of metastatic tumors, multiple drugs can be used, such as the therapeutic approach used against (HIV) virus, because monotherapy can produce some initial responses. However, it lacks durability, so effective and durable therapy is required which minimizes the formation of resistance and improves durability. Phytochemicals can provide a better option due to their compatibility with multiple agents, which needs to be explored in detail. Some of the commonly explored and proven phytochemicals with anticancer and antimetastatic activities are apigenin, allicin, α-carotene, baicalein, and berberine curcumin, wogonin, formononetin, gambojic/ursolic/ellagic acid, papaya pectins, sulforaphane, and isothiocyanates [44].
In the recent past, increasing proof of toxicity related to chemotherapeutic agents, even at the smallest therapeutic doses, has paved the way for exploring a nature-based approach and various medicinal plant extracts/phytochemicals have been utilized at different stages of cancer progression [45]; their general mechanisms are shown in Figure 3.

2.3. Natural Products and Cancer Immunotherapy

Tumor immunotherapy is relatively safe and effective, such as immune checkpoint blockade, and is considered a major breakthrough in recent decades. Both types (innate and adaptive) of anti-tumor immune responses can be initiated. Tumor-suppressing natural agents exert their action by upregulating immune responses in the tumor microenvironment, primarily via M1/M2-type macrophages. Recently, Sun et al. found that resveratrol can inhibit the polarization of human monocyte-derived macrophages into M2, and they also observed that lung cancer medium induced polarized M2 macrophages were modified to M1 polarized macrophages after resveratrol treatment by increasing IL-12 and TNF-ɑ and reducing IL-10 levels. Additionally, M2 macrophage markers’ expression was decreased, such as MRC1, chiI3, CCL24, and Retnla [47]. Moreover, Loo et al. successfully increased the immunity in postoperative patients with Rhodiola algida mixture by increasing the expression of IL-2, IL-4, GM-CSF, and mRNA content, thereby decreasing the increased risk of oral cancer [48].
Improving the functional activity of dendritic cells in cancer patients can significantly improve the treatment outcome by activating T cells, as mature dendritic cells secrete IL-12, which acts on T cells and encourage Th1 cell differentiation [49]. Chang et al. showed that the treatment of DCs with Astragalus and Codonopsis polysaccharide could promote the proliferation of CD4+ and CD8+ T cells in mice, and DC-based vaccine increased the anti-metastatic efficacy in 4T1 cells bearing mice by increasing the expression of CD40, CD80, and CD86 levels on the surface of DCs [50]. Similarly, it has been shown that shikonin, combined with B16 tumor cell lysate, could activate DCs to a more mature state with increased surface levels of CD86 and MHC II, which activates Th1 cells and improve its functions. Subsequently, the shikonin-tumor cell load DCs vaccine improved the cytotoxicity of splenocytes to cancer cells, hampered tumor growth, and increased mice’s survival [51].
NK cells help in differentiating stem cells and undifferentiated cancer cells by secreting IFN-γ and TNF-α, which limit the tumor growth by remodeling/regulating the tumor microenvironment. Natural products activate NK cells to stop tumor growth and metastasis [52]. Wu et al. have shown that lupeol can activate NK cells, which further activates the PI3K/Akt and Wnt/β- catenin pathway by increasing the expression of IFN-γ, PFP, and CD107a, as a result, inhibiting the proliferation of gastric cells (N87, HGC27 and BGC823) [53]. In addition, Hou et al. reported that total flavonoids obtained from Hippophae rhamnoides increased the cytotoxicity of NK cells against tumor cells by increasing the expression of NKp44, NKp46, granzymes B, and perforin [54].
MDSCs, a heterogeneous group of cells, promotes tumor progression, angiogenesis, and metastasis to healthy cells/tissues. Recently, Zhou et al. studied the effects of icariin on 4T1-Neu xenograft mice and noted that the number of MDSCs decreased in the spleen of the mice. Furthermore, icariin also restored the secretion IFN-γ by CD8+ T cells in the mice and reduced nitric oxide and ROS levels in MSDCs in vivo, while in vitro treatment differentiated into dendritic cells and macrophages [55]. The Shugan Jianpi formula has recently been reported to effectively regulate tumor microenvironment by suppressing CD8+ cells apoptosis, reducing MDSC proliferation and tumor activity in 4T1 tumor cells bearing mice [56].
Peripheral T lymphocytes CD4+ and CD8+ play a key part in anti-tumor immunity, and CD4+ T cells help in recruiting CD8+ T cells, while CD8+ T cells directly attack and kill tumor cells. CD4+ T cells can be differentiated into many subgroups of cells, such as Th1, Th2, Th9, Th17, Tfh, Treg cells, etc. The Th1/Th2 imbalance plays a vital role in the occurrence of tumors [57], and natural products increase Th1 development, which leads to the suppression of tumors. Wei et al. reported that with the addition of Tetramethylpyrazine phosphate (plant extracts) to the culture of lung cancer patient-derived PBMC cells and healthy subject cells, the expression of IFN-γ and IL-2 was increased, and Th2 cytokines decreased in patient-derived cells; the killing efficiency of macrophages was also improved [58]. Eckol is derived from Ecklonia cava, brown algae, and has shown potent antiproliferative and immunomodulatory activities. In a recent in vivo antitumoral study, the cytotoxic effects of eckol were evaluated by using sarcoma S180 xenograft model. The results stated that all the pro-apoptotic and antiproliferative proteins were upregulated, eckol also stimulated (MPS) and activated dendritic cells, which produced Th1 responses and enhanced CD4+/CD8+ T lymphocyte ratio, as a result cytotoxic T lymphocytes responses were increased [59]. Takei et al. observed that Ginsenoside promotes the transformation of naive T cells into Th1 cells by acting on DCs, increasing IFN-γ level, and in CTL assays, it was found that the Ginsenoside induced more IFN-γ than the mature DCs, which indicates that combined treatment may exert strong immunogenic response [60].
In recent times, cancer immunotherapy has gained some success by engineering antigen-specific T cells, e.g., TCR-T and CAR-T cells. Especially in hematological cancer, the CAR-T targets the CD19 and has been used and allowed for clinical use by the US FDA. Similarly, TCR-T cells have been used to treat cancer; however, MHC-restriction hinders their potential clinical use. There are no reports available which show the use of CAR-T cells and TCR-T cells combined with natural products, but this is an emerging area of research and needs to be explored, and we might obtain some successful outcomes in treating various cancers [61].
Astragalus polysaccharides have been shown to inhibit the proliferation of Treg cells in a dose- and time-dependent manner by reestablishing the cytokine balance and decreasing the expression of Foxp3 in the microenvironment of HCC. Moreover, Astragalus polysaccharides can block the (SDF-1) or its receptor (responsible for recruiting Treg cells into the HCC microenvironment) via the CXCR4/CXCL12 signaling pathway, thus inhibiting Treg cells migration [62]. Astragalus polysaccharides also act as an immunomodulatory agent by activating the TLR4 pathway, which inhibits TGF-β, resulting in a decrease in the number of Treg cells [63]. Similarly, Radix glycyrrhizae polysaccharides have been shown to downregulate the number of Treg cells in the microenvironment of H22 tumor xenograft mice by decreasing the expression of Foxp3 in Treg cells while upregulating the Th1/Th2 ratio in serum; as a result, inhibiting tumor growth [64].
Recently, targeting immune checkpoint molecules (PD-1 and CTLA4) has been considered a promising approach to treat various diseases such as cancer. These checkpoint molecules play an essential role in regulating immune homeostasis [65]. Currently, immune checkpoint molecules include programmed death receptor 1 (PD-1), cytotoxic T-lymphocyte associated antigen 4 (CTLA4), lymphocyte activation gene 3 (LAG3), and T cell immunoglobulin mucin 3 (Tim3). Presently, very little work is done on using natural products to target immune checkpoint molecules, and this research area needs more exploration [66]. Zhang et al. showed that Qiyusanlong decoction suppressed tumor growth in Lewis lung carcinoma-bearing mice by decreasing both mRNA and PD-1/PDL-1 protein levels in the tumor [67]. Similarly, another study revealed that Gegen-qinlian decoction increased the effects of PD-1 blockade in colon cancer via remodeling of intestinal microbiota, which demonstrated that phytochemicals could block PD-1 and thereby could be a possible approach for treating various cancers [68]. Zhao et al. showed that curcumin suppresses Treg cells (CD4+ CD25+) function by decreasing CTLA4 and Foxp3 [69].

2.4. Natural Products in Combination with Other Chemotherapeutic Drugs

Currently, combinatorial therapy has gained importance due to the administration of multiple chemotherapeutics with different biochemical targets, which increases efficacy as well as safety, and this concept is widely applied to different tumors. Many studies have reported the combination of widely used phytochemicals like (Resveratrol, curcumin, and thymoquinone) with other antitumoral drugs and have achieved significant success in preclinical studies [70]. Combination therapy reduces the toxicities associated with traditionally used chemotherapeutics such as doxorubicin (cardiomyopathy) and cisplatin (nephrotoxicity and immunosuppression), etc. The other advantage of combination therapy is overcoming drug resistance, which reduces adverse events by using low doses with the same efficacy and sometimes better pharmacological effects (synergism). Many preclinical studies and clinical trials are conducted to use phytochemicals as adjuvant therapy in various cancers. Chemoresistance has been commonly endured by many traditional chemotherapeutics, which is the main concern and hurdle in tumor therapy along with radiation resistance. Natural products are ideal candidates for chemosensitizing tumor cells and increasing the efficacy of existing drugs [71]. Silybinin has been shown to help doxorubicin in overcoming drug resistance in colorectal cancer by inhibiting GLUT1 expression. Many natural agents act as antioxidants, such as ROS scavenging/antioxidation, thereby reducing the toxicities related to the generation of free radicals. However, drug–drug interaction could be the main disadvantage of combinatorial therapy [72].
Curcumin has been used in combination with many chemotherapeutic drugs for the treatment of various cancers. For example, curcumin sensitizes tumor cells to 5-fluorouracil, which inhibits the tumor growth in HCT116 cells/xenograft model via inhibition of AMPK/UKL1 modulation of the AKT pathway [73]. Similarly, curcumin augmented doxorubicin’s anti-tumor activity by suppressing cell migration and inducing apoptosis in neuroblastoma SH-SY5Y cells via upregulation of p21 p53, and TIMP1 and downregulation of MMP2 [74]. Resveratrol and paclitaxel combination is reported to produce synergistic effects through the stimulation of TRPM2 channel in glioblastoma DBTRG cell line, which resulted in oxidative stress and induced apoptosis, which shows that resveratrol can be used in combination with other chemotherapeutic agents which can enhance the treatment efficiency [75]. Shen et al. reported that noscapine increases the sensitivity of cisplatin by modulating the cell cycle (G2/M phase arrest) and inducing apoptosis (downregulating XIAP and NF-kβ, while upregulating caspase-3) in cisplatin-resistant ovarian SKOV3 cells and xenograft in mice [76]. In addition, Neferine has been shown to increase the antitumor activity of cisplatin in lung cancer A549 cells via G1 phase cell cycle arrest, hypergeneration of ROS, upregulating Bax, BAk, p53, and c-myc levels while downregulating Bcl-2, FAK, VEGF, MMP-2/-9, and loss of membrane potential (ΔΨM) [77]. Wang et al. used cryptotanshinone in combination with paclitaxel against tongue squamous cell carcinoma (CAL 27 and SCC) cell lines, which increased the antitumor activity by inhibiting the migration and proliferation of cells as well as inducing apoptosis due to inhibition of JAK/STAT3 signaling pathway [78].
Similarly, Zhang et al. showed that a combination of formononetin and temozolomide could produce synergistic effects in C6 glioma cells through upregulation of Bax and caspase-3/-9 expression while downregulating Bcl-2 and MMP-2/-9 expressions [79]. Furthermore, Tseng et al. reported that aloe emodin enhanced the cytotoxicity of tamoxifen in different breast cancer cells (MDA-MB 231, MCF-7, HCC1954, and BT-474 cells) via suppression of Ras/ERK and PI3K/mTOR signaling pathway [80]. Similarly, a combination of capsaicin and docetaxel has been shown to have higher cytotoxicity against prostate cancer (LNCaP and PC3) cell lines and tumor xenograft in mice by inhibiting the growth and proliferation via reducing Akt, mTOR, and S6 phosphorylation while increasing (AMPK) phosphorylation [81].
Recently, Saikia et al. reported that combined use of heteronemin with cytarabine increases its sensitivity against leukemia (HL-60) cells by downregulating downstream targets of Ras signaling pathway such as AP-1, NF-kβ, AMPK, and c-myc; as a result it produces synergistic effects, inhibits the growth, and induces apoptosis [82]. Furthermore, the combined use of celastrol and triptolide has been shown to produce synergistic effects against various cancer models in vitro (H1299, H460, SKOV3, OVCAR3, Hela, SIHA, and SW480 cell lines) and in vivo studies (xenograft in mice) by inhibiting growth and proliferation through increasing ROS, G2/M phase arrest, decreasing Akt, surviving, and EGFR expressions [83]. Moreover, the combined use of berberine and sorafenib is effective in liver carcinoma because berberine increases the sensitivity of HCC (HepG2 and SMMC-7721) cells to sorafenib, which ultimately inhibits the growth and proliferation and induces apoptosis by increasing the expression of cleaved PARP and cleaved caspase-3, while decreasing the expression of Bcl-2 and VEGF markers [84]. Desai et al. studied the combined effects of biochanin A and Temozolomide against U-87 and T-98 MG cells and observed that the growth and proliferation of the cells were synergistically inhibited and the expression of p-p53 was increased, while the expression of p-Akt, EGFR, p-ERK, and MT-MMP1 was decreased [85]. Mangiferin increased the sensitivity of oxaliplatin towards MCF-7, Hela, and HT29 cell lines by increasing the expression of caspase-3, S-phase cycle arrest, and reducing Nf-kβ activation, which induced apoptosis [86].
In a recent study, dietary phytochemical fisetin was shown to produce synergistic effects in combination with sorafenib against cervical cancer (Hela) cells by activating DR5-mediated caspase-3/8 activity and mitochondrial apoptotic pathway [87]. In another study, the combined use of ursolic acid and temozolomide has been shown to produce synergistic effects in in vitro and in vivo studies in mice by overcoming temozolomide-resistance due to the downregulation of O6-methylguanine DNA-methyltransferase (MGMT), which resulted in an increase in cytotoxicity and inhibited the proliferation of the tumor cells, which indicates the potential of ursolic acid as a monotherapeutic anticancer agent as well as a chemosensitizer [88]. Similarly, a combination of lipoic acid and other antineoplastic drugs (5-FU and Doxorubicin) resulted in a decrease in p53-mediated stabilization of p21, which resulted in the synergistic killing of colorectal cancer cells [89]. Recently, Huang et al. showed that vitamin D sensitizes oral squamous carcinoma cells to cisplatin by inhibiting lipocalin-2 (LCN2) modulated Nf-kβ signaling pathway via ribosomal protein S3 (RPS3), which partially reversed the cisplatin resistance [90]. Moreover, the combined use of embelin (XIAP inhibitor) and celastrol (NF-kβ inhibitor) in acute myeloid leukemia (HL-60) cell line has been shown to produce synergistic and additive effects by downregulating COX-2 and survivin, which shows that simultaneous inhibition of the two tumor signaling pathways can improve the effectiveness of the chemotherapy [91].
The combination of withaferin A and oxaliplatin produced synergistic effects in pancreatic cancer (in vitro and in vivo studies) by increasing the generation of ROS, which ultimately led to the inhibition of the PI3K/AKT signaling pathway [92]. Yang et al. reported that tectorigenin increases the sensitivity of paclitaxel in paclitaxel-resistant ovarian tumor (SKVO3, MPSC1, and A27800) cell lines by activating caspase-3/-8/-9 and downregulating XIAP, Bcl-2, COX-2, FLIP, IkB, IKK, and Akt, which suggests that it increases the sensitivity and cytotoxicity by inactivating Akt/IkB/IKK/NF-kβ pathway [93]. Moreover, glabridin, a flavonoid, has been shown to reverse the P-gp mediated resistance to the conventional chemotherapeutic drugs (Doxorubicin and paclitaxel) in P-gp overexpressing breast cancer (MDA-MB-231/MDR1 and MCF-7/ADR) cell lines by suppressing P-gp and downregulating the functional activity of P-gp- ATPase and thereby reversing the multidrug resistance [94]. Similarly, Xia et al. showed that gambojic acid promotes gemcitabine’s sensitivity in pancreatic cancer (in vitro and in vivo studies) by inhibiting the ERK/E2F1/RRM2 signaling pathway, which decreased the proliferation and growth of tumors. Moreover, the study also indicated that gambojic acid can be used as a monotherapeutic agent, as well as a chemosensitizer [95]. In addition, the combined use of forbesione with 5-fluorouracil produced synergistic effects in inhibiting growth and proliferation in cholangiocarcinoma cancer (in vitro and in vivo studies using Ham-1 cells), through downregulating Bcl-2 and increasing expression of p53, Bax, Apaf-1, and caspase 3/-9, which induces apoptosis and results in the killing of the tumor cells [96]. Similarly, Su et al. showed that osthole enhances cisplatin’s effects by suppressing NRF2 expression and blocking the PI3K/Akt signaling pathway, thus inhibiting the progression of cisplatin-resistant cervical cancer (SiHa/cDDP) cells in vitro and in an in vivo xenograft model [97].

3. Natural Products against Some Common Types of Cancer

3.1. Lung Cancer

Lung cancer is highly occurring and a standout amongst the most widely recognized human malignancies in both developed and developing nations, with 2.1 million new lung cancer cases and 1.8 million deaths anticipated in 2018, nearly one of every five (18.4%) malignancy deaths [98,99]. Even though various treatment alternatives are available, including surgery, radiotherapy, chemotherapy, and targeted therapy, half of all newly diagnosed cancers are now at an advanced stage, where the effects of treatment are constrained. Treatment in such cases might be restricted to palliative care, leading to a low 5-year survival rate. Lung cancer comprises a couple of subtypes; for example, adenoma-carcinoma (AdCa), the most widely recognized subtype in non-smokers and females; squamous cell carcinoma (SqCC), and small cell lung cancer (SCLC). A few instances of current treatment for malignant lung growth incorporate medical procedures, chemotherapy, and radiotherapy, whereby the choice of treatment relies upon the subtype and stage of lung cancer. Nevertheless, there is a dire need to find other alternative chemotherapeutic agents, either fabricated as a more effective drug or in combination with the current therapeutic agents to have a more prominent and synergistic anticancer activity. Recently, Aucubin has been reported to produce antitumoral effects in lung cancer (A549) cells by inducing cell cycle arrest at G0/G1 phase, inducing p53 activation, and also increasing the activity of the Fas/Fas-ligand system [100]. Similarly, Xu et al. reported that cytisine, a natural alkaloid, produces significant cytotoxic effects against lung cancer A549, NCI-H23, and NCI-H460 cells and in vivo (rat model) by inducing apoptosis via an increase in ROS and loss of membrane potential; increasing BAD, cleaving PARP, and cleaving caspase-3 expressions; and decreasing Bcl-2, pro-PARP, and pro-caspase-3. Moreover, phosphorylation of p38, JNK, and I-kB was increased, while there was a significant decrease in the phosphorylation of ERK, NF-kβ, and STAT3. Furthermore, cytisine arrested the cell cycle at the G2/M phase, which was related to the inhibition of the Akt signaling pathway [101].
In another study, Han et al. showed that L-securinine from dried leaves of Securinega suffruticosa inhibited the proliferation of A549 cancer cells, decreased the expression of DKK1 genes, and promoted the methylation of DKK1 promoter in comparison to 5-azacytidine [102]. Similarly, gedunin isolated from Azadirachta indica induced apoptosis and inhibited the growth and proliferation of lung cancer A549 cells by generating ROS, decreasing the membrane potential, and causing DNA damage via downregulation of PIK3CA, EGFR, AKT, and autophagy. Furthermore, gedunin also disrupted the interaction between Hsp90: Bcl-1:Bcl-2 [103]. In another study, vanillin, the main ingredient of Vanilla planifolia, suppressed the CSC-like behavior of lung cancer NCI-H460 cells by inducing Akt-proteasomal degradation and decreasing downstream CSC transcription factors (Oct4 and Nanog) [104]. Additionally, cepharanthine isolated from Stephania cepharantha inhibited the growth and proliferation of lung cancer H1299 and A549 cells by increasing ROS, loss of membrane potential, and cell cycle arrest [105].

3.2. Breast Cancer

In 2018, there were about 2.1 million newly diagnosed female breast malignant growth cases around the world, representing nearly one of every four tumor cases among women. Breast cancer is the most widely diagnosed cancer by far in most nations (154 of 185) and is also the primary source of cancer-related deaths in more than 100 nations. For malignant breast tumor treatment, numerous alternatives are provided, for example, medical procedures, hormonal treatment, radiation treatment, chemotherapy, and targeted treatment [106]. However, there are also certain limitations (i.e., narrow therapeutic index with non-specific toxic consequences for healthy tissues, increasing chance of infection, etc.), and also the severe side effects persist for months or years, even after treatment completion. Therefore, further studies are required to explicate the underlying fundamental mechanisms of breast cancer to develop new therapeutic strategies. Breast cancer is the collection of abnormal cells, presumably credited to the imbalanced proliferation of cells, apoptosis, and the cluttered autophagy regulation. Various natural products were accounted for as possible anti-cancer agents or thought of as immediate or indirect sources of new chemotherapeutic adjuvants to upgrade the efficacy or enhance the symptoms through autophagy regulation [107]. Higenamine has been demonstrated to increase the anticancer activity (apoptosis and G2/M cell cycle arrest) of cucurbitacin B in breast cancer (T47D and SkBr3) cells by inhibiting Akt and CDK2 [108].
In another study, Citral, isolated from Cymbopogon citrates, suppressed the growth and proliferation of 4T1 breast cancer cells implanted in the nude BALB/c mice [109]. Similarly, euphol inhibited the growth and proliferation of breast cancer (T47D) cells by inducing G1 phase cell cycle arrest via downregulation of cyclin D1 and upregulation of p21 and p27 expressions [110]. In another study, Reddy et al. demonstrated that strophanthidin produces dose-dependent cytotoxic effects against breast (MCF7), liver (HepG2), and lung (A549) cancer cells by attenuating MAPK, Wnt/β-catenin, and PI3K/Akt/mTOR signaling pathways [111]. Moreover, jatrophone isolated from Jatropha isabelli inhibited the proliferation and EMT of triple-negative breast cancer cells by interfering with the cell cycle and Wnt/β-catenin signaling pathways [112]. Deng et al. reported that rotenone induced apoptosis in MCF7 breast cancer cells by producing ROS, chromatin condensation, upregulating Bax, downregulating Bcl2, and cleaving PARP. Furthermore, cJNK and P38 MAPKs were activated, and ERK (1/2) signaling pathways were inactivated [113]. Similarly, Dhandayuthapani et al. demonstrated bromelain to be effective against breast cancer by promoting apoptosis via upregulation of c-JNK, p38 kinase, and activating caspase 3 and 9 [114]. In another study, flavokawain B induced apoptosis and inhibited the proliferation of breast cancer (MCF-7 and MDA-MB231) cells by upregulating various kinases such as p-Akt, p-JNK, p-CREB, p-p53, and p-WNK1 and downregulating Nf-kβ, COX2, MMP-9, GLUT1, and VEGFA [115].

3.3. Ovarian Cancer

Ovarian cancer is one of the leading causes of death associated with the female reproductive system in the Western world. Ovarian cancer is a standout amongst the most deadly gynecological diseases in the female reproductive system, influencing approximately one out of 75 women in the United States. Even though the first-line treatment may profit about 80% of patients with ovarian malignancy, 75% of those patients still experience tumor recurrence, which causes a major worry surrounding the treatment of ovarian cancer in patients. Cisplatin is the most generally utilized chemotherapeutic entity for treating ovarian cancer. Treatment dereliction and death in over 90% of patients with metastatic malady are believed to be brought about by drug resistance. Adverse reactions and resistance developed to platinum-based chemotherapy have turned into an obstacle for ovarian cancer treatment. Therefore, it is important to search for new compounds to treat cancer and reduce the associated side effects of the treatments.
Natural bioactive agents have received an increased consideration in cancer treatment lately [116]. For example, diarylheptanoid hirsutenone potentiates the TRAIL-induced apoptotic activity in ovarian cancer (OVCAR-3 and SKOV-3) cells by enhancing the activation of caspase-8 and BID dependent signaling pathways, which further activates caspase-3/-9, exhibiting the significance of hirsutenone in combination with TRAIL treatment [117]. Paclitaxel, first isolated from Taxus brevifolia, has been used in the management of lung, breast, ovarian, prostate, sarcoma, leukemia, and endometrial cancer; it acts by interfering in spindle function, stabilizing microtubules, inducing apoptosis, cell cycle arrest, and inhibiting the growth of tumor cells [118]. Moreover, barbamine isolated from Berberis amurensis effectively inhibited the growth of cancer in vitro (SKOV3 and Es2 cell lines) and in vivo (SKOV3 xenograft model) by inducing apoptosis and targeting the Wnt/β-catenin pathway [119]. Recently, Nobilietin has also been shown to produce strong antitumor effects in ovarian cancer (HOCCs) cells by increasing the expression of cleaved PARP in a dose-dependent manner and inhibiting proliferation, causing DNA damage and inducing apoptosis. Many in vitro and in vivo studies have demonstrated that various flavonoids produce potent cytotoxic activities by inducing apoptosis, cell cycle arrest, and inhibiting angiogenesis [120].

3.4. Colon Cancer

Colorectal cancer (CRC) is prominent amongst the most occurring tumors worldwide, and as of late, it is the third-highest contributor to cancer-associated deaths in the United States of America and China [121]. It represents over 90% of the malignant tumors of the large bowel, and the other 10% comprises lymphoma and squamous cell carcinoma. The occurrence of colorectal cancer is continually increasing due to weak prognosis in patients possessing widely metastasized tumors, and the fundamental underlying mechanism of metastasis is not completely clear [122]. A few epidemiological investigations have exhibited the relationship of colon tumor growth with dietary propensities, for example, low fiber diet, high fat intake, and low calcium/micronutrient consumption [123]. The inverse correlation between the utilization of vegetables and fruits with different cancers has led researchers to examine the advantages of dietary elements in chemoprevention [124]. Recently, Morin has been shown to induce apoptosis and inhibit the growth of colon cancer (SW480) cells through the generation of ROS and decrease in mitochondrial membrane potential. Furthermore, it increased PARP, Bax, and cleaved caspase-3/-8/-9 expressions [125]. In another study, inflexinol isolated from Isodon excisus has shown the potential to strongly inhibit the growth and proliferation of colon cancer cells by inducing apoptosis through inactivation of NF-kβ in in vitro (SW620 and HCT116 cell lines) and in vivo (SW620 xenograft) rat model [126]. In addition, strychnine from Nux vomica has shown significant antiproliferative effects in vitro (SW480, Lovo, and DLD1) and in vivo (DLD1 xenografted model) by targeting the Wnt/β-catenin pathway by enhancing APC and decreasing β-catenin and c-Myc. Furthermore, the expression of DKK1 was also enhanced, known for negatively regulating the Wnt/β-catenin pathway [127].
In another study, Gao et al. showed that berginin significantly suppressed the proliferation of HCT116 cells by accumulating intracellular ROS, DNA damage, and G1 phase arrest through inhibition of PI3K/AKT/mTOR pathway [128]. Similarly, psoralidin inhibited the viability and proliferation of SW480 cells by inhibiting the Nf-kβ and Bcl-2/Bax signaling pathway [129]. In a recent study, matrine was also shown to produce antitumor effects in vitro against multiple colorectal cancer (LS174T, SW1116, Caco-2 and RKO) cells and in vivo in (LS174T xenografted) mice via G1/G0 phase arrest, reducing Bcl-2 and caspase-3, while upregulating cleaved caspase-3 and Bax expressions [130]. Magnolol isolated from Magnolia officinalis shows significant antiproliferative effects in vitro and in vivo against HT-29 and CT-26 cell lines by inducing apoptosis through intrinsic and extrinsic pathways inhibiting PKCδ/NF-kβ signaling pathway [131]. Similarly, Mi et al. showed that imperatonin isolated from Angelica dahurica suppresses the tumor growth, proliferation, and angiogenesis of human colon cancer (HeLa, Hep3B, and HCT116) cells through HIF-1α targeting through mTOR/p70S6K/4E-BP1 and MAPK signaling pathways [132]. Moreover, aloperinole is an alkaloid extracted from the leaves of Sophora alopecuroides which produces potent cytotoxic effects against colon cancer HCT116 cells by inducing apoptosis and inhibiting the proliferation by G2/M phase cell cycle arrest and increasing the expression of Bax, p21, and p53, while decreasing Bcl-2, CD1, and B1. Furthermore, it also inhibited PI3K/Akt and JAK/Stat3 pathway [133]. Many preclinical and clinical studies have indicated that bioactive components of fruits and vegetables may help in the deterrence of colon cancer.

3.5. Brain Cancer

Brain tumors represent 85–90% of all significant CNS tumors. Glioblastoma (GBM) represents around half of all aggressive brain tumors and is related to low survival. Multimodal treatments, including surgery, pursued by adjuvant chemoradiation therapy (CRT) with temozolomide (TMZ), is a standardized treatment for GBM. The TMZ increases overall survival (OS) from 7.7 to 13.5 months and from 7.9 to 10.0 months in the GBM patients, which is extremely poor [134]. This poor survival is likely due to numerous variables, including systemic toxicity due to higher TMZ dosages, BBB impermeability, CRT resistance, and progression of refractory tumors [135]. Hence, identifying a novel chemotherapeutic agent that can modulate the BBB, restrain tumor development, and stop tumors recurrence is vital for better patient prognosis. In recent studies, ginkgetin extracted from Cephalotaxus fortune has been shown to effectively inhibit the growth, invasion, and proliferation of medulloblastoma (Daoy and D283 MB) cell lines by arresting cell cycle at (G2/M phase), and also inhibit Wnt/β catenin signaling pathway [136]. Similarly, Cao et al. reported that toosendanin shows higher cytotoxic effects in vitro and in vivo against glioblastoma multiforme (U87 and C6 cell lines) by inducing apoptosis via upregulation of ERβ and functional activation of p53 [137]. Recently, Noman et al. isolated prestegane B from Thymelaea microphylla, evaluated it for antiproliferative properties against C6 and Hela cell lines, and found that it shows higher antiproliferative and radical scavenging ability than some of the other known drugs [138]. Extracts of Viscum album are widely used as complementary medicine for cancer therapy, and recently some of the extracts such as aviscumine, iscador Qu, and ML-1 have been shown to regulate the expression of genes associated with the cell migration, invasion, adhesion, and even cell architecture formation in glioma cells by inhibiting TGF-β, SMAD2, and MMP-2/-9. These ingredients have the potential to be used as an adjuvant therapy for the treatment of glioblastoma [139].

3.6. Liver Cancer

Liver cancer was the sixth most commonly reported cancer and the fourth leading cause of cancer-related deaths worldwide in 2018, with around 841,000 new cases and 782,000 deaths every year. Liver cancer is the most widely recognized tumor of the digestive system, with a high mortality rate. Based on conformational evidence, numerous natural dietary compounds are potential hotspots for the prevention and management of liver malignancy [140]. Hepatocellular carcinoma (HCC) is the primary type of malignant liver tumors (70–80%), trailed by intrahepatic cholangiocarcinoma. The major risk factors for liver cancer are hepatitis B/hepatitis C disease, alcohol usage, aflatoxin B1, and metabolic issues. Liver cancer is mostly aggressive with a poor prognosis, having a low five-year survival rate below 9% [141]. Proper interventions such as liver resection, transplantation, and percutaneous removal are considered as the best approaches with curing potential for liver cancer. However, because of various injuries and extrahepatic metastasis, merely 20% of liver cancer patients are appropriate for a medical procedure. On the contrary, chemotherapeutic medications for liver cancer are constrained, and sorafenib is the most well-known medicine. The phase III clinical trials exhibited that sorafenib could increase the overall survival and progression time. However, its clinical gains are modest, and it was noted that sorafenib was only valuable for around 30% of patients, and resistance developed in half a year. Moreover, other issues such as hepatotoxicity, drug resistance, recurrence, and other unwanted side effects exist in present treatments, which compels researchers to look for an alternative therapy [142].
Recently, echinoside A and ds-echinoside, isolated from pearsonothuria graeffei, inhibited tumor growth and proliferation in the HepG2 cells by G0/G1 phase arrest and induced apoptosis via the mitochondrial pathway [143]. In another study, oleuropein has been shown to effectively inhibit the cell viability and proliferation of hepatocellular carcinoma (HepG2) cells by inducing apoptosis through ROS production, increasing Bax, and decreasing Bcl-2. Moreover, oleuropein also targeted the PI3k/Akt signaling pathway [144]. Similarly, crocin isolated from saffron has been shown to produce autophagic apoptosis and inhibit the growth and proliferation of human hepatic cancer (HCCLM3 and HepG2) cells via upregulation of LC3-II and constraining the functional activities of key proteins involved in Akt/mTOR signaling pathway such as p-Akt, p-mTOR, and p-p70S6K [145]. In addition, Alpinia oxyphylla extracts inhibited the proliferation and tumor growth in hepatocellular carcinoma (HepG2, Hep3B, Bel-7402, and SMMC-7721) cell lines and also in vivo Hep3B xenograft by increasing Bax and caspase-3/-9 expressions and decreasing Bcl-2 levels. Furthermore, it also upregulated PTEN and downregulated PI3k and inhibited the phosphorylation of Akt [146]. Alpidine, a cyclic depsipeptide isolated from Aplidium albicans, has produced antiproliferative effects in various tumors such as melanoma, breast and lung cancer via activation of JNK/p38 MAPK pathway and induces apoptosis [147]. Recently, corilagin has been shown to produce significant cytotoxic effects in vitro and in vivo against liver cancer SMMC7721 cells via G2/M phase cycle arrest, and by downregulating p-Akt and cyclin B1/cdc2 and upregulating p-p53 and p-p21Cip1 [148].

3.7. Head and Neck Cancer

Head and neck squamous cell carcinoma (HNSCCs) is among the sixth most common malignancies globally and accounts for 90% of all the head and neck tumors occurring at different anatomical sites. Alcohol consumption, tobacco use, and infection with the human papillomavirus (HPV) are thought to be the oncogenic drivers. Treatment options involve surgical resection, radiotherapy, and targeted chemotherapy, but are mostly ineffective and subsequent relapse frequently occurs due to tumor resistance. Along with treatment failure, these therapies result in higher morbidity and reduced quality of life due to their severely toxic nature [149]. This inadequate therapeutic response can be associated with alterations in intracellular signaling pathways and significant changes in the extracellular tumor microenvironment. Therefore, a better, safe, and effective alternative is needed to address these challenges to improve the treatment standard and reduce the treatment cost. Over the decade, various novel approaches have been considered to understand and target specific disease targets but have fallen short of achieving clinically significant results [150].
Phytochemicals have shown enormous potential in chemosensitizing and inhibiting the growth of HNSCCs in various in vitro and in vivo models. Recently, active constituents of Piper methysticum (kava) such as flavokawain A, flavokawain B, and yangonin have been shown to exhibit antiproliferative effects against OSCC cancer (BICR56 and H400) cells but are non-toxic to normal keratinocytes (OKF6), showing promise for in vivo translation. Additionally, Pycnogenol, isolated from pine bark, inhibited the cell viability and neoplastic transformation in (HSC-3 and TPA-treated JB6) cells by inducing apoptosis through ROS generation, and also increased the expression of cleaved PARP, Caspase-3, and Bax [151]. Moreover, α-mangostin extracted from Garcinia mangostana inhibited the growth and proliferation of human OSSC cells (HSC-2/-3/-4) by inducing apoptosis through a decrease in mitochondrial membrane potential and translocation of cytochrome c and also arrested cell cycle at G1 phase by downregulating (CDks/cyclins) [152]. Thymol isolated from thyme and oregano possesses potent antitumor activity and induces mitochondrial dysfunction and apoptosis in OSCC in vivo model [153]. Polyphenon E, in combination with erlotinib, has been used in clinical trials for the treatment of HSSCs (NCT01116336). Similarly, cucurbitacin B, in combination with other chemotherapeutic drugs, has shown antitumor activity against HSCC and breast cancer [154]. Several in vitro studies have shown that grape seed proanthocyanidins and EGCG produce cytotoxic effects by inducing apoptosis and cell cycle arrest at the G0/G1 phase in HSCC cell lines through various mechanisms [155].

3.8. Prostate Cancer

Prostate cancer is the most commonly diagnosed deadliest cancer in men after lung cancer in the world, having a mortality rate of 3.6% in 2018 and 1.05% increase is expected by 2040 [156]. The estimated rate of prostate cancer diagnosed in men is (one in nine) and mortality (1 in 41), which is quite alarming [157]. The etiological evidence of prostate cancer incidence, progression, and development is not very clear, and various risk factors such as age, race, family history, genetic/somatic mutations, and lifestyle have been linked to it. In particular, age has been a critical factor, and 66 years is the average age of diagnosis [158]. Like other cancers, early diagnosis and treatment are extremely important, but the asymptomatic nature of the disease and complicated diagnostic procedures and treatment expenses make it difficult to manage it properly. Different bioactive compounds have been evaluated and have shown potential to be used for the prevention and treatment of prostate cancer. For example, Shukla et al. tested apigenin against prostate cancer and found that the tumorigenesis in TRAMP mice was suppressed through inhibition of the Nf-kβ pathway [159]. Similarly, afzelin derived from Nymphaea odoratum exhibits anti-tumoral activities against prostate cancer (LNCaP and PC-3) cells by arresting cell cycle at G0 phase and also inhibits the expression of LIM domain kinase-1 [160]. Recently, plectronthoic acid, a novel compound, demonstrated potent significant anti-tumor activities by inhibiting proliferation, induced G0/G1 phase arrest in prostate cancer (PC3, DU145, and CW22Rv1) cells via upregulation of p21/CIP1 and p27/KIP1, and also suppressed mTOR/S6K signaling pathway [161]. In another study, anacardic acid has been shown to inhibit the proliferation and induce apoptosis in human LNCaP cells by autophagy through ER stress/DAPK3/Akt pathway [162]. Delphinidin induces p53 mediated apoptosis by suppressing HDAC function and activating acetylation of p53 in LNCap cells [163]. Moreover, lycorine extracted from Amaryllidaceae plants inhibited the growth, proliferation, migration, and invasion of multiple prostate cancer cells in vitro (LNCaP, PC-3M, DU145, and 22 RV1) and in vivo (PC-3M xenograft) model by abrogating p-STAT3 expression, reversed EMT via STAT3-mediated twist decrease, and inhibited EGF-induced JAK/STAT signaling [164].
Lall et al. established that fisetin effectively inhibits the synthesis of hyaluronan (HA) and increases antiangiogenic high molecular mass-HA and can be used to manage prostate cancer [165]. Similarly, punicalagin, a polyphenol, exhibited apoptotic activity and inhibited the proliferation of PC3 and LNCaP cells via upregulating caspase-3/-8 expressions and decreasing vascular network formation in the CAM model, which also confirmed its antiangiogenic effect [166]. Furthermore, Zeylenone isolated from Uvaria grandiflora Roxb decreased the cell viability, invasion, and metastatic growth of human PCa (DU145) cells by downregulating the Wnt/β-catenin pathway [167]. Moreover, mangiferin has been reported to possess immunomodulatory, apoptotic, antiangiogenic, and gene regulatory effects in vitro, ex vivo, and in vivo in different cancers, especially prostate cancer [168]. Similarly, anethole inhibited the proliferation of prostate cancer PC-3 cells and induced apoptosis by generating ROS, decreasing mitochondrial membrane potential, DNA damage, activating caspase-3 and -9, increasing Bax/Bcl-2 ratio, and leading to G2/M phase cell cycle arrest [169]. In addition, Paeonol inhibited tumor growth in vitro and in vivo by activating intrinsic and extrinsic apoptotic pathways and also inhibited p13k/Akt signaling pathway [170]. Similarly, oleandrin inhibits tumor progression by deregulating multiple pathways such as MAPK, Nf-kβ, and p13k/Akt pathway [171]. Ramamoorthy et al. reported that reserpine isolated from Rauwolfia serpentine induced apoptosis and arrested cell cycle at the G2 phase in prostate cancer (PC3) cells [172].

3.9. Hematological Cancer

Adverse events associated with the use of current therapies such as toxicities, neuropathy, or incessant relapsing have paved the way for using phytochemicals in the treatment of various hematological malignancies such as leukemia, lymphomas, multiple myelomas, Wald Enstrom’s macroglobulinemia, and other hematological diseases such as hemolytic anemia and thrombocytopenia. The frequently used natural products in hematological malignancies are anthracyclines and anthracenediones, vinca alkaloids, isothiocyanates, podophyllotoxin derivatives, polyphenols, and other antioxidants [173,174]. These phytochemicals are also used to sensitize conventional therapeutics such as dexamethasone. Curcumin, agaricus, and neovastat are currently in clinical trials for the treatment of multiple myelomas [175]. Vinca alkaloids such as vincristine and vinblastine are widely used as chemotherapeutic agents for various solid tumors and hematological disorders. Alkaloids such as harringtonine and isoharringtonine, which are isolated from Cephalotaxus harringtonia, act as chemotherapeutic agents and are used in the treatment of acute and chronic myelogenous leukemia and myelodysplastic syndromes. Isothiocyanates are another useful family of electrophilic bioactive compounds such as sinigrin, glucotropaeolin, gluconasturtiin, and glucoraphanin, and are involved in the inhibition of leukemic cells growth and regulate progression and differentiation of tumors, cell cycle, and apoptotic mechanisms. In a recent study, Wu et al. reported that rocaglamide breaks the TRAIL-induced resistance in vitro and in vivo in multiple myeloma and acute leukemia by inhibiting c-FLIP (main factor in TRAIL therapy resistance) expression and thereby increasing the effects of camptothecin, which shows that combined use of rocaglamide can be a suitable therapeutic option [176]. In addition, conophylline, an alkaloid isolated from Tabernaemontana divaricata, suppressed the pancreatic cancer desmoplasia and associated cytokines (IL-6, IL-8, CCL2, and CXCL12) produced by cancer-associated fibroblasts (CAF) and stellate cells. Moreover, the inhibitory and apoptotic effects of conophylline increased more when used in combination with gemcitabine [177]. Most recently, Tubulosine from Alangium salvifolium wang has selectively inhibited JAK3 signals by binding to ATP-binding active site of the kinase (JAK3), thereby reducing the progression and survival of hematopoietic cancer, as shown against HDLM-2, L540, U266, and BKO-84 cancer cells [178].
Similarly, betulinic acid has shown promising cytotoxic effects in various tumors. Phytochemicals provide a wide range of cellular effects by preventing carcinogens from reaching targeted sites and helping in ROS detoxification, increasing immunosurveillance to eliminate transformed cells, activating DNA repair mechanisms, and inhibiting proliferative tumor pathways [31]. In a recent study by Karami et al., gaillardin, a sesquiterpene lactone isolated from Inula ocular-christi, has been shown to produce higher cytotoxic effects in leukemic cells (NALM-6 and MOT-4 with IC50 of 6.1 and 7.3 µM, respectively) by arresting G0/G1 phase and inducing apoptosis, but it shows no cytotoxicity in normal cells [179]. Medicarpin, a natural phytoalexin, has been reported to sensitize myeloid leukemia cells to TRAIL treatment by upregulating pro-apoptotic markers (Cytochrome-c, tBid, Bax, CHOP) and downregulating anti-apoptotic proteins (Bcl-xl, Bcl-2, c-FLIP, XIAP, and survivin). Furthermore, it causes G2/M cell cycle arrest and also increases DR5 expression via activation of the ROS-JNK-CHOP signaling pathway [180]. Furthermore, taxodione isolated from taxodium distichum has been found to induce apoptosis in leukemia k562 cancer cells by generating ROS. Moreover, BCR-ABL, Akt, and STAT5 are sequestered in the mitochondria and are unable to stimulate proliferation [181].

3.10. Miscellaneous Cancer

Many different types of cancers have been affecting and threatening the lives of people, which include skin, stomach, oral cavity, rectum, gastric, gallbladder, pancreas, cervix uteri, penis, kidney, bladder, thyroid, Hodgkin lymphoma, non-Hodgkin lymphoma, leukemia, and osteosarcoma [182]. The interest in using natural products to treat cancer has been rising due to the better efficacy, low toxicity, and lower cost. Recently, Shivamadhu et al. showed that praecetrullus fistulosus lectins inhibited the growth and proliferation of a variety of tumor cells such as Hela, MCF-7, K562, and HT29 cells and also increased the survival of EAC bearing mice by decreasing MMP 2/9 activity and inducing apoptosis in the tumor cells [183]. In another study, oxolane derivatives from Morus alba, such as odisolane, have been proven to inhibit angiogenesis by reducing VEGF, p-ERK effectively, and p-Akt expressions, and can be employed to stop neovascularization in tumors [184]. Recently, hispiloscine and hispidacine alkaloids isolated from Ficus hispida linn showed excellent cytotoxic activity against various cell lines such as breast, lung, and colon, and also demonstrated vasorelaxant activity in rats [185]. Recently, Kim et al. demonstrated that bakuchiol inhibited the viability and EGF-induced neoplastic transformation of skin epithelial carcinoma A431 cell line in vitro and in vivo studies through the inhibition of Blk, Hck, p38MAPK, and Akt/p70S6k pathways [186]. Similarly, cryptolepine inhibited non-melanoma skin cancer proliferation by damaging DNA, S- phase cell cycle arrest, and decreasing membrane potential [187]. Curcubitacin B suppressed the invasion and proliferation of gastric cancer cells through STAT3 inhibition, which also induced apoptosis and furthermore in combination with cisplatin produced increased cytotoxic effects, which indicate that curcibitacin B is a promising STAT3 inhibitor [188]. Phycocyanin is another promising molecule isolated from seaweed, and it exerts its effects by arresting the cell cycle at the G2/M phase in MDA-MB-231, HT29, and A549 cells by decreasing cyclin E and CDk2 expressions and upregulating p21. Furthermore, it activates the mitochondrial apoptotic pathway while inhibiting the proliferative pathways such as PI3K/Akt/mTOR, MAPK, and Nf-kβ pathway [189].
In another study, Liang et al. demonstrated that isovitexin suppresses the stemness of osteosarcoma (U2OS and MG63) cells and induces apoptosis by disrupting the DNMT1/miR-34a/Bcl-2 nexus [190]. Furthermore, melatonin, a natural hormone produced by the pineal gland of animals, has been reported to inhibit almost all the hallmarks of tumor such as angiogenesis, metastasis, dysregulated metabolism, immune evasion, replicative immortality, and proliferative signaling, and induces apoptosis in various tumors an in vitro and in vivo studies [191]. Furthermore, purpurogallin inhibits the anchorage-dependent/independent growth of the esophageal squamous cell carcinoma in vitro and in vivo through the inhibition of MEK1/2 and ERK1/2 signaling pathways and, moreover, induces cell cycle arrest at S and G2 phases by decreasing Cyclin A1 and cyclin B2 and also activates PARP which induces apoptosis [192]. Xie et al. reported that mahanimbine inhibits the growth, proliferation, and viability of bladder cancer cells by inducing apoptosis with an increase in the expression of Bax and a decrease in Bcl2 levels, and also causes cell cycle arrest at Go/G1 phase. Furthermore it also induced autophagic death by increasing the LC3II and p62 expressions [193]. Similarly, taurine, an important amino acid present in different tissues of the body, has been proven to have antitumor effects in breast, lung, liver, colon, and prostate cancer by suppressing proliferation, invasion, and metastasis and inducing apoptosis [194]. Some of the common natural products which are undergoing different pre-clinical research are presented in Table 1.

4. Novel Formulations of Natural Products for Chemotherapy

The major concern of traditional therapies is their poor accessibility to the tumor site; therefore, higher doses are required to achieve a desired pharmacological response, which causes adverse drug reactions. The advancement of innovative nanotechnologies in medicines can significantly improve the treatment at clinical settings by overcoming the existing limitations associated with the diagnosis and treatment of various fatal diseases. Despite a wide range of antitumor effects exhibited by natural compounds in vitro preclinical studies, various hurdles still exist in translating these promising results into in vivo or clinical trials, resulting in failure and expectations. Despite remarkable health benefits, natural products’ full clinical potential has not yet been unlocked due to low aqueous solubility, poor absorption, and lower bioavailability, and shorter retention time in the biological environment. After administration, natural agents need to interact with various physico-chemical barriers that can alter their structure and affect their antitumoral activity. Therefore, novel formulation strategies are adopted to prevent the degradation of the natural compound and their parent structure, which helps retain their chemopreventive and chemotherapeutic activities [412].
Nanobased formulations in the size range of 30–100 nm have a higher surface area and can easily pass through the membranes and be absorbed. These unique characteristics of nanoparticles make them significantly attractive because it can easily overcome the inherent poor solubility and absorption problem of many natural compounds and other chemical entities. Various nano-based formulations have been employed for drug delivery in order to achieve targeted delivery, improve aqueous solubility and bioavailability, and also enhance the retention time, thereby minimizing the adverse effects and toxicities and also protecting the drug molecules from the detrimental effects of the bio environment such as enzymatic attack, pH fluctuation, and biochemical degradation [413]. Nanoencapsulation can protect and deliver the natural agents in their natural structural form and target it to the body’s specific tissues. Various nanostructures are designed and developed for drug delivery applications such as polymeric nanoparticles, liposomes, micelles, nanogels, solid lipid nanoparticles, dendrimers, nanocapsules, nanoemulsion, metallic, and ceramic nanoparticles.
Recently, Muosa et al. prepared ellagic acid (EA) or diindolylmethane (DIM) loaded PEG-PLGA nanoparticles, tested them against pancreatic cancer SUIT2-luciferase cells, and observed that the nanoformulation of the bioactive compounds produced more significant effects than the compounds alone by inhibiting the viability, angiogenesis, and growth of the pancreatic tumor cells [414]. Similarly, β-sterol loaded PLGA nanoparticles showed better antitumoral activity against breast cancer MDA-MB-231 and MCF-7 cell lines by inhibiting the proliferation and growth of the cells [415]. Recently, Feng et al. demonstrated that sequential delivery of α-mangostin and triptolide loaded polymeric micelles can improve the permeation and therapeutic efficacy in pancreatic cancer by inactivating cancer-associated fibroblast triggered by TGF-β and, as a result, can increase the perfusion at the tumor site, induce apoptosis, and inhibit proliferation in the orthotopic model of pancreatic ductal adenocarcinoma [416]. In a recent study, PEGylated betulinic acid loaded liposomes were developed and were shown to have better tumor inhibitory effects in in vitro (HepG2 and Hela cells) and in vivo (U14 cells xenograft) cancer models [417]. Recently, Liu et al. prepared a coordination nanoassembly of luteolin and ferric ions, which increased the solubility, stability, absorption, and efficacy, and acted as a chemotherapeutic agent as well as photothermal agent, which shows that it can overcome the problem of stability and increase its therapeutic efficacy [418].
A potential clinically effective therapeutic approach is the combinatorial delivery of bioactive molecules, producing synergistic or additive effects at lower doses, and minimizing the associated toxicities. Ahmadi et al. co-encapsulated hydroxystyrol and doxorubicin in PLGA-co-Acrylic acid nanoparticles and delivered them to HT-29 colon cancer cells, which resulted in higher apoptosis and cell cycle arrest and regulated the genes expression better than the single-loaded formulations and free drugs [419]. Similarly, the liquid crystalline nanoparticle of resveratrol and pemetrexed was developed for the management of lung cancer and tested against A549 lung cancer cells and urethane induced lung cancer model in mice, and the results indicated that the cytotoxicity, cellular uptake, and antitumoral activity was significantly improved compared to the drugs alone [420]. Furthermore, paclitaxel loaded artesunate-phospholipid liposomes have been tested against MCF-7, HepG2, and A549 cell lines, and it showed better cellular internalization and synergistically enhanced the in vitro antitumoral efficacy against all the tested cell lines [421]. Moreover, Wang et al. developed a hyaluronic acid-coated PEI-PLGA nanosystem for co-delivery of gambogic acid and pTRAIL for triple-negative breast cancer (TNBC) therapy. They observed that combinatorial nanosystem significantly inhibited tumor growth and proliferation by inducing apoptosis in vitro and in vivo tumor xenograft model of TNBC [422].
Similarly, Bian et al. investigated the combinatorial anticancer effects of curcumin and sorafenib-loaded lactosylated nanoparticles against the hepatocellular carcinoma HepG2 cell line HCC tumor xenograft model. They found that the combinatorial nanosystem even reduced the smallest tumor volume and exhibited the strongest inhibitory potential with lower systemic toxicity [423]. In addition, Wang et al. developed a novel PB@MIL-100 (Fe) dual metal-organic framework loaded with artemisinin for dual photothermal/chemotherapeutic purpose and evaluated in vitro and in vivo studies and found that the d-MOFs system produced synergistic effects with lower toxicity and can be a good option to be utilized as a photothermal-chemotherapeutic nanomedicine for the cancer therapy [424].
Nanocrystals are an important and emerging carrier system for delivering therapeutics. Recently, Wang et al. synthesized hyaluronic acid-coated camptothecin-loaded nanocrystals, evaluated them against various cancer cell lines such as MDA-MB-231, MCF-7, HepG2, and IMR-90 cells, and demonstrated that HA decorated CP nanocrystals improved the cytotoxicity, specificity, anti-migration activity, and antiproliferative activity [425]. Gupta et al. conjugated/encapsulated berberine in PAMAM dendrimers and evaluated them against MCF-7 and MDA-MB468 cell lines and in vivo pharmacokinetic studies in albino rats. They found that the conjugated berberine produced more significant and improved anticancer effects than the berberine encapsulated dendrimers [426]. Ding et al. reported a novel nanogel carrier system for co-delivery of EGCG and siRNA for drug-resistant breast cancer therapy and demonstrated that the multicomponent nanogel carrier has more pronounced cytotoxic effects and showed improved anticancer activity in in vitro MDA-MB231 cell line and in vivo tumor xenograft model [427]. Nanoparticles can specifically target cancer cells and enhance the specificity and efficacy of therapeutic modalities, resulting in improved patient compliance, response, and survival. However, the nanomaterial used for drug delivery applications must be biocompatible and biodegradable so that the unloaded material can be degraded into non-toxic metabolites and removed from the body via circulation. Various nanomaterials for cancer treatment are at the preliminary stages of research, and more significant in vivo data with excellent clinical efficacy are required in order to bring it to clinical settings for commercial use.
Moreover, various new natural compounds of different structures, isolated from various plants, have been deemed as models, leads, or heads of series, and their structural alteration has resulted in compounds with pharmacological action and remarkable therapeutic potential outcomes (Figure 4). This research area, which is persistently growing and is of tremendous current interest, investigates new natural products originating from various sources. The significance of natural compounds in anticancer medication exploration could go past molecular diversity and novelty in structures (Table 2). The discovery of new natural compound structures with critical biological significance and new mechanisms of activity could likewise pioneer innovative research.
The increasing trend in natural products research shows that natural products have sustained their credibility as a viable and fertile area for drug discovery and advancement. Otherwise, there would have been a substantial therapeutic deficit in many most important clinical areas, such as cancer, cardiovascular disease, and neurodegenerative diseases. The most encouraging part is the continuous introduction of new natural product chemotypes with intriguing structures and biological activities, which are highly diverse and respond to much wider screening opportunities provided by a plethora of newly discovered targets. Natural products can provide a basic structural unit, and many lead molecules can be built upon by providing some useful input from the synthetic chemical libraries and utilizing modern computational and high throughput allied technologies.

5. Conclusions and Future Perspectives

In summary, the development of cancer is a complex process that is influenced by many factors. The present review highlights the importance of natural molecules in cancer management, since natural products provide inexhaustible sources of compounds with unique structures and new mechanisms of action. Moreover, natural compounds, either used alone or in combination, can be beneficial in the treatment and prevention of cancer. However, further studies are needed to describe the effect of natural compounds on cancer progression. In addition, it is evident that there is still much to be explored based on the natural diversity of the world, and because of technological advancements there are several new prototypes for pharmacologically active compounds appearing in screening programs, which will accelerate the exploration of natural compounds for different diseases such as cancer.
The growing incidence of cancer and limitations of conventional therapy such as higher costs, toxicities due to non-specific targeting, increased chemoresistance, and metastasis has led to tumor recurrence and posed a serious challenge to design and develop an alternative biocompatible, eco-friendly, and cost-effective strategy to combat cancer. Primary prevention is a better way to control cancer, benefiting both present and future generations, and estimates suggest that 30 to 50% of incident cancers can be prevented by translating prior knowledge of causes into effective interventions. Under this scenario, natural products are expected to revolutionize cancer treatment in the next decade due to their higher efficacy, biodegradability, and biocompatibility. Though the clinical applications of natural compounds are limited due to their poor aqueous solubility, rapid catabolism, targeting specificity, poor intestinal absorption, and reduced bioavailability, although nanotechnology-based formulations are used to overcome these limitations. Moreover, there is a lack of insight into the molecular interactions of natural compounds with various signaling molecules at the preclinical stage; for this purpose, in silico approaches such as molecular docking are needed to understand the interactions of natural products in different signaling pathways and identify their associated carcinogenesis biomarkers which can be further confirmed by many in vitro and in vivo models. Additionally, methodological flaws in various clinical studies such as smaller sample size, lack of placebo or control groups, and shorter trial duration have been observed. Thus, for many natural products, it is important to perform large-scale and well-controlled clinical trials to confirm and validate their anticancer activities, adverse effects, and safeties prior to their usage as anticancer drugs/adjuvants. Furthermore, extensive standardization based on modern approaches by evaluating their efficacy, composition, dosage-regimen, safety, quality, bioavailability, manufacturing practices, and regulatory and approval procedures need to be performed on the promising natural products to meet the quality and criteria of the international standard. Interestingly, pharmaceutical industries have vast experience and knowledge about drug development, and, therefore, combining the advantages of modern and traditional medicines can increase the development and availability of natural compounds for patient use.
A novel integrated approach for drug discovery, taking advantage of the ethnopharmacological knowledge validated and confirmed through the interdisciplinary efforts involving natural products chemistry, pharmacology, medicinal chemistry, biochemistry, cellular and molecular biology, is required to harvest the true potential of natural products. The higher biofunctionality and biodiversity of natural products make them attractive and unique from a drug discovery perspective, which needs to be further explored using modern screening techniques, which can provide various new prototypes that are pharmacologically active, thus increasing the speed of exploitation of natural compounds. Moreover, advancements in analytical techniques and computational chemistry coupled with artificial intelligence can help in the facilitation and identification of new natural active molecules for pharmacological evaluation. Further research on natural products and their limitations may lead to the discovery of new potent anticancer therapeutics with enhanced efficacy, safety, and quality for chemoprevention and chemotherapy.
In summary, this review is intended to increase scientists’ and researchers’ awareness of natural products’ diverse benefits, which can be utilized to develop safer and newer cancer treatments, as well as provide a solid foundation for future studies on natural compounds in cancer therapy. In the future, further studies are needed to investigate combinations of natural products comprising more than two natural products or combinations of existing chemotherapeutic drugs with less frequently utilized natural products. Moreover, many of these products were tested on limited cancer types, so their spectrum of activity should be expanded. Bioavailability limits the effectiveness of naturally occurring compounds. Thus, researchers must focus not only on the efficacy of the compound, which is of high interest, but also on drug delivery systems able to overcome pharmacokinetic issues, in addition to studying derivatives with a high degree of biological efficacy and availability.

Author Contributions

Conceptualization, A.N.; visualization, P.H.; methodology, A.N.: software, Y.L.; investigation, A.N.; formal analysis, J.Z.; data curation, A.N.; writing—original draft preparation, A.N.; validation, W.Z.; resources, Q.Z.; funding acquisition, Q.Z.; supervision, Q.Z. and M.Y.; project administration, Q.Z. and M.Y.; and writing—review and editing, A.N., Q.Z., and M.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China, grant number 82060719, and Jiangxi University of Chinese Medicine (Traditional Chinese Medicine Preparation Technology and Equipment Innovation Team, funding number CXTD22006).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We are thankful to Jiangxi University of Chinese Medicine, Nanchang, China for providing digital resources.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Not applicable.

References

  1. Cadoná, F.C.; Dantas, R.F.; de Mello, G.H.; Silva, F.P., Jr. Natural products targeting into cancer hallmarks: An update on caffeine, theobromine, and (+)-catechin. Crit. Rev. Food Sci. Nutr. 2021, 62, 7222–7241. [Google Scholar] [CrossRef] [PubMed]
  2. Brennan, P.; Davey-Smith, G. Identifying Novel Causes of Cancers to Enhance Cancer Prevention: New Strategies Are Needed. JNCI J. Natl. Cancer Inst. 2022, 114, 353–360. [Google Scholar] [CrossRef] [PubMed]
  3. Sung, H.; Ferlay, J.; Siegel, R.L.; Laversanne, M.; Soerjomataram, I.; Jemal, A.; Bray, F. Global cancer statistics 2020: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J. Clin. 2021, 71, 209–249. [Google Scholar] [CrossRef] [PubMed]
  4. Hassanpour, S.H.; Dehghani, M. Review of cancer from perspective of molecular. J. Cancer Res. Pract. 2017, 4, 127–129. [Google Scholar] [CrossRef]
  5. Rahman, M.M.; Sarker, M.T.; Tumpa, M.A.A.; Yamin, M.; Islam, T.; Park, M.N.; Islam, M.R.; Rauf, A.; Sharma, R.; Cavalu, S. Exploring the recent trends in perturbing the cellular signaling pathways in cancer by natural products. Front. Pharmacol. 2022, 13, 950109. [Google Scholar] [CrossRef] [PubMed]
  6. Huang, B.; Zhang, Y. Teaching an old dog new tricks: Drug discovery by repositioning natural products and their derivatives. Drug Discov. Today 2022, 27, 1936–1944. [Google Scholar] [CrossRef]
  7. Muhammad, N.; Usmani, D.; Tarique, M.; Naz, H.; Ashraf, M.; Raliya, R.; Tabrez, S.; Zughaibi, T.A.; Alsaieedi, A.; Hakeem, I.J.; et al. The role of natural products and their multitargeted approach to treat solid cancer. Cells 2022, 11, 2209. [Google Scholar] [CrossRef]
  8. Dickens, E.; Ahmed, S. Principles of cancer treatment by chemotherapy. Surgery 2018, 36, 134–138. [Google Scholar]
  9. Liu, S.; Khan, A.R.; Yang, X.; Dong, B.; Ji, J.; Zhai, G. The reversal of chemotherapy-induced multidrug resistance by nanomedicine for cancer therapy. J. Control. Release 2021, 335, 1–20. [Google Scholar] [CrossRef]
  10. Lee, S.; Rauch, J.; Kolch, W. Targeting MAPK Signaling in Cancer: Mechanisms of Drug Resistance and Sensitivity. Int. J. Mol. Sci. 2020, 21, 1102. [Google Scholar] [CrossRef] [Green Version]
  11. Bashraheel, S.S.; Domling, A.; Goda, S.K. Update on targeted cancer therapies, single or in combination, and their fine tuning for precision medicine. Biomed. Pharmacother. 2020, 125, 110009. [Google Scholar] [CrossRef]
  12. Sharifi-Rad, J.; Quispe, C.; Patra, J.K.; Singh, Y.D.; Panda, M.K.; Das, G.; Adetunji, C.O.; Michael, O.S.; Sytar, O.; Polito, L.; et al. Paclitaxel: Application in Modern Oncology and Nanomedicine-Based Cancer Therapy. Oxidative Med. Cell. Longev. 2021, 2021, 3687700. [Google Scholar] [CrossRef]
  13. Nonnekens, J.; Hoeijmakers, J.H. After surviving cancer, what about late life effects of the cure? EMBO Mol. Med. 2017, 9, 4–6. [Google Scholar] [CrossRef]
  14. Wigmore, P.M.; Mustafa, S.; El-Beltagy, M.; Lyons, L.; Umka, J.; Bennett, G. Effects of 5-FU. In Chemo Fog; Springer: New York, NY, USA, 2010; pp. 157–164. [Google Scholar]
  15. Cardona-Mendoza, A.; Olivares-Niño, G.; Díaz-Báez, D.; Lafaurie, G.I.; Perdomo, S.J. Chemopreventive and Anti-tumor Potential of Natural Products in Oral Cancer. Nutr. Cancer 2022, 74, 779–795. [Google Scholar] [CrossRef]
  16. Gao, Q.; Feng, J.; Liu, W.; Wen, C.; Wu, Y.; Liao, Q.; Zou, L.; Sui, X.; Xie, T.; Zhang, J.; et al. Opportunities and challenges for co-delivery nanomedicines based on combination of phytochemicals with chemotherapeutic drugs in cancer treatment. Adv. Drug Deliv. Rev. 2022, 188, 114445. [Google Scholar] [CrossRef]
  17. Pepper, J.W.; Findlay, C.S.; Kassen, R.; Spencer, S.L.; Maley, C.C. Synthesis: Cancer research meets evolutionary biology. Evol. Appl. 2009, 2, 62–70. [Google Scholar] [CrossRef]
  18. Yagüe, E.; Arance, A.; Kubitza, L.; O’Hare, M.; Jat, P.; Ogilvie, C.M.; Hart, I.R.; Higgins, C.F.; Raguz, S. Ability to acquire drug resistance arises early during the tumorigenesis process. Cancer Res. 2007, 67, 1130–1137. [Google Scholar]
  19. Zhu, Y.; Ouyang, Z.; Du, H.; Wang, M.; Wang, J.; Sun, H.; Kong, L.; Xu, Q.; Ma, H.; Sun, Y. New opportunities and challenges of natural products research: When target identification meets single-cell multiomics. Acta Pharm. Sin. B 2022, 12, 4011–4039. [Google Scholar] [CrossRef]
  20. Newman, D.J.; Cragg, G.M. Natural Products as Sources of New Drugs from 1981 to 2014. J. Nat. Prod. 2016, 79, 629–661. [Google Scholar] [CrossRef] [Green Version]
  21. Newman, D.J.; Cragg, G.M. Natural products as sources of new drugs over the nearly four decades from January 1981 to September 2019. J. Nat. Prod. 2020, 83, 770–803. [Google Scholar] [CrossRef] [Green Version]
  22. Patridge, E.; Gareiss, P.; Kinch, M.S.; Hoyer, D. An analysis of FDA-approved drugs: Natural products and their derivatives. Drug Discov. Today 2016, 21, 204–207. [Google Scholar] [CrossRef] [PubMed]
  23. Priya, S.; Satheeshkumar, P.K. 5-Natural Products From Plants: Recent Developments in Phytochemicals, Phytopharmaceuticals, and Plant-Based Neutraceuticals as Anticancer Agents. In Functional and Preservative Properties of Phytochemicals; Prakash, B., Ed.; Academic Press: Cambridge, MA, USA, 2020; pp. 145–163. [Google Scholar]
  24. Šudomová, M.; Berchová-Bímová, K.; Marzocco, S.; Liskova, A.; Kubatka, P.; Hassan, S.T. Berberine in human oncogenic herpesvirus infections and their linked cancers. Viruses 2021, 13, 1014. [Google Scholar] [CrossRef] [PubMed]
  25. Liskova, A.; Samec, M.; Koklesova, L.; Brockmueller, A.; Zhai, K.; Abdellatif, B.; Siddiqui, M.; Biringer, K.; Kudela, E.; Pec, M.; et al. Flavonoids as an effective sensitizer for anti-cancer therapy: Insights into multi-faceted mechanisms and applicability towards individualized patient profiles. EPMA J. 2021, 12, 155–176. [Google Scholar] [CrossRef] [PubMed]
  26. Atanasov, A.G.; Zotchev, S.B.; Dirsch, V.M.; The International Natural Product Sciences Taskforce; Supuran, C.T. Natural products in drug discovery: Advances and opportunities. Nat. Rev. Drug Discov. 2021, 20, 200–216. [Google Scholar] [CrossRef] [PubMed]
  27. Lachance, H.; Wetzel, S.; Kumar, K.; Waldmann, H. Charting, navigating, and populating natural product chemical space for drug discovery. J. Med. Chem. 2012, 55, 5989–6001. [Google Scholar] [CrossRef]
  28. Dey, P.; Kundu, A.; Kumar, A.; Gupta, M.; Lee, B.M.; Bhakta, T.; Dash, S.; Kim, H.S. Analysis of Alkaloids (Indole Alkaloids, Isoquinoline Alkaloids, Tropane Alkaloids). In Recent Advances in Natural Products Analysis; Elsevier: Amsterdam, The Netherlands, 2020; pp. 505–567. [Google Scholar]
  29. Ramos, A.C.; Peláez, R.; López, J.L.; Caballero, E.; Medarde, M.; San Feliciano, A. Heterolignanolides. Furo-and thieno-analogues of podophyllotoxin and thuriferic acid. Tetrahedron 2001, 57, 3963–3977. [Google Scholar] [CrossRef]
  30. Talib, W.H.; Daoud, S.; Mahmod, A.I.; Hamed, R.A.; Awajan, D.; Abuarab, S.F.; Odeh, L.H.; Khater, S.; Al Kury, L.T. Plants as a Source of Anticancer Agents: From Bench to Bedside. Molecules 2022, 27, 4818. [Google Scholar] [CrossRef]
  31. Găman, A.M.; Egbuna, C.; Găman, M.-A. Natural Bioactive Lead Compounds Effective Against Haematological Malignancies. In Phytochemicals as Lead Compounds for New Drug Discovery; Elsevier: Amsterdam, The Netherlands, 2020; pp. 95–115. [Google Scholar]
  32. Wang, H.; Zhang, K.; Liu, J.; Yang, J.; Tian, Y.; Yang, C.; Li, Y.; Shao, M.; Su, W.; Song, N. Curcumin regulates cancer progression: Focus on ncRNAs and molecular signaling pathways. Front. Oncol. 2021, 11, 660712. [Google Scholar] [CrossRef]
  33. Khatoon, E.; Banik, K.; Harsha, C.; Sailo, B.L.; Thakur, K.K.; Khwairakpam, A.D.; Vikkurthi, R.; Devi, T.B.; Gupta, S.C.; Kunnumakkara, A.B. Phytochemicals in Cancer Cell Chemosensitization: Current Knowledge and Future Perspectives. In Seminars in Cancer Biology; Academic Press: Cambridge, MA, USA, 2020. [Google Scholar]
  34. Bordoloi, D.; Roy, N.K.; Monisha, J.; Padmavathi, G.; Kunnumakkara, A.B. Multi-targeted agents in cancer cell chemosensitization: What we learnt from curcumin thus far. Recent Pat. Anti-Cancer Drug Discov. 2016, 11, 67–97. [Google Scholar] [CrossRef]
  35. Monisha, J.; Jaiswal, A.; Banik, K.; Choudhary, H.; Singh, A.K.; Bordoloi, D.; Kunnumakkara, A.B. Cancer Cell Chemoresistance: A Prime Obstacle in Cancer Therapy. In Cancer Cell Chemoresistance and Chemosensitization; World Scientific: Singapore, 2018; pp. 15–49. [Google Scholar]
  36. Maurya, S.K.; Shadab, G.; Siddique, H.R. Chemosensitization of Therapy Resistant Tumors: Targeting Multiple Cell Signaling Pathways by Lupeol, A Pentacyclic Triterpene. Curr. Pharm. Des. 2020, 26, 455–465. [Google Scholar] [CrossRef]
  37. Yu, J.; Zhong, B.; Chen, X. Induction of Programmed Necrosis by Phytochemicals in Colorectal Cancer. In Drug Resistance in Colorectal Cancer: Molecular Mechanisms and Therapeutic Strategies; Elsevier: Amsterdam, The Netherlands, 2020; pp. 117–133. [Google Scholar]
  38. Hsu, Y.-H.; Chen, S.-Y.; Wang, S.-Y.; Lin, J.-A.; Yen, G.-C. Pterostilbene Enhances Cytotoxicity and Chemosensitivity in Human Pancreatic Cancer Cells. Biomolecules 2020, 10, 709. [Google Scholar] [CrossRef]
  39. Datta, S.; Sinha, D. EGCG maintained Nrf2-mediated redox homeostasis and minimized etoposide resistance in lung cancer cells. J. Funct. Foods 2019, 62, 103553. [Google Scholar] [CrossRef]
  40. Tripathi, S.K.; Panda, M.; Biswal, B.K. Emerging role of plumbagin: Cytotoxic potential and pharmaceutical relevance towards cancer therapy. Food Chem. Toxicol. 2019, 125, 566–582. [Google Scholar] [CrossRef]
  41. Lan, C.-Y.; Chen, S.-Y.; Kuo, C.-W.; Lu, C.-C.; Yen, G.-C. Quercetin facilitates cell death and chemosensitivity through RAGE/PI3K/AKT/mTOR axis in human pancreatic cancer cells. J. Food Drug Anal. 2019, 27, 887–896. [Google Scholar] [CrossRef]
  42. Muthusamy, G.; Gunaseelan, S.; Prasad, N.R. Ferulic acid reverses P-glycoprotein-mediated multidrug resistance via inhibition of PI3K/Akt/NF-κB signaling pathway. J. Nutr. Biochem. 2019, 63, 62–71. [Google Scholar] [CrossRef]
  43. Irani, S. Emerging insights into the biology of metastasis: A review article. Iran. J. Basic Med. Sci. 2019, 22, 833–847. [Google Scholar]
  44. Kapinova, A.; Kubatka, P.; Liskova, A.; Baranenko, D.; Kruzliak, P.; Matta, M.; Büsselberg, D.; Malicherova, B.; Zulli, A.; Kwon, T.K.; et al. Controlling metastatic cancer: The role of phytochemicals in cell signaling. J. Cancer Res. Clin. Oncol. 2019, 145, 1087–1109. [Google Scholar] [CrossRef]
  45. Yang, P.; Jiang, Y.; Pan, Y.; Ding, X.; Rhea, P.; Ding, J.; Hawke, D.H.; Felsher, D.; Narla, G.; Lu, Z.; et al. Mistletoe extract Fraxini inhibits the proliferation of liver cancer by down-regulating c-Myc expression. Sci. Rep. 2019, 9, 6428. [Google Scholar] [CrossRef] [Green Version]
  46. Koh, Y.-C.; Ho, C.-T.; Pan, M.-H. Recent advances in cancer chemoprevention with phytochemicals. J. Food Drug Anal. 2020, 28, 14–37. [Google Scholar] [CrossRef] [Green Version]
  47. Sun, L.; Chen, B.; Jiang, R.; Li, J.; Wang, B. Resveratrol inhibits lung cancer growth by suppressing M2-like polarization of tumor associated macrophages. Cell. Immunol. 2017, 311, 86–93. [Google Scholar] [CrossRef]
  48. Loo, W.T.; Jin, L.; Chow, L.W.; Cheung, M.N.; Wang, M. Rhodiola algida improves chemotherapy-induced oral mucositis in breast cancer patients. Expert Opin. Investig. Drugs 2010, 19, S91–S100. [Google Scholar] [CrossRef] [PubMed]
  49. Galluzzi, L.; Senovilla, L.; Vacchelli, E.; Eggermont, A.; Fridman, W.H.; Galon, J.; Sautès-Fridman, C.; Tartour, E.; Zitvogel, L.; Kroemer, G. Trial watch: Dendritic cell-based interventions for cancer therapy. Oncoimmunology 2012, 1, 1111–1134. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Chang, W.T.; Lai, T.H.; Chyan, Y.J.; Yin, S.Y.; Chen, Y.H.; Wei, W.C.; Yang, N.-S. Specific medicinal plant polysaccharides effectively enhance the potency of a DC-based vaccine against mouse mammary tumor metastasis. PLoS ONE 2015, 10, e0122374. [Google Scholar] [CrossRef] [PubMed]
  51. Huang, H.F.; Zeng, Z.; Chen, M.Q. Roles of Kupffer cells in liver transplantation. Hepato-Gastroenterology 2012, 59, 1251–1257. [Google Scholar] [CrossRef] [PubMed]
  52. Fang, F.; Xiao, W.; Tian, Z. NK Cell-Based Immunotherapy for Cancer. In Seminars in Immunology; Academic Press: Cambridge, MA, USA, 2017; pp. 37–54. [Google Scholar]
  53. Wu, X.-T.; Liu, J.-Q.; Lu, X.-T.; Chen, F.-X.; Zhou, Z.-H.; Wang, T.; Zhu, S.-P.; Fei, S.-J. The enhanced effect of lupeol on the destruction of gastric cancer cells by NK cells. Int. Immunopharmacol. 2013, 16, 332–340. [Google Scholar] [CrossRef] [PubMed]
  54. Hou, D.; Wang, D.; Ma, X.; Chen, W.; Guo, S.; Guan, H. Effects of total flavonoids of sea buckthorn (Hippophae rhamnoides L.) on cytotoxicity of NK92-MI cells. Int. J. Immunopathol. Pharmacol. 2017, 30, 353–361. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Zhou, J.; Wu, J.; Chen, X.; Fortenbery, N.; Eksioglu, E.; Kodumudi, K.N.; Epling-Burnette, P.K.; Dong, J.; Djeu, J.Y.; Wei, S. Icariin and its derivative, ICT, exert anti-inflammatory, anti-tumor effects, and modulate myeloid derived suppressive cells (MDSCs) functions. Int. Immunopharmacol. 2011, 11, 890–898. [Google Scholar] [CrossRef] [Green Version]
  56. Lu, Y.-T.; Li, J.; Qi, X.; Pei, Y.-x.; Shi, W.-G.; Lin, H.-S. Effects of Shugan Jianpi Formula (疏肝健脾方) on myeloid-derived suppression cells-mediated depression breast cancer mice. Chin. J. Integr. Med. 2017, 23, 453–460. [Google Scholar] [CrossRef]
  57. Murakami, H.; Ogawara, H.; Hiroshi, H. Th1/Th2 cells in patients with multiple myeloma. Hematology 2004, 9, 41–45. [Google Scholar] [CrossRef]
  58. Wei, H.; Sun, R.; Xiao, W.; Feng, J.; Zhen, C.; Xu, X.; Tian, Z. Type two cytokines predominance of human lung cancer and its reverse by traditional Chinese medicine TTMP. Cell. Mol. Immunol. 2004, 1, 63–70. [Google Scholar]
  59. Zhang, M.-Y.; Guo, J.; Hu, X.-M.; Zhao, S.-Q.; Li, S.-L.; Wang, J. An in vivo anti-tumor effect of eckol from marine brown algae by improving the immune response. Food Funct. 2019, 10, 4361–4371. [Google Scholar] [CrossRef]
  60. Takei, M.; Tachikawa, E.; Hasegawa, H.; Lee, J.-J. Dendritic cells maturation promoted by M1 and M4, end products of steroidal ginseng saponins metabolized in digestive tracts, drive a potent Th1 polarization. Biochem. Pharmacol. 2004, 68, 441–452. [Google Scholar] [CrossRef]
  61. Wang, Y.; Zhang, Q.; Chen, Y.; Liang, C.-L.; Liu, H.; Qiu, F.; Dai, Z. Antitumor effects of immunity-enhancing traditional Chinese medicine. Biomed. Pharmacother. 2020, 121, 109570. [Google Scholar] [CrossRef]
  62. Li, Q.; Bao, J.-M.; Li, X.-L.; Zhang, T.; Shen, X.-H. Inhibiting effect of Astragalus polysaccharides on the functions of CD4+ CD25highTreg cells in the tumor microenvironment of human hepatocellular carcinoma. Chin. Med. J. 2012, 125, 786–793. [Google Scholar]
  63. Du, X.; Chen, X.; Zhao, B.; Lv, Y.; Zhang, H.; Liu, H.; Chen, Z.; Chen, Y.; Zeng, X. Astragalus polysaccharides enhance the humoral and cellular immune responses of hepatitis B surface antigen vaccination through inhibiting the expression of transforming growth factor β and the frequency of regulatory T cells. FEMS Immunol. Med. Microbiol. 2011, 63, 228–235. [Google Scholar] [CrossRef]
  64. He, X.; Li, X.; Liu, B.; Xu, L.; Zhao, H.; Lu, A. Down-regulation of Treg cells and up-regulation of TH1/TH2 cytokine ratio were induced by polysaccharide from Radix Glycyrrhizae in H22 hepatocarcinoma bearing mice. Molecules 2011, 16, 8343–8352. [Google Scholar] [CrossRef]
  65. Kasagi, S.; Kawano, S.; Kumagai, S. PD-1 and autoimmunity. Crit. Rev. Immunol. 2011, 31, 265–295. [Google Scholar]
  66. Chikuma, S.; Kanamori, M.; Mise-Omata, S.; Yoshimura, A. Suppressors of cytokine signaling: Potential immune checkpoint molecules for cancer immunotherapy. Cancer Sci. 2017, 108, 574–580. [Google Scholar] [CrossRef] [Green Version]
  67. Zhang, X.; Tong, J.; Li, Z. Qiyusanlong decoction inhibits the level of PD-1/PD-L1 in mice bearing Lewis lung carcinoma. Xi Bao Yu Fen Zi Mian Yi Xue Za Zhi Chin. J. Cell. Mol. Immunol. 2016, 32, 770–774. [Google Scholar]
  68. Lv, J.; Jia, Y.; Li, J.; Kuai, W.; Li, Y.; Guo, F.; Xu, X.; Zhao, Z.; Lv, J.; Li, Z. Gegen Qinlian decoction enhances the effect of PD-1 blockade in colorectal cancer with microsatellite stability by remodelling the gut microbiota and the tumour microenvironment. Cell Death Dis. 2019, 10, 415. [Google Scholar] [CrossRef] [Green Version]
  69. Zhao, G.-J.; Lu, Z.-Q.; Tang, L.-M.; Wu, Z.-S.; Wang, D.-W.; Zheng, J.-Y.; Qiu, Q.-M. Curcumin inhibits suppressive capacity of naturally occurring CD4+ CD25+ regulatory T cells in mice in vitro. Int. Immunopharmacol. 2012, 14, 99–106. [Google Scholar] [CrossRef] [PubMed]
  70. Lin, S.R.; Chang, C.H.; Hsu, C.F.; Tsai, M.J.; Cheng, H.; Leong, M.K.; Sung, P.J.; Chen, J.C.; Weng, C.F. Natural compounds as potential adjuvants to cancer therapy: Preclinical evidence. Br. J. Pharmacol. 2020, 177, 1409–1423. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Bishayee, A.; Sethi, G. Bioactive Natural Products in Cancer Prevention and Therapy: Progress and Promise. In Seminars in Cancer Biology; Elsevier: Amsterdam, The Netherlands, 2016; pp. 1–3. [Google Scholar]
  72. Wang, L.; Sun, J.; Gao, P.; Su, K.; Wu, H.; Li, J.; Lou, W. Wnt1-inducible signaling protein 1 regulates laryngeal squamous cell carcinoma glycolysis and chemoresistance via the YAP1/TEAD1/GLUT1 pathway. J. Cell. Physiol. 2019, 234, 15941–15950. [Google Scholar] [CrossRef]
  73. Zhang, P.; Lai, Z.-L.; Chen, H.-F.; Zhang, M.; Wang, A.; Jia, T.; Sun, W.-Q.; Zhu, X.-M.; Chen, X.-F.; Zhao, Z.; et al. Curcumin synergizes with 5-fluorouracil by impairing AMPK/ULK1-dependent autophagy, AKT activity and enhancing apoptosis in colon cancer cells with tumor growth inhibition in xenograft mice. J. Exp. Clin. Cancer Res. 2017, 36, 190. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Namkaew, J.; Jaroonwitchawan, T.; Rujanapun, N.; Saelee, J.; Noisa, P. Combined effects of curcumin and doxorubicin on cell death and cell migration of SH-SY5Y human neuroblastoma cells. Vitr. Cell. Dev. Biol.-Anim. 2018, 54, 629–639. [Google Scholar] [CrossRef]
  75. Öztürk, Y.; Günaydın, C.; Yalçın, F.; Nazıroğlu, M.; Braidy, N. Resveratrol enhances apoptotic and oxidant effects of paclitaxel through TRPM2 channel activation in DBTRG glioblastoma cells. Oxidative Med. Cell. Longev. 2019, 2019, 4619865. [Google Scholar] [CrossRef]
  76. Shen, W.; Liang, B.; Yin, J.; Li, X.; Cheng, J. Noscapine Increases the Sensitivity of Drug-Resistant Ovarian Cancer Cell Line SKOV3/DDP to Cisplatin by Regulating Cell Cycle and Activating Apoptotic Pathways. Cell Biochem. Biophys. 2015, 72, 203–213. [Google Scholar] [CrossRef]
  77. Sivalingam, K.S.; Paramasivan, P.; Weng, C.F.; Viswanadha, V.p. Neferine Potentiates the Antitumor Effect of Cisplatin in Human Lung Adenocarcinoma Cells Via a Mitochondria-Mediated Apoptosis Pathway. J. Cell. Biochem. 2017, 118, 2865–2876. [Google Scholar] [CrossRef]
  78. Wang, Y.; Lu, H.-L.; Liu, Y.-D.; Yang, L.-Y.; Jiang, Q.-K.; Zhu, X.-J.; Fan, H.-N.; Qian, Y. Cryptotanshinone sensitizes antitumor effect of paclitaxel on tongue squamous cell carcinoma growth by inhibiting the JAK/STAT3 signaling pathway. Biomed. Pharmacother. 2017, 95, 1388–1396. [Google Scholar] [CrossRef]
  79. Zhang, X.; Ni, Q.; Wang, Y.; Fan, H.-w.; Li, Y. Synergistic anticancer effects of formononetin and temozolomide on glioma C6 cells. Biol. Pharm. Bull. 2018, 8, 1194–1202. [Google Scholar] [CrossRef] [Green Version]
  80. Tseng, H.-S.; Wang, Y.-F.; Tzeng, Y.-M.; Chen, D.-R.; Liao, Y.-F.; Chiu, H.-Y.; Hsieh, W.-T. Aloe-Emodin Enhances Tamoxifen Cytotoxicity by Suppressing Ras/ERK and PI3K/mTOR in Breast Cancer Cells. Am. J. Chin. Med. 2017, 45, 337–350. [Google Scholar] [CrossRef] [PubMed]
  81. Sánchez, B.G.; Bort, A.; Mateos-Gómez, P.A.; Rodríguez-Henche, N.; Díaz-Laviada, I. Combination of the natural product capsaicin and docetaxel synergistically kills human prostate cancer cells through the metabolic regulator AMP-activated kinase. Cancer Cell Int. 2019, 19, 54. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Saikia, M.; Retnakumari, A.P.; Anwar, S.; Anto, N.P.; Mittal, R.; Shah, S.; Pillai, K.S.; Balachandran, V.S.; Peter, V.; Thomas, R.; et al. Heteronemin, a marine natural product, sensitizes acute myeloid leukemia cells towards cytarabine chemotherapy by regulating farnesylation of Ras. Oncotarget 2018, 9, 18115–18127. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Jiang, Q.-W.; Cheng, K.-J.; Mei, X.-L.; Qiu, J.-G.; Zhang, W.-J.; Xue, Y.-Q.; Qin, W.-M.; Yang, Y.; Zheng, D.-W.; Chen, Y.; et al. Synergistic anticancer effects of triptolide and celastrol, two main compounds from thunder god vine. Oncotarget 2015, 6, 32790–32804. [Google Scholar] [CrossRef] [PubMed]
  84. Huang, Y.; Wang, K.; Gu, C.; Yu, G.; Zhao, D.; Mai, W.; Zhong, Y.; Liu, S.; Nie, Y.; Yang, H. Berberine, a natural plant alkaloid, synergistically sensitizes human liver cancer cells to sorafenib. Oncol. Rep. 2018, 40, 1525–1532. [Google Scholar] [CrossRef] [Green Version]
  85. Desai, V.; Jain, A.; Shaghaghi, H.; Summer, R.; Lai, J.C.K.; Bhushan, A. Combination of biochanin A and temozolomide impairs tumor growth by modulating cell metabolism in glioblastoma multiforme. Anticancer Res. 2019, 39, 57–66. [Google Scholar] [CrossRef] [Green Version]
  86. Du Plessis-Stoman, D.; Du Preez, J.G.H.; van de Venter, M. Combination treatment with oxaliplatin and mangiferin causes increased apoptosis and downregulation of NFκB in cancer cell lines. Afr. J. Tradit. Complement. Altern. Med. 2011, 8, 177–184. [Google Scholar]
  87. Lin, M.-T.; Lin, C.-L.; Lin, T.-Y.; Cheng, C.-W.; Yang, S.-F.; Lin, C.-L.; Wu, C.-C.; Hsieh, Y.-H.; Tsai, J.-P. Synergistic effect of fisetin combined with sorafenib in human cervical cancer HeLa cells through activation of death receptor-5 mediated caspase-8/caspase-3 and the mitochondria-dependent apoptotic pathway. Tumor Biol. 2016, 37, 6987–6996. [Google Scholar] [CrossRef]
  88. Zhu, Z.; Du, S.; Ding, F.; Guo, S.; Ying, G.; Yan, Z. Ursolic acid attenuates temozolomide resistance in glioblastoma cells by downregulating O(6)-methylguanine-DNA methyltransferase (MGMT) expression. Am. J. Transl. Res. 2016, 8, 3299–3308. [Google Scholar]
  89. Neitzel, C.; Seiwert, N.; Göder, A.; Diehl, E.; Weber, C.; Nagel, G.; Stroh, S.; Rasenberger, B.; Christmann, M.; Fahrer, J. Lipoic Acid Synergizes with Antineoplastic Drugs in Colorectal Cancer by Targeting p53 for Proteasomal Degradation. Cells 2019, 8, 794. [Google Scholar] [CrossRef] [Green Version]
  90. Huang, Z.; Zhang, Y.; Li, H.; Zhou, Y.; Zhang, Q.; Chen, R.; Jin, T.; Hu, K.; Li, S.; Wang, Y.; et al. Vitamin D promotes the cisplatin sensitivity of oral squamous cell carcinoma by inhibiting LCN2-modulated NF-κB pathway activation through RPS3. Cell Death Dis. 2019, 10, 936. [Google Scholar] [CrossRef] [Green Version]
  91. Pazhang, Y.; Jaliani, H.Z.; Imani, M.; Dariushnejad, H. Synergism between NF-kappa B inhibitor, celastrol, and XIAP inhibitor, embelin, in an acute myeloid leukemia cell line, HL-60. J. Cancer Res. Ther. 2016, 12, 155. [Google Scholar]
  92. Li, X.; Zhu, F.; Jiang, J.; Sun, C.; Wang, X.; Shen, M.; Tian, R.; Shi, C.; Xu, M.; Peng, F.; et al. Synergistic antitumor activity of withaferin A combined with oxaliplatin triggers reactive oxygen species-mediated inactivation of the PI3K/AKT pathway in human pancreatic cancer cells. Cancer Lett. 2015, 357, 219–230. [Google Scholar] [CrossRef]
  93. Yang, Y.-I.; Lee, K.-T.; Park, H.-J.; Kim, T.J.; Choi, Y.S.; Shih, I.-M.; Choi, J.-H. Tectorigenin sensitizes paclitaxel-resistant human ovarian cancer cells through downregulation of the Akt and NFκB pathway. Carcinogenesis 2012, 33, 2488–2498. [Google Scholar] [CrossRef] [Green Version]
  94. Qian, J.; Xia, M.; Liu, W.; Li, L.; Yang, J.; Mei, Y.; Meng, Q.; Xie, Y. Glabridin resensitizes p-glycoprotein-overexpressing multidrug-resistant cancer cells to conventional chemotherapeutic agents. Eur. J. Pharmacol. 2019, 852, 231–243. [Google Scholar] [CrossRef]
  95. Xia, G.; Wang, H.; Song, Z.; Meng, Q.; Huang, X.; Huang, X. Gambogic acid sensitizes gemcitabine efficacy in pancreatic cancer by reducing the expression of ribonucleotide reductase subunit-M2 (RRM2). J. Exp. Clin. Cancer Res. 2017, 36, 107. [Google Scholar] [CrossRef] [Green Version]
  96. Boueroy, P.; Hahnvajanawong, C.; Boonmars, T.; Saensa-ard, S.; Wattanawongdon, W.; Kongsanthia, C.; Salao, K.; Wongwajana, S.; Anantachoke, N.; Reutrakul, V. Synergistic Effect of Forbesione From Garcinia hanburyi in Combination with 5-Fluorouracil on Cholangiocarcinoma. Asian Pac. J. Cancer Prev. 2017, 18, 3343–3351. [Google Scholar] [CrossRef]
  97. Su, J.; Zhang, F.; Li, X.; Liu, Z. Osthole promotes the suppressive effects of cisplatin on NRF2 expression to prevent drug-resistant cervical cancer progression. Biochem. Biophys. Res. Commun. 2019, 514, 510–517. [Google Scholar] [CrossRef]
  98. Bray, F.; Ferlay, J.; Soerjomataram, I.; Siegel, R.L.; Torre, L.A.; Jemal, A. Global cancer statistics 2018: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J. Clin. 2018, 68, 394–424. [Google Scholar] [CrossRef] [Green Version]
  99. Ye, W.; Sun, W.; Chen, R.; Wang, Z.; Cui, X.; Zhang, H.; Qian, S.; Zheng, Q.; Zhou, Y.; Wan, J.; et al. Pharmacokinetics in rat plasma and tissue distribution in mice of galangin determined by UHPLC–MS/MS. Acta Chromatogr. 2019, 31, 120–125. [Google Scholar] [CrossRef]
  100. Hung, J.-Y.; Yang, C.-J.; Tsai, Y.-M.; Huang, H.-W.; Huang, M.-S. Antiproliferative activity of aucubin is through cell cycle arrest and apoptosis in human non-small cell lung cancer A549 cells. Clin. Exp. Pharmacol. Physiol. 2008, 35, 995–1001. [Google Scholar] [CrossRef] [PubMed]
  101. Xu, W.-T.; Li, T.-Z.; Li, S.-M.; Wang, C.; Wang, H.; Luo, Y.-H.; Piao, X.-J.; Wang, J.-R.; Zhang, Y.; Zhang, T.; et al. Cytisine exerts anti-tumour effects on lung cancer cells by modulating reactive oxygen species-mediated signalling pathways. Artif. Cells Nanomed. Biotechnol. 2020, 48, 84–95. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Han, S.; Yang, X.; Pan, Y.; Qi, Q.; Shen, J.; Fang, H.; Ji, Z. L-securinine inhibits the proliferation of A549 lung cancer cells and promotes DKK1 promoter methylation. Oncol. Lett. 2017, 14, 4243–4248. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Hasan, A.; Haque, E.; Hameed, R.; Maier, P.N.; Irfan, S.; Kamil, M.; Nazir, A.; Mir, S.S. Hsp90 inhibitor gedunin causes apoptosis in A549 lung cancer cells by disrupting Hsp90:Beclin-1:Bcl-2 interaction and downregulating autophagy. Life Sci. 2020, 256, 118000. [Google Scholar] [CrossRef] [PubMed]
  104. Srinual, S.; Chanvorachote, P.; Pongrakhananon, V. Suppression of cancer stem-like phenotypes in NCI-H460 lung cancer cells by vanillin through an Akt-dependent pathway. Int. J. Oncol. 2017, 50, 1341–1351. [Google Scholar] [CrossRef] [Green Version]
  105. Hua, P.; Sun, M.; Zhang, G.; Zhang, Y.; Tian, X.; Li, X.; Cui, R.; Zhang, X. Cepharanthine induces apoptosis through reactive oxygen species and mitochondrial dysfunction in human non-small-cell lung cancer cells. Biochem. Biophys. Res. Commun. 2015, 460, 136–142. [Google Scholar] [CrossRef]
  106. Hutchinson, L. Breast cancer: Challenges, controversies, breakthroughs. Nat. Rev. Clin. Oncol. 2010, 7, 669–670. [Google Scholar] [CrossRef] [Green Version]
  107. Zhou, X.; Yue, G.G.-L.; Tsui, S.K.-W.; Pu, J.; Fung, K.-P.; Lau, C.B.-S. Elaborating the role of natural products on the regulation of autophagy and their potentials in breast cancer therapy. Curr. Cancer Drug Targets 2018, 18, 239–255. [Google Scholar] [CrossRef]
  108. Jin, Z.-Q.; Hao, J.; Yang, X.; He, J.-H.; Liang, J.; Yuan, J.-W.; Mao, Y.; Liu, D.; Cao, R.; Wu, X.-Z.; et al. Higenamine enhances the antitumor effects of cucurbitacin B in breast cancer by inhibiting the interaction of AKT and CDK2. Oncol. Rep. 2018, 40, 2127–2136. [Google Scholar] [CrossRef] [Green Version]
  109. Nigjeh, S.E.; Yeap, S.K.; Nordin, N.; Rahman, H.; Rosli, R. In vivo anti-tumor effects of citral on 4T1 breast cancer cells via induction of apoptosis and downregulation of aldehyde dehydrogenase activity. Molecules 2019, 24, 3241. [Google Scholar] [CrossRef]
  110. Wang, L.; Wang, G.; Yang, D.; Guo, X.; Xu, Y.; Feng, B.; Kang, J. Euphol arrests breast cancer cells at the G1 phase through the modulation of cyclin D1, p21 and p27 expression. Mol. Med. Rep. 2013, 8, 1279–1285. [Google Scholar] [CrossRef] [Green Version]
  111. Reddy, D.; Ghosh, P.; Kumavath, R. Strophanthidin Attenuates MAPK, PI3K/AKT/mTOR, and Wnt/β-Catenin Signaling Pathways in Human Cancers. Front. Oncol. 2020, 9, 1469. [Google Scholar] [CrossRef] [Green Version]
  112. Fatima, I.; El-Ayachi, I.; Taotao, L.; Lillo, M.A.; Krutilina, R.; Seagroves, T.N.; Radaszkiewicz, T.W.; Hutnan, M.; Bryja, V.; Krum, S.A.; et al. The natural compound Jatrophone interferes with Wnt/β-catenin signaling and inhibits proliferation and EMT in human triple-negative breast cancer. PLoS ONE 2017, 12, e0189864. [Google Scholar] [CrossRef] [Green Version]
  113. Deng, Y.-T.; Huang, H.-C.; Lin, J.-K. Rotenone induces apoptosis in MCF-7 human breast cancer cell-mediated ROS through JNK and p38 signaling. Mol. Carcinog. 2010, 49, 141–151. [Google Scholar] [CrossRef]
  114. Dhandayuthapani, S.; Perez, H.D.; Paroulek, A.; Chinnakkannu, P.; Kandalam, U.; Jaffe, M.; Rathinavelu, A. Bromelain-Induced Apoptosis in GI-101A Breast Cancer Cells. J. Med. Food 2012, 15, 344–349. [Google Scholar] [CrossRef]
  115. Abu, N.; Akhtar, M.N.; Yeap, S.K.; Lim, K.L.; Ho, W.Y.; Abdullah, M.P.; Ho, C.L.; Omar, A.R.; Ismail, J.; Alitheen, N.B. Flavokawain B induced cytotoxicity in two breast cancer cell lines, MCF-7 and MDA-MB231 and inhibited the metastatic potential of MDA-MB231 via the regulation of several tyrosine kinases in vitro. BMC Complement. Altern. Med. 2016, 16, 86. [Google Scholar] [CrossRef] [Green Version]
  116. Stewart, B.; Wild, C.P. World Cancer Report 2014; World Health Organization: Geneva, Switzerland, 2014. [Google Scholar]
  117. Lee, C.S.; Jang, E.-R.; Kim, Y.J.; Myung, S.C.; Kim, W.; Lee, M.W. Diarylheptanoid hirsutenone enhances apoptotic effect of TRAIL on epithelial ovarian carcinoma cell lines via activation of death receptor and mitochondrial pathway. Investig. New Drugs 2012, 30, 548–557. [Google Scholar] [CrossRef]
  118. Weaver, B.A. How Taxol/paclitaxel kills cancer cells. Mol. Biol. Cell 2014, 25, 2677–2681. [Google Scholar] [CrossRef]
  119. Zhang, H.; Jiao, Y.; Shi, C.; Song, X.; Chang, Y.; Ren, Y.; Shi, X. Berbamine suppresses cell proliferation and promotes apoptosis in ovarian cancer partially via the inhibition of Wnt/β-catenin signaling. Acta Biochim. Biophys. Sin. 2018, 50, 532–539. [Google Scholar] [CrossRef] [Green Version]
  120. Chen, T.G.; Li, L.Y.; Wei, Y.R.; Zhang, L.W. Screening, identification and activity evaluation of pancreatic lipase inhibition in Prunella vulgaris. China J. Chin. Mater. Med. 2018, 43, 4665–4671. [Google Scholar]
  121. Siegel, R.L.; Miller, K.D.; Jemal, A. Colorectal cancer mortality rates in adults aged 20 to 54 years in the United States, 1970–2014. JAMA 2017, 318, 572–574. [Google Scholar] [CrossRef] [PubMed]
  122. Sun, J.; Ding, C.; Yang, Z.; Liu, T.; Zhang, X.; Zhao, C.; Wang, J. The long non-coding RNA TUG1 indicates a poor prognosis for colorectal cancer and promotes metastasis by affecting epithelial-mesenchymal transition. J. Transl. Med. 2016, 14, 42. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Imai, H.; Sawada, K.; Sato, A.; Nishi, K.; Sasaki, T.; Takahashi, T.; Ohori, H. Complete resection of liver metastases of colorectal cancer after high efficacy bevacizumab, S-1, and CPT-11 combination chemotherapy. Gan Kagaku Ryoho Cancer Chemother. 2015, 42, 101–104. [Google Scholar]
  124. Murthy, K.C.; Jayaprakasha, G.; Patil, B.S. Obacunone and obacunone glucoside inhibit human colon cancer (SW480) cells by the induction of apoptosis. Food Chem. Toxicol. 2011, 49, 1616–1625. [Google Scholar] [CrossRef] [PubMed]
  125. Sithara, T.; Arun, K.; Syama, H.; Reshmitha, T.; Nisha, P. Morin inhibits proliferation of SW480 colorectal cancer cells by inducing apoptosis mediated by reactive oxygen species formation and uncoupling of warburg effect. Front. Pharmacol. 2017, 8, 640. [Google Scholar] [CrossRef] [Green Version]
  126. Ban, J.O.; Oh, J.H.; Hwang, B.Y.; Moon, D.C.; Jeong, H.-S.; Lee, S.; Kim, S.; Lee, H.; Kim, K.-B.; Han, S.B.; et al. Inflexinol inhibits colon cancer cell growth through inhibition of nuclear factor-κB activity via direct interaction with p50. Mol. Cancer Ther. 2009, 8, 1613–1624. [Google Scholar] [CrossRef] [Green Version]
  127. Ren, H.; Zhao, J.; Fan, D.; Wang, Z.; Zhao, T.; Li, Y.; Zhao, Y.; Adelson, D.; Hao, H. Alkaloids from nux vomica suppresses colon cancer cell growth through Wnt/β-catenin signaling pathway. Phytother. Res. 2019, 33, 1570–1578. [Google Scholar] [CrossRef]
  128. Gao, X.; Wang, Y.; Zhang, J.; Lin, L.; Yao, Q.; Xiang, G. Bergenin suppresses the growth of colorectal cancer cells by inhibiting PI3K/AKT/mTOR signaling pathway. Trop. J. Pharm. Res. 2017, 16, 2307–2313. [Google Scholar] [CrossRef]
  129. Jin, Z.; Yan, W.; Jin, H.; Ge, C.; Xu, Y. Differential effect of psoralidin in enhancing apoptosis of colon cancer cells via nuclear factor-κB and B-cell lymphoma-2/B-cell lymphoma-2-associated X protein signaling pathways. Oncol. Lett. 2016, 11, 267–272. [Google Scholar] [CrossRef] [Green Version]
  130. Gu, Y.-Y.; Chen, M.-H.; May, B.H.; Liao, X.-Z.; Liu, J.-H.; Tao, L.-T.; Man-yuen Sze, D.; Zhang, A.L.; Mo, S.-L. Matrine induces apoptosis in multiple colorectal cancer cell lines in vitro and inhibits tumour growth with minimum side effects in vivo via Bcl-2 and caspase-3. Phytomedicine 2018, 51, 214–225. [Google Scholar] [CrossRef]
  131. Su, C.-M.; Weng, Y.-S.; Kuan, L.-Y.; Chen, J.-H.; Hsu, F.-T. Suppression of PKCδ/NF-κB Signaling and Apoptosis Induction through Extrinsic/Intrinsic Pathways Are Associated Magnolol-Inhibited Tumor Progression in Colorectal Cancer In Vitro and In Vivo. Int. J. Mol. Sci. 2020, 21, 3527. [Google Scholar] [CrossRef]
  132. Mi, C.; Ma, J.; Wang, K.S.; Zuo, H.X.; Wang, Z.; Li, M.Y.; Piao, L.X.; Xu, G.H.; Li, X.; Quan, Z.S.; et al. Imperatorin suppresses proliferation and angiogenesis of human colon cancer cell by targeting HIF-1α via the mTOR/p70S6K/4E-BP1 and MAPK pathways. J. Ethnopharmacol. 2017, 203, 27–38. [Google Scholar] [CrossRef]
  133. Zhang, L.; Zheng, Y.; Deng, H.; Liang, L.; Peng, J. Aloperine induces G2/M phase cell cycle arrest and apoptosis in HCT116 human colon cancer cells. Int. J. Mol. Med. 2014, 33, 1613–1620. [Google Scholar] [CrossRef] [Green Version]
  134. McNeill, K.A. Epidemiology of brain tumors. Neurol. Clin. 2016, 34, 981–998. [Google Scholar] [CrossRef]
  135. 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]
  136. Ye, Z.-N.; Yu, M.-Y.; Kong, L.-M.; Wang, W.-H.; Yang, Y.-F.; Liu, J.-Q.; Qiu, M.-H.; Li, Y. Biflavone ginkgetin, a novel Wnt inhibitor, suppresses the growth of medulloblastoma. Nat. Prod. Bioprospect. 2015, 5, 91–97. [Google Scholar] [CrossRef] [Green Version]
  137. Cao, L.; Qu, D.; Wang, H.; Zhang, S.; Jia, C.; Shi, Z.; Wang, Z.; Zhang, J.; Ma, J. Toosendanin exerts an anti-cancer effect in glioblastoma by inducing estrogen receptor β- and p53-mediated apoptosis. Int. J. Mol. Sci. 2016, 17, 1928. [Google Scholar] [CrossRef] [Green Version]
  138. Noman, L.; Oke-Altuntas, F.; Zellagui, A.; Sahin Yaglioglu, A.; Demirtas, I.; Cardoso, S.M.; Akkal, S.; Gherraf, N.; Rhouati, S. A novel benzimidazole and other constituents with antiproliferative and antioxidant properties from Thymelaea microphylla Coss. et Dur. Nat. Prod. Res. 2017, 31, 2032–2041. [Google Scholar] [CrossRef]
  139. Schötterl, S.; Hübner, M.; Armento, A.; Veninga, V.; Wirsik, N.M.; Bernatz, S.; Lentzen, H.; Mittelbronn, M.; Naumann, U. Viscumins functionally modulate cell motility-associated gene expression. Int. J. Oncol. 2017, 50, 684–696. [Google Scholar] [CrossRef] [Green Version]
  140. Meng, X.; Li, Y.; Li, S.; Gan, R.-Y.; Li, H.-B. Natural Products for Prevention and Treatment of Chemical-Induced Liver Injuries. Compr. Rev. Food Sci. Food Saf. 2018, 17, 472–495. [Google Scholar] [CrossRef] [Green Version]
  141. Kotecha, R.; Takami, A.; Espinoza, J.L. Dietary phytochemicals and cancer chemoprevention: A review of the clinical evidence. Oncotarget 2016, 7, 52517–52529. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Zhou, Y.; Li, Y.; Zhou, T.; Zheng, J.; Li, S.; Li, H.-B. Dietary natural products for prevention and treatment of liver cancer. Nutrients 2016, 8, 156. [Google Scholar] [CrossRef]
  143. Zhao, Q.; Xue, Y.; Wang, J.-F.; Li, H.; Long, T.-T.; Li, Z.; Wang, Y.-M.; Dong, P.; Xue, C.-H. In vitro and in vivo anti-tumour activities of echinoside A and ds-echinoside A from Pearsonothuria graeffei. J. Sci. Food Agric. 2012, 92, 965–974. [Google Scholar] [CrossRef] [PubMed]
  144. Yan, C.-M.; Chai, E.-Q.; Cai, H.-Y.; Miao, G.-Y.; Ma, W. Oleuropein induces apoptosis via activation of caspases and suppression of phosphatidylinositol 3-kinase/protein kinase B pathway in HepG2 human hepatoma cell line. Mol. Med. Rep. 2015, 11, 4617–4624. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Yao, C.; Liu, B.-B.; Qian, X.-D.; Li, L.-Q.; Cao, H.-B.; Guo, Q.-S.; Zhou, G.-F. Crocin induces autophagic apoptosis in hepatocellular carcinoma by inhibiting Akt/mTOR activity. OncoTargets Ther. 2018, 11, 2017–2028. [Google Scholar] [CrossRef]
  146. Hui, F.; Qin, X.; Zhang, Q.; Li, R.; Liu, M.; Ren, T.; Zhao, M.; Zhao, Q. Alpinia oxyphylla oil induces apoptosis of hepatocellular carcinoma cells via PI3K/Akt pathway in vitro and in vivo. Biomed. Pharmacother. 2019, 109, 2365–2374. [Google Scholar] [CrossRef]
  147. García-Fernández, L.F.; Losada, A.; Alcaide, V.; Álvarez, A.M.; Cuadrado, A.; González, L.; Nakayama, K.; Nakayama, K.I.; Fernández-Sousa, J.M.; Muñoz, A.; et al. Aplidin™ induces the mitochondrial apoptotic pathway via oxidative stress-mediated JNK and p38 activation and protein kinase C δ. Oncogene 2002, 21, 7533–7544. [Google Scholar] [CrossRef] [Green Version]
  148. Ming, Y.; Zheng, Z.; Chen, L.; Zheng, G.; Liu, S.; Yu, Y.; Tong, Q. Corilagin inhibits hepatocellular carcinoma cell proliferation by inducing G2/M phase arrest. Cell Biol. Int. 2013, 37, 1046–1054. [Google Scholar] [CrossRef]
  149. Alsahafi, E.; Begg, K.; Amelio, I.; Raulf, N.; Lucarelli, P.; Sauter, T.; Tavassoli, M. Clinical update on head and neck cancer: Molecular biology and ongoing challenges. Cell Death Dis. 2019, 10, 540. [Google Scholar] [CrossRef] [Green Version]
  150. Cramer, J.D.; Burtness, B.; Le, Q.T.; Ferris, R.L. The changing therapeutic landscape of head and neck cancer. Nat. Rev. Clin. Oncol. 2019, 16, 669–683. [Google Scholar] [CrossRef]
  151. Yang, I.-H.; Shin, J.-A.; Kim, L.-H.; Kwon, K.H.; Cho, S.-D. The caspase 3-dependent apoptotic effect of pycnogenol in human oral squamous cell carcinoma HSC-3 cells. J. Clin. Biochem. Nutr. 2016, 58, 40–47. [Google Scholar] [CrossRef] [Green Version]
  152. Kwak, H.-H.; Kim, I.-R.; Kim, H.-J.; Park, B.-S.; Yu, S.-B. α-Mangostin Induces Apoptosis and Cell Cycle Arrest in Oral Squamous Cell Carcinoma Cell. Evid.-Based Complement. Altern. Med. 2016, 2016, 9060649. [Google Scholar] [CrossRef]
  153. De La Chapa, J.J.; Singha, P.K.; Lee, D.R.; Gonzales, C.B. Thymol inhibits oral squamous cell carcinoma growth via mitochondria-mediated apoptosis. J. Oral Pathol. Med. 2018, 47, 674–682. [Google Scholar] [CrossRef]
  154. Chattopadhyay, I. Role of Nutrigenetics and Nutrigenomics in Cancer Chemoprevention. In Pharmacotherapeutic Botanicals for Cancer Chemoprevention; Springer: Singapore, 2020; pp. 167–188. [Google Scholar]
  155. Katiyar, S.K. Emerging phytochemicals for the prevention and treatment of head and neck cancer. Molecules 2016, 21, 1610. [Google Scholar] [CrossRef] [Green Version]
  156. Rawla, P. Epidemiology of prostate cancer. World J. Oncol. 2019, 10, 63. [Google Scholar] [CrossRef] [Green Version]
  157. Salehi, B.; Fokou, P.V.T.; Yamthe, L.R.T.; Tali, B.T.; Adetunji, C.O.; Rahavian, A.; Mudau, F.N.; Martorell, M.; Setzer, W.N.; Rodrigues, C.F.; et al. Phytochemicals in Prostate Cancer: From Bioactive Molecules to Upcoming Therapeutic Agents. Nutrients 2019, 11, 1483. [Google Scholar] [CrossRef] [Green Version]
  158. Zaidi, S.; Gandhi, J.; Joshi, G.; Smith, N.L.; Khan, S.A. The anticancer potential of metformin on prostate cancer. Prostate Cancer Prostatic Dis. 2019, 22, 351–361. [Google Scholar] [CrossRef]
  159. Shukla, S.; Shankar, E.; Fu, P.; MacLennan, G.T.; Gupta, S. Suppression of NF-κB and NF-κB-regulated gene expression by apigenin through IκBα and IKK pathway in TRAMP mice. PLoS ONE 2015, 10, e0138710. [Google Scholar] [CrossRef] [Green Version]
  160. Zhu, K.-C.; Sun, J.-M.; Shen, J.-G.; Jin, J.-Z.; Liu, F.; Xu, X.-L.; Chen, L.; Liu, L.-T.; Lv, J.-J. Afzelin exhibits anti-cancer activity against androgen-sensitive LNCaP and androgen-independent PC-3 prostate cancer cells through the inhibition of LIM domain kinase 1. Oncol. Lett. 2015, 10, 2359–2365. [Google Scholar] [CrossRef] [Green Version]
  161. Akhtar, N.; Syed, D.N.; Khan, M.I.; Adhami, V.M.; Mirza, B.; Mukhtar, H. The pentacyclic triterpenoid, plectranthoic acid, a novel activator of AMPK induces apoptotic death in prostate cancer cells. Oncotarget 2016, 7, 3819–3831. [Google Scholar] [CrossRef] [Green Version]
  162. Tan, J.; Jiang, X.; Yin, G.; He, L.; Liu, J.; Long, Z.; Jiang, Z.; Yao, K. Anacardic acid induces cell apoptosis of prostatic cancer through autophagy by ER stress/DAPK3/Akt signaling pathway. Oncol. Rep. 2017, 38, 1373–1382. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Jeong, M.-H.; Ko, H.; Jeon, H.; Sung, G.-J.; Park, S.-Y.; Jun, W.J.; Lee, Y.-H.; Lee, J.; Lee, S.-w.; Yoon, H.-G.; et al. Delphinidin induces apoptosis via cleaved HDAC3-mediated p53 acetylation and oligomerization in prostate cancer cells. Oncotarget 2016, 7, 56767–56780. [Google Scholar] [CrossRef] [PubMed]
  164. Hu, M.; Peng, S.; He, Y.; Qin, M.; Cong, X.; Xing, Y.; Liu, M.; Yi, Z. Lycorine is a novel inhibitor of the growth and metastasis of hormone-refractory prostate cancer. Oncotarget 2015, 6, 15348–15361. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Lall, R.K.; Syed, D.N.; Khan, M.I.; Adhami, V.M.; Gong, Y.; Lucey, J.A.; Mukhtar, H. Dietary flavonoid fisetin increases abundance of high-molecular-mass hyaluronan conferring resistance to prostate oncogenesis. Carcinogenesis 2016, 37, 918–928. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Adaramoye, O.; Erguen, B.; Nitzsche, B.; Höpfner, M.; Jung, K.; Rabien, A. Punicalagin, a polyphenol from pomegranate fruit, induces growth inhibition and apoptosis in human PC-3 and LNCaP cells. Chem.-Biol. Interact. 2017, 274, 100–106. [Google Scholar] [CrossRef]
  167. Zeng, S.; Zhu, B.; Zeng, J.; Wu, W.; Jiang, C. Zeylenone represses the progress of human prostate cancer by downregulating the Wnt/β-catenin pathway. Mol. Med. Rep. 2018, 18, 5572–5578. [Google Scholar] [CrossRef] [Green Version]
  168. Núñez Selles, A.J.; Daglia, M.; Rastrelli, L. The potential role of mangiferin in cancer treatment through its immunomodulatory, anti-angiogenic, apoptopic, and gene regulatory effects. BioFactors 2016, 42, 475–491. [Google Scholar] [CrossRef]
  169. Elkady, A.I. Anethole inhibits the proliferation of human prostate cancer cells via induction of cell cycle arrest and apoptosis. Anti-Cancer Agents Med. Chem. Former. Curr. Med. Chem.-Anti-Cancer Agents 2018, 18, 216–236. [Google Scholar] [CrossRef]
  170. Xu, Y.; Zhu, J.-Y.; Lei, Z.-M.; Wan, L.-J.; Zhu, X.-W.; Ye, F.; Tong, Y.-Y. Anti-proliferative effects of paeonol on human prostate cancer cell lines DU145 and PC-3. J. Physiol. Biochem. 2017, 73, 157–165. [Google Scholar] [CrossRef]
  171. Kanwal, N.; Rasul, A.; Hussain, G.; Anwar, H.; Shah, M.A.; Sarfraz, I.; Riaz, A.; Batool, R.; Shahbaz, M.; Hussain, A.; et al. Oleandrin: A bioactive phytochemical and potential cancer killer via multiple cellular signaling pathways. Food Chem. Toxicol. 2020, 143, 111570. [Google Scholar] [CrossRef]
  172. Ramamoorthy, M.D.; Kumar, A.; Ayyavu, M.; Dhiraviam, K.N.; Ayyau, M. Reserpine induces apoptosis and cell cycle arrest in hormone independent prostate cancer cells through mitochondrial membrane potential failure. Anti-Cancer Agents Med. Chem. Former. Curr. Med. Chem.-Anti-Cancer Agents 2018, 18, 1313–1322. [Google Scholar] [CrossRef]
  173. Royston, K.J.; Tollefsbol, T.O. The epigenetic impact of cruciferous vegetables on cancer prevention. Curr. Pharmacol. Rep. 2015, 1, 46–51. [Google Scholar] [CrossRef]
  174. Doll, R.; Peto, R. The causes of cancer: Quantitative estimates of avoidable risks of cancer in the United States today. JNCI J. Natl. Cancer Inst. 1981, 66, 66–1192. [Google Scholar] [CrossRef]
  175. Kang, B.; Park, H.; Kim, B. Anticancer Activity and Underlying Mechanism of Phytochemicals against Multiple Myeloma. Int. J. Mol. Sci. 2019, 20, 2302. [Google Scholar] [CrossRef] [Green Version]
  176. Wu, Y.; Giaisi, M.; Köhler, R.; Chen, W.-M.; Krammer, P.H.; Li-Weber, M. Rocaglamide breaks TRAIL-resistance in human multiple myeloma and acute T-cell leukemia in vivo in a mouse xenogtraft model. Cancer Lett. 2017, 389, 70–77. [Google Scholar] [CrossRef]
  177. Ishii, N.; Araki, K.; Yokobori, T.; Hagiwara, K.; Gantumur, D.; Yamanaka, T.; Handa, T.; Tsukagoshi, M.; Igarashi, T.; Watanabe, A.; et al. Conophylline suppresses pancreatic cancer desmoplasia and cancer-promoting cytokines produced by cancer-associated fibroblasts. Cancer Sci. 2019, 110, 334–344. [Google Scholar] [CrossRef] [Green Version]
  178. Kim, B.-H.; Yi, E.H.; Jee, J.-G.; Jeong, A.J.; Sandoval, C.; Park, I.-C.; Baeg, G.H.; Ye, S.-K. Tubulosine selectively inhibits JAK3 signalling by binding to the ATP-binding site of the kinase of JAK3. J. Cell. Mol. Med. 2020, 24, 7427–7438. [Google Scholar] [CrossRef]
  179. Karami, A.; Hamzeloo-Moghadam, M.; Yami, A.; Barzegar, M.; Mashati, P.; Gharehbaghian, A. Antiproliferative Effect of Gaillardin from Inula oculus-christi in Human Leukemic Cells. Nutr. Cancer 2020, 72, 1043–1056. [Google Scholar] [CrossRef]
  180. Trivedi, R.; Maurya, R.P.; Mishra, D.P. Medicarpin, a legume phytoalexin sensitizes myeloid leukemia cells to TRAIL-induced apoptosis through the induction of DR5 and activation of the ROS-JNK-CHOP pathway. Cell Death Dis. 2014, 5, e1465. [Google Scholar] [CrossRef] [Green Version]
  181. Uchihara, Y.; Tago, K.; Funakoshi-Tago, M. The mechanisms of taxodione-induced apoptosis in BCR-ABL-positive leukemia cells. Nihon Yakurigaku Zasshi Folia Pharmacol. Jpn. 2019, 153, 147–154. [Google Scholar] [CrossRef]
  182. Wu, C.; Li, M.; Meng, H.; Liu, Y.; Niu, W.; Zhou, Y.; Zhao, R.; Duan, Y.; Zeng, Z.; Li, X.; et al. Analysis of status and countermeasures of cancer incidence and mortality in China. Sci. China Life Sci. 2019, 62, 640–647. [Google Scholar] [CrossRef] [PubMed]
  183. Shivamadhu, M.C.; Srinivas, B.K.; Jayarama, S.; Chandrashekaraiah, S.A. Anti-cancer and anti-angiogenic effects of partially purified lectin from Praecitrullus fistulosus fruit on in vitro and in vivo model. Biomed. Pharmacother. 2017, 96, 1299–1309. [Google Scholar] [CrossRef] [PubMed]
  184. Lee, S.R.; Park, J.Y.; Yu, J.S.; Lee, S.O.; Ryu, J.-Y.; Choi, S.-Z.; Kang, K.S.; Yamabe, N.; Kim, K.H. Odisolane, a Novel Oxolane Derivative, and Antiangiogenic Constituents from the Fruits of Mulberry (Morus alba L.). J. Agric. Food Chem. 2016, 64, 3804–3809. [Google Scholar] [CrossRef] [PubMed]
  185. Yap, V.A.; Loong, B.-J.; Ting, K.-N.; Loh, S.H.-S.; Yong, K.-T.; Low, Y.-Y.; Kam, T.-S.; Lim, K.-H. Hispidacine, an unusual 8,4′-oxyneolignan-alkaloid with vasorelaxant activity, and hispiloscine, an antiproliferative phenanthroindolizidine alkaloid, from Ficus hispida Linn. Phytochemistry 2015, 109, 96–102. [Google Scholar] [CrossRef] [PubMed]
  186. Kim, J.-E.; Kim, J.H.; Lee, Y.; Yang, H.; Heo, Y.-S.; Bode, A.M.; Lee, K.W.; Dong, Z. Bakuchiol suppresses proliferation of skin cancer cells by directly targeting Hck, Blk, and p38 MAP kinase. Oncotarget 2016, 7, 14616–14627. [Google Scholar] [CrossRef] [Green Version]
  187. Pal, H.C.; Katiyar, S.K. Cryptolepine, a plant alkaloid, inhibits the growth of non-melanoma skin cancer cells through inhibition of topoisomerase and induction of DNA damage. Molecules 2016, 21, 1758. [Google Scholar] [CrossRef] [Green Version]
  188. Xu, J.; Chen, Y.; Yang, R.; Zhou, T.; Ke, W.; Si, Y.; Yang, S.; Zhang, T.; Liu, X.; Zhang, L.; et al. Cucurbitacin B inhibits gastric cancer progression by suppressing STAT3 activity. Arch. Biochem. Biophys. 2020, 684, 108314. [Google Scholar] [CrossRef]
  189. Jiang, L.; Wang, Y.; Yin, Q.; Liu, G.; Liu, H.; Huang, Y.; Li, B. Phycocyanin: A potential drug for cancer treatment. J. Cancer 2017, 8, 3416–3429. [Google Scholar] [CrossRef] [Green Version]
  190. Liang, X.; Xu, C.; Cao, X.; Wang, W. Isovitexin Suppresses Cancer Stemness Property And Induces Apoptosis Of Osteosarcoma Cells by Disruption of The DNMT1/miR-34a/Bcl-2 Axis. Cancer Manag. Res. 2019, 11, 8923. [Google Scholar] [CrossRef] [Green Version]
  191. Talib, W.H. Melatonin and cancer hallmarks. Molecules 2018, 23, 518. [Google Scholar] [CrossRef] [Green Version]
  192. Xie, X.; Zu, X.; Liu, F.; Wang, T.; Wang, X.; Chen, H.; Liu, K.; Wang, P.; Liu, F.; Zheng, Y.; et al. Purpurogallin is a novel mitogen-activated protein kinase kinase 1/2 inhibitor that suppresses esophageal squamous cell carcinoma growth in vitro and in vivo. Mol. Carcinog. 2019, 58, 1248–1259. [Google Scholar] [CrossRef]
  193. Xie, H.; Zhang, T.; Yang, N.; Li, Z.; Liu, Y. Anticancer effects of Mahanimbine alkaloid on the human bladder cancer cells are due to the induction of G0/G1 cell cycle arrest, apoptosis and autophagy. J. BUON 2020, 25, 1166–1171. [Google Scholar]
  194. Tu, S.; Zhang, X.L.; Wan, H.F.; Xia, Y.Q.; Liu, Z.Q.; Yang, X.H.; Wan, F.S. Effect of taurine on cell proliferation and apoptosis human lung cancer A549 cells. Oncol. Lett. 2018, 15, 5473–5480. [Google Scholar] [CrossRef]
  195. Rath, B.; Hochmair, M.; Plangger, A.; Hamilton, G. Anticancer activity of fascaplysin against lung cancer cell and small cell lung cancer circulating tumor cell lines. Mar. Drugs 2018, 16, 383. [Google Scholar] [CrossRef] [Green Version]
  196. Xie, Y.-J.; Gao, W.-N.; Wu, Q.-B.; Yao, X.-J.; Jiang, Z.-B.; Wang, Y.-W.; Wang, W.-J.; Li, W.; Hussain, S.; Liu, L.; et al. Chelidonine selectively inhibits the growth of gefitinib-resistant non-small cell lung cancer cells through the EGFR-AMPK pathway. Pharmacol. Res. 2020, 159, 104934. [Google Scholar] [CrossRef]
  197. Kim, J.H.; Cho, E.B.; Lee, J.; Jung, O.; Ryu, B.J.; Kim, S.H.; Cho, J.Y.; Ryou, C.; Lee, S.Y. Emetine inhibits migration and invasion of human non-small-cell lung cancer cells via regulation of ERK and p38 signaling pathways. Chem.-Biol. Interact. 2015, 242, 25–33. [Google Scholar] [CrossRef]
  198. Yu, X.; Lin, H.; Wang, Y.; Lv, W.; Zhang, S.; Qian, Y.; Deng, X.; Feng, N.; Yu, H.; Qian, B. D-limonene exhibits antitumor activity by inducing autophagy and apoptosis in lung cancer. OncoTargets Ther. 2018, 11, 1833–1847. [Google Scholar] [CrossRef] [Green Version]
  199. Zhai, D.-D.; Supaibulwatana, K.; Zhong, J.-J. Inhibition of tumor cell proliferation and induction of apoptosis in human lung carcinoma 95-D cells by a new sesquiterpene from hairy root cultures of Artemisia annua. Phytomedicine 2010, 17, 856–861. [Google Scholar] [CrossRef]
  200. Wang, J.; Zhang, B.; Guo, Y.; Li, G.; Xie, Q.; Zhu, B.; Gao, J.; Chen, Z. Artemisinin Inhibits Tumor Lymphangiogenesis by Suppression of Vascular Endothelial Growth Factor C. Pharmacology 2008, 82, 148–155. [Google Scholar] [CrossRef]
  201. Chen, J.; Huang, X.; Tao, C.; Xiao, T.; Li, X.; Zeng, Q.; Ma, M.; Wu, Z. Artemether attenuates the progression of non-small cell lung cancer by inducing apoptosis, cell cycle arrest and promoting cellular senescence. Biol. Pharm. Bull. 2019, 42, 1720–1725. [Google Scholar] [CrossRef] [Green Version]
  202. Li, Y.; He, K.; Huang, Y.; Zheng, D.; Gao, C.; Cui, L.; Jin, Y.H. Betulin induces mitochondrial cytochrome c release associated apoptosis in human cancer cells. Mol. Carcinog. 2010, 49, 630–640. [Google Scholar] [CrossRef] [PubMed]
  203. Mertens-Talcott, S.U.; Noratto, G.D.; Li, X.; Angel-Morales, G.; Bertoldi, M.C.; Safe, S. Betulinic acid decreases ER-negative breast cancer cell growth in vitro and in vivo: Role of Sp transcription factors and microRNA-27a: ZBTB10. Mol. Carcinog. 2012, 52, 591–602. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Hsu, T.-I.; Wang, M.-C.; Chen, S.-Y.; Huang, S.-T.; Yeh, Y.-M.; Su, W.-C.; Chang, W.-C.; Hung, J.-J. Betulinic acid decreases specificity protein 1 (Sp1) level via increasing the sumoylation of sp1 to inhibit lung cancer growth. Mol. Pharmacol. 2012, 82, 1115–1128. [Google Scholar] [CrossRef]
  205. Zhu, M.; Wang, M.; Jiang, Y.; Wu, H.; Lu, G.; Shi, W.; Cong, D.; Song, S.; Liu, K.; Wang, H. Gambogic Acid Induces Apoptosis of Non-Small Cell Lung Cancer (NSCLC) Cells by Suppressing Notch Signaling. Med. Sci. Monit. Int. Med. J. Exp. Clin. Res. 2018, 24, 7146–7151. [Google Scholar] [CrossRef] [PubMed]
  206. Mei, C.; Zhou, S.; Zhu, L.; Ming, J.; Zeng, F.; Xu, R. Antitumor effects of Laminaria extract fucoxanthin on lung cancer. Mar. Drugs 2017, 15, 39. [Google Scholar] [CrossRef] [Green Version]
  207. You, C.; Sun, Y.; Zhang, S.; Tang, G.; Zhang, N.; Li, C.; Tian, X.; Ma, S.; Luo, Y.; Sun, W.; et al. Trichosanthin enhances sensitivity of non-small cell lung cancer (NSCLC) TRAIL-resistance cells. Int. J. Biol. Sci. 2018, 14, 217–227. [Google Scholar] [CrossRef] [Green Version]
  208. Ruan, J.S.; Zhou, H.; Yang, L.; Wang, L.; Jiang, Z.S.; Sun, H.; Wang, S.M. Ursolic Acid Attenuates TGF-β1-Induced Epithelial-Mesenchymal Transition in NSCLC by Targeting Integrin αVβ5/MMPs Signaling. Oncol. Res. 2019, 27, 593–600. [Google Scholar] [CrossRef]
  209. Song, Y.; Kong, L.; Sun, B.; Gao, L.; Chu, P.; Ahsan, A.; Qaed, E.; Lin, Y.; Peng, J.; Ma, X.; et al. Induction of autophagy by an oleanolic acid derivative, SZC017, promotes ROS-dependent apoptosis through Akt and JAK2/STAT3 signaling pathway in human lung cancer cells. Cell Biol. Int. 2017, 41, 1367–1378. [Google Scholar] [CrossRef]
  210. Wang, C.; Cui, C. Inhibition of lung cancer proliferation by wogonin is associated with the activation of apoptosis and generation of reactive oxygen species. Balk. Med. J. 2020, 37, 29–33. [Google Scholar] [CrossRef]
  211. Jiang, Z.-Q.; Li, M.-H.; Qin, Y.-M.; Jiang, H.-Y.; Zhang, X.; Wu, M.-H. Luteolin inhibits tumorigenesis and induces apoptosis of non-small cell lung cancer cells via regulation of MicroRNA-34a-5p. Int. J. Mol. Sci. 2018, 19, 447. [Google Scholar] [CrossRef] [Green Version]
  212. Zou, J.; Yang, Y.; Yang, Y.; Liu, X. Polydatin suppresses proliferation and metastasis of non-small cell lung cancer cells by inhibiting NLRP3 inflammasome activation via NF-κB pathway. Biomed. Pharmacother. 2018, 108, 130–136. [Google Scholar] [CrossRef]
  213. Tang, Q.; Wu, J.; Zheng, F.; Chen, Y.; Hann, S.S. Emodin increases expression of insulin-like growth factor binding protein 1 through activation of MEK/ERK/AMPKα and interaction of PPARγ and Sp1 in lung cancer. Cell. Physiol. Biochem. 2017, 41, 339–357. [Google Scholar] [CrossRef]
  214. Yao, C.-C.; Tu, Y.-R.; Jiang, J.; Ye, S.-F.; Du, H.-X.; Zhang, Y. β-elemene reverses the drug resistance of lung cancer A549/DDP cells via the mitochondrial apoptosis pathway. Oncol. Rep. 2014, 31, 2131–2138. [Google Scholar] [CrossRef]
  215. Jiang, Y.; Zhang, Y.; Luan, J.; Duan, H.; Zhang, F.; Yagasaki, K.; Zhang, G. Effects of bufalin on the proliferation of human lung cancer cells and its molecular mechanisms of action. Cytotechnology 2010, 62, 573–583. [Google Scholar] [CrossRef] [Green Version]
  216. Huang, A.-C.; Yang, M.-D.; Hsiao, Y.-T.; Lin, T.-S.; Ma, Y.-S.; Peng, S.-F.; Hsia, T.-C.; Cheng, Y.-D.; Kuo, C.-L.; Chung, J.-G. Bufalin inhibits gefitinib resistant NCI-H460 human lung cancer cell migration and invasion in vitro. J. Ethnopharmacol. 2016, 194, 1043–1050. [Google Scholar] [CrossRef]
  217. Zhao, R.; Chen, M.; Jiang, Z.; Zhao, F.; Xi, B.; Zhang, X.; Fu, H.; Zhou, K. Platycodin-D Induced Autophagy in Non-Small Cell Lung Cancer Cells via PI3K/Akt/mTOR and MAPK Signaling Pathways. J. Cancer 2015, 6, 623–631. [Google Scholar] [CrossRef] [Green Version]
  218. Zhang, J.; Wang, J.; Jiang, J.-Y.; Liu, S.-D.; Fu, K.; Liu, H.-Y. Tanshinone IIA induces cytochrome c-mediated caspase cascade apoptosis in A549 human lung cancer cells via the JNK pathway. Int. J. Oncol. 2014, 45, 683–690. [Google Scholar] [CrossRef] [Green Version]
  219. Mou, H.; Zheng, Y.; Zhao, P.; Bao, H.; Fang, W.; Xu, N. Celastrol induces apoptosis in non-small-cell lung cancer A549 cells through activation of mitochondria-and Fas/FasL-mediated pathways. Toxicol. Vitr. 2011, 25, 1027–1032. [Google Scholar] [CrossRef]
  220. Piantino, C.B.; Salvadori, F.A.; Ayres, P.P.; Kato, R.B.; Srougi, V.; Leite, K.R.; Srougi, M. An evaluation of the anti-neoplastic activity of curcumin in prostate cancer cell lines. Int. Braz. J. Urol. 2009, 35, 354–361. [Google Scholar] [CrossRef] [Green Version]
  221. Wang, C.; Song, X.; Shang, M.; Zou, W.; Zhang, M.; Wei, H.; Shao, H. Curcumin exerts cytotoxicity dependent on reactive oxygen species accumulation in non-small-cell lung cancer cells. Future Oncol. 2019, 15, 1243–1253. [Google Scholar] [CrossRef]
  222. Wu, C.; Zhuang, Y.; Jiang, S.; Tian, F.; Teng, Y.; Chen, X.; Zheng, P.; Liu, S.; Zhou, J.; Wu, J.; et al. Cinnamaldehyde induces apoptosis and reverses epithelial-mesenchymal transition through inhibition of Wnt/β-catenin pathway in non-small cell lung cancer. Int. J. Biochem. Cell Biol. 2017, 84, 58–74. [Google Scholar] [CrossRef] [PubMed]
  223. Lu, X.; Liu, W.; Wu, J.; Li, M.; Wang, J.; Wu, J.; Luo, C. A polysaccharide fraction of adlay seed (Coix lachryma-jobi L.) induces apoptosis in human non-small cell lung cancer A549 cells. Biochem. Biophys. Res. Commun. 2013, 430, 846–851. [Google Scholar] [CrossRef] [PubMed]
  224. Hu, C.; Jiang, R.; Cheng, Z.; Lu, Y.; Gu, L.; Li, H.; Li, L.; Gao, Q.; Chen, M.; Zhang, X. Ophiopogonin-B suppresses epithelial-mesenchymal transition in human lung adenocarcinoma cells via the Linc00668/miR-432-5p/EMT axis. J. Cancer 2019, 10, 2849–2856. [Google Scholar] [CrossRef] [PubMed]
  225. Zhong, L.-R.; Zheng, J.; Sun, Q.; Wei, K.; Hu, Y. Radix Tetrastigma hemsleyani flavone inhibits proliferation, migration, and invasion of human lung carcinoma A549 cells. OncoTargets Ther. 2016, 9, 635–641. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  226. Chou, P.-Y.; Huang, G.-J.; Pan, C.-H.; Chien, Y.-C.; Chen, Y.-Y.; Wu, C.-H.; Sheu, M.-J.; Cheng, H.-C. Trilinolein inhibits proliferation of human non-small cell lung carcinoma A549 through the modulation of pi3k/akt pathway. Am. J. Chin. Med. 2011, 39, 803–815. [Google Scholar] [CrossRef]
  227. Hsu, H.-Y.; Lin, T.-Y.; Lu, M.-K.; Leng, P.-J.; Tsao, S.-M.; Wu, Y.-C. Fucoidan induces Toll-like receptor 4-regulated reactive oxygen species and promotes endoplasmic reticulum stress-mediated apoptosis in lung cancer. Sci. Rep. 2017, 7, srep44990. [Google Scholar] [CrossRef] [Green Version]
  228. Xia, S.; Zhang, X.; Li, C.; Guan, H. Oridonin inhibits breast cancer growth and metastasis through blocking the Notch signaling. Saudi Pharm. J. 2017, 25, 638–643. [Google Scholar] [CrossRef]
  229. Guo, W.-Q.; Chen, Y.-G.; Shi, R.-Z.; He, K.; Wang, J.-F.; Shao, J.-H.; Wan, J.-B.; Gao, J.-L. 20(S)-Protopanaxdiol Suppresses the Abnormal Granule-Monocyte Differentiation of Hematopoietic Stem Cells in 4T1 Breast Cancer-Bearing Mouse. Evid.-Based Complement. Altern. Med. 2020, 2020, 8747023. [Google Scholar] [CrossRef] [Green Version]
  230. Jiang, J.; Jedinak, A.; Sliva, D. Ganodermanontriol (GDNT) exerts its effect on growth and invasiveness of breast cancer cells through the down-regulation of CDC20 and uPA. Biochem. Biophys. Res. Commun. 2011, 415, 325–329. [Google Scholar] [CrossRef]
  231. Xiaoping, C.; Yan, C.; Shuibing, L.; Youguo, C.; Jianyun, L.; Lanping, L. Free radical scavenging of Ganoderma lucidum polysaccharides and its effect on antioxidant enzymes and immunity activities in cervical carcinoma rats. Carbohydr. Polym. 2009, 77, 389–393. [Google Scholar] [CrossRef]
  232. Zhao, L.; Dong, Y.; Chen, G.; Hu, Q. Extraction, purification, characterization and antitumor activity of polysaccharides from Ganoderma lucidum. Carbohydr. Polym. 2010, 80, 783–789. [Google Scholar] [CrossRef]
  233. Yang, R.; Hanwell, H.; Zhang, J.; Tsao, R.; Meckling, K.A. Antiproliferative Activity of Pomiferin in Normal (MCF-10A) and Transformed (MCF-7) Breast Epithelial Cells. J. Agric. Food Chem. 2011, 59, 13328–13336. [Google Scholar] [CrossRef]
  234. Dong, Y.; Yin, S.; Li, J.; Jiang, C.; Ye, M.; Hu, H. Bufadienolide compounds sensitize human breast cancer cells to TRAIL-induced apoptosis via inhibition of STAT3/Mcl-1 pathway. Apoptosis 2011, 16, 394–403. [Google Scholar] [CrossRef]
  235. Yue, G.G.-L.; Xie, S.; Lee, J.K.-M.; Kwok, H.-F.; Gao, S.; Nian, Y.; Wu, X.-X.; Wong, C.-K.; Qiu, M.-H.; Lau, C.B.-S. New potential beneficial effects of actein, a triterpene glycoside isolated from Cimicifuga species, in breast cancer treatment. Sci. Rep. 2016, 6, 35263. [Google Scholar] [CrossRef] [Green Version]
  236. Zhou, R.; Xu, L.; Ye, M.; Liao, M.; Du, H.; Chen, H. Formononetin inhibits migration and invasion of MDA-MB-231 and 4T1 breast cancer cells by suppressing MMP-2 and MMP-9 through PI3K/AKT signaling pathways. Horm. Metab. Res. 2014, 46, 753–760. [Google Scholar] [CrossRef] [Green Version]
  237. Li, W.; Song, K.; Wang, S.; Zhang, C.; Zhuang, M.; Wang, Y.; Liu, T. Anti-tumor potential of astragalus polysaccharides on breast cancer cell line mediated by macrophage activation. Mater. Sci. Eng. C 2019, 98, 685–695. [Google Scholar] [CrossRef]
  238. Cao, C.; Huang, W.; Zhang, N.; Wu, F.; Xu, T.; Pan, X.; Peng, C.; Han, B. Narciclasine induces autophagy-dependent apoptosis in triple-negative breast cancer cells by regulating the AMPK-ULK1 axis. Cell Prolif. 2018, 51, e12518. [Google Scholar] [CrossRef]
  239. Kong, Y.; Li, F.; Nian, Y.; Zhou, Z.; Yang, R.; Qiu, M.-H.; Chen, C. KHF16 is a Leading Structure from Cimicifuga foetida that Suppresses Breast Cancer Partially by Inhibiting the NF-κB Signaling Pathway. Theranostics 2016, 6, 875–886. [Google Scholar] [CrossRef] [Green Version]
  240. Hien, T.T.; Kim, H.G.; Han, E.H.; Kang, K.W.; Jeong, H.G. Molecular mechanism of suppression of MDR1 by puerarin from Pueraria lobata via NF-κB pathway and cAMP-responsive element transcriptional activity-dependent up-regulation of AMP-activated protein kinase in breast cancer MCF-7/adr cells. Mol. Nutr. Food Res. 2010, 54, 918–928. [Google Scholar] [CrossRef]
  241. Moon, A.; Kim, E.-S.; Jeong, C.-S. Genipin, a constituent of Gardenia jasminoides Ellis, induces apoptosis and inhibits invasion in MDA-MB-231 breast cancer cells. Oncol. Rep. 2012, 27, 567–572. [Google Scholar] [CrossRef] [Green Version]
  242. Du, J.; Wang, X.-F.; Zhou, Q.-M.; Zhang, T.-L.; Lu, Y.-Y.; Zhang, H.; Su, S.-B. Evodiamine induces apoptosis and inhibits metastasis in MDA-MB-231 human breast cancer cells in vitro and in vivo. Oncol. Rep. 2013, 30, 685–694. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  243. Wang, H.; Chen, X.; Li, T.; Xu, J.; Ma, Y. A myrsinol diterpene isolated from a traditional herbal medicine, LANGDU reverses multidrug resistance in breast cancer cells. J. Ethnopharmacol. 2016, 194, 1–5. [Google Scholar] [CrossRef] [PubMed]
  244. Mohsenikia, M.; Alizadeh, A.M.; Khodayari, S.; Khodayari, H.; Kouhpayeh, S.A.; Karimi, A.; Zamani, M.; Azizian, S.; Mohagheghi, M.A. The protective and therapeutic effects of alpha-solanine on mice breast cancer. Eur. J. Pharmacol. 2013, 718, 1–9. [Google Scholar] [CrossRef] [PubMed]
  245. Fang, Z.-Y.; Ren, Y.-D.; Du, S.-Y.; Zhang, M.; Wang, Y.-S.; Fang, L.; Zhang, H. Melosuavine I, an apoptosis-inducing bisindole alkaloid from Melodinus suaveolens. Fitoterapia 2019, 133, 175–179. [Google Scholar] [CrossRef] [PubMed]
  246. Wang, L.; Peng, Y.; Shi, K.; Wang, H.; Lu, J.; Li, Y.; Ma, C. Osthole inhibits proliferation of human breast cancer cells by inducing cell cycle arrest and apoptosis. J. Biomed. Res. 2015, 29, 132–138. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  247. Yan, Z.; Liu, G.; Liang, M.; Xu, Y. Ophiopogonin D inhibits cell proliferation and induces apoptosis of human laryngocarcinoma through downregulation of cyclin B1 and MMP-9 and upregulation of p38-MAPK signaling. Oncol. Lett. 2019, 17, 1877–1882. [Google Scholar] [CrossRef] [Green Version]
  248. Kim, S.-J.; Kim, A.K. Anti-breast cancer activity of Fine Black ginseng (Panax ginseng Meyer) and ginsenoside Rg5. J. Ginseng Res. 2015, 39, 125–134. [Google Scholar] [CrossRef] [Green Version]
  249. Huang, C.; Chen, X.; Guo, B.; Huang, W.; Shen, T.; Sun, X.; Xiao, P.; Zhou, Q. Induction of apoptosis by Icariside II through extrinsic and intrinsic signaling pathways in human breast cancer MCF7 cells. Biosci. Biotechnol. Biochem. 2012, 76, 1322–1328. [Google Scholar] [CrossRef]
  250. Zhao, H.; Yang, Z.; Wang, X.; Zhang, X.; Wang, M.; Wang, Y.; Mei, Q.; Wang, Z. Triptolide inhibits ovarian cancer cell invasion by repression of matrix metalloproteinase 7 and 19 and upregulation of E-cadherin. Exp. Mol. Med. 2012, 44, 633–641. [Google Scholar] [CrossRef] [Green Version]
  251. Wang, J.-H.; Nao, J.-F.; Zhang, M.; He, P. 20(s)-ginsenoside Rg3 promotes apoptosis in human ovarian cancer HO-8910 cells through PI3K/Akt and XIAP pathways. Tumor Biol. 2014, 35, 11985–11994. [Google Scholar] [CrossRef]
  252. Woo, J.-H.; Ahn, J.-H.; Jang, D.S.; Lee, K.-T.; Choi, J.-H. Effect of kumatakenin isolated from cloves on the apoptosis of cancer cells and the alternative activation of tumor-associated macrophages. J. Agric. Food Chem. 2017, 65, 7893–7899. [Google Scholar] [CrossRef]
  253. Hua, F.; Li, C.-H.; Chen, X.-G.; Liu, X.-P. Daidzein exerts anticancer activity towards SKOV3 human ovarian cancer cells by inducing apoptosis and cell cycle arrest, and inhibiting the Raf/MEK/ERK cascade. Int. J. Mol. Med. 2018, 41, 3485–3492. [Google Scholar] [CrossRef] [Green Version]
  254. Chang, H.-L.; Su, J.-H.; Yeh, Y.-T.; Lee, Y.-C.; Chen, H.-M.; Wu, Y.-C.; Yuan, S.-S.F. Protoapigenone, a novel flavonoid, inhibits ovarian cancer cell growth in vitro and in vivo. Cancer Lett. 2008, 267, 85–95. [Google Scholar] [CrossRef]
  255. Li, J.; Jiang, K.; Zhao, F. Icariin regulates the proliferation and apoptosis of human ovarian cancer cells through microRNA-21 by targeting PTEN, RECK and Bcl-2. Oncol. Rep. 2015, 33, 2829–2836. [Google Scholar] [CrossRef] [Green Version]
  256. Gao, L.; Chen, M.; Ouyang, Y.; Li, R.; Zhang, X.; Gao, X.; Lin, S.; Wang, X. Icaritin induces ovarian cancer cell apoptosis through activation of p53 and inhibition of Akt/mTOR pathway. Life Sci. 2018, 202, 188–194. [Google Scholar] [CrossRef]
  257. Bae, H.; Song, G.; Lim, W. Stigmasterol Causes Ovarian Cancer Cell Apoptosis by Inducing Endoplasmic Reticulum and Mitochondrial Dysfunction. Pharmaceutics 2020, 12, 488. [Google Scholar] [CrossRef]
  258. Jiang, L.; Cao, X.-C.; Cao, J.-G.; Liu, F.; Quan, M.-F.; Sheng, X.-F.; Ren, K.-Q. Casticin induces ovarian cancer cell apoptosis by repressing FoxM1 through the activation of FOXO3a. Oncol. Lett. 2013, 5, 1605–1610. [Google Scholar] [CrossRef] [Green Version]
  259. Zhang, Y.-Y.; Zhang, F.; Zhang, Y.-S.; Thakur, K.; Zhang, J.-G.; Liu, Y.; Kan, H.; Wei, Z.-J. Mechanism of juglone-induced cell cycle arrest and apoptosis in Ishikawa Human endometrial cancer cells. J. Agric. Food Chem. 2019, 67, 7378–7389. [Google Scholar] [CrossRef]
  260. Chen, J.; Li, Z.; Chen, A.Y.; Ye, X.; Luo, H.; Rankin, G.O.; Chen, Y.C. Inhibitory effect of baicalin and baicalein on ovarian cancer cells. Int. J. Mol. Sci. 2013, 14, 6012–6025. [Google Scholar] [CrossRef] [Green Version]
  261. Du, Y.; Feng, J.; Wang, R.; Zhang, H.; Liu, J. Effects of flavonoids from Potamogeton crispus L. on proliferation, migration, and invasion of human ovarian cancer cells. PLoS ONE 2015, 10, e0130685. [Google Scholar] [CrossRef] [Green Version]
  262. Greenshields, A.L.; Shepherd, T.G.; Hoskin, D.W. Contribution of reactive oxygen species to ovarian cancer cell growth arrest and killing by the anti-malarial drug artesunate. Mol. Carcinog. 2017, 56, 75–93. [Google Scholar] [CrossRef] [PubMed]
  263. Huang, K.; Li, L.A.; Meng, Y.G.; You, Y.Q.; Fu, X.Y.; Song, L. Arctigenin promotes apoptosis in ovarian cancer cells via the iNOS/NO/STAT3/survivin signalling. Basic Clin. Pharmacol. Toxicol. 2014, 115, 507–511. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  264. Zhang, F.; Zhang, Y.-Y.; Sun, Y.-S.; Ma, R.-H.; Thakur, K.; Zhang, J.-G.; Wei, Z.-J. Asparanin A from Asparagus officinalis L. Induces G0/G1 Cell Cycle Arrest and Apoptosis in Human Endometrial Carcinoma Ishikawa Cells via Mitochondrial and PI3K/AKT Signaling Pathways. J. Agric. Food Chem. 2020, 68, 213–224. [Google Scholar] [CrossRef] [PubMed]
  265. Jiang, G.; Liu, J.; Ren, B.; Zhang, L.; Owusu, L.; Liu, L.; Zhang, J.; Tang, Y.; Li, W. Anti-tumor and chemosensitization effects of Cryptotanshinone extracted from Salvia miltiorrhiza Bge. on ovarian cancer cells in vitro. J. Ethnopharmacol. 2017, 205, 33–40. [Google Scholar] [CrossRef] [PubMed]
  266. Yang, Z.; Li, C.; Wang, X.; Zhai, C.; Yi, Z.; Wang, L.; Liu, B.; Du, B.; Wu, H.; Guo, X.; et al. Dauricine induces apoptosis, inhibits proliferation and invasion through inhibiting NF-κB signaling pathway in colon cancer cells. J. Cell. Physiol. 2010, 225, 266–275. [Google Scholar] [CrossRef]
  267. Xia, S.; Miao, Y.; Liu, S. Withaferin A induces apoptosis by ROS-dependent mitochondrial dysfunction in human colorectal cancer cells. Biochem. Biophys. Res. Commun. 2018, 503, 2363–2369. [Google Scholar] [CrossRef]
  268. De Almeida, G.C.; Oliveira, L.F.S.; Predes, D.; Fokoue, H.H.; Kuster, R.M.; Oliveira, F.L.; Mendes, F.A.; Abreu, J.G. Piperine suppresses the Wnt/β-catenin pathway and has anti-cancer effects on colorectal cancer cells. Sci. Rep. 2020, 10, 11681. [Google Scholar] [CrossRef]
  269. Ray, P.; Guha, D.; Chakraborty, J.; Banerjee, S.; Adhikary, A.; Chakraborty, S.; Das, T.; Sa, G. Crocetin exploits p53-induced death domain (PIDD) and FAS-associated death domain (FADD) proteins to induce apoptosis in colorectal cancer. Sci. Rep. 2016, 6, 32979. [Google Scholar] [CrossRef]
  270. Lee, S.K.; Nam, K.-A.; Heo, Y.-H. Cytotoxic activity and G2/M cell cycle arrest mediated by antofine, a phenanthroindolizidine alkaloid isolated from Cynanchum paniculatum. Planta Med. 2003, 69, 21–25. [Google Scholar] [CrossRef]
  271. Sithara, T.; Dhanya, B.P.; Arun, K.B.; Sini, S.; Dan, M.; Kokkuvayil Vasu, R.; Nisha, P. Zerumbone, a cyclic sesquiterpene from Zingiber zerumbet induces apoptosis, cell cycle arrest, and antimigratory effects in SW480 colorectal cancer cells. J. Agric. Food Chem. 2018, 66, 602–612. [Google Scholar] [CrossRef]
  272. Xiong, Y.; Xiong, Y.-J.; Liu, D.-Y.; Shen, R.-R. Pancratistatin Inhibits the Growth of Colorectal Cancer Cells by Inducing Apoptosis, Autophagy, and G2/M Cell Cycle Arrest. Med. Sci. Monit. 2019, 25, 6015–6022. [Google Scholar] [CrossRef]
  273. Dong, G.-Z.; Shim, A.-R.; Hyeon, J.S.; Lee, H.J.; Ryu, J.-H. Inhibition of Wnt/β-catenin pathway by dehydrocostus lactone and costunolide in colon cancer cells. Phytother. Res. 2015, 29, 680–686. [Google Scholar] [CrossRef]
  274. Zhu, P.; Wu, Y.; Yang, A.; Fu, X.; Mao, M.; Liu, Z. Catalpol suppressed proliferation, growth and invasion of CT26 colon cancer by inhibiting inflammation and tumor angiogenesis. Biomed. Pharmacother. 2017, 95, 68–76. [Google Scholar] [CrossRef]
  275. Go, H.; Hwang, H.-J.; Nam, T.-J. A glycoprotein from Laminaria japonica induces apoptosis in HT-29 colon cancer cells. Toxicol. Vitr. 2010, 24, 1546–1553. [Google Scholar] [CrossRef]
  276. Han, Y.-H.; Kee, J.-Y.; Hong, S.-H. Rosmarinic acid activates AMPK to inhibit metastasis of colorectal cancer. Front. Pharmacol. 2018, 9, 68. [Google Scholar] [CrossRef]
  277. Murthy, K.N.C.; Jayaprakasha, G.K.; Patil, B.S. The natural alkaloid berberine targets multiple pathways to induce cell death in cultured human colon cancer cells. Eur. J. Pharmacol. 2012, 688, 14–21. [Google Scholar] [CrossRef]
  278. Han, M.H.; Kim, G.-Y.; Yoo, Y.H.; Choi, Y.H. Sanguinarine induces apoptosis in human colorectal cancer HCT-116 cells through ROS-mediated Egr-1 activation and mitochondrial dysfunction. Toxicol. Lett. 2013, 220, 157–166. [Google Scholar] [CrossRef]
  279. Sichaem, J.; Surapinit, S.; Siripong, P.; Khumkratok, S.; Jong-aramruang, J.; Tip-pyang, S. Two new cytotoxic isomeric indole alkaloids from the roots of Nauclea orientalis. Fitoterapia 2010, 81, 830–833. [Google Scholar] [CrossRef]
  280. Ji, S.; Tang, S.; Li, K.; Li, Z.; Liang, W.; Qiao, X.; Wang, Q.; Yu, S.; Ye, M. Licoricidin inhibits the growth of SW480 human colorectal adenocarcinoma cells in vitro and in vivo by inducing cycle arrest, apoptosis and autophagy. Toxicol. Appl. Pharmacol. 2017, 326, 25–33. [Google Scholar] [CrossRef]
  281. Jayameena, P.; Sivakumarik, A.K.; Rajesh, S. Rutin: A potential anticancer drug against human colon cancer (Hct116) cells. Int. J. Biol. Pharm. Allied Sci. 2018, 7, 1731–1745. [Google Scholar]
  282. Dou, J.; Wang, Z.; Ma, L.; Peng, B.; Mao, K.; Li, C.; Su, M.; Zhou, C.; Peng, G. Baicalein and baicalin inhibit colon cancer using two distinct fashions of apoptosis and senescence. Oncotarget 2018, 9, 20089–20102. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  283. Tabana, Y.M.; Hassan, L.E.A.; Ahamed, M.B.K.; Dahham, S.S.; Iqbal, M.A.; Saeed, M.A.A.; Khan, S.S.; Sandai, D.; Majid, A.S.A.; Oon, C.E.; et al. Scopoletin, an active principle of tree tobacco (Nicotiana glauca) inhibits human tumor vascularization in xenograft models and modulates ERK1, VEGF-A, and FGF-2 in computer model. Microvasc. Res. 2016, 107, 17–33. [Google Scholar] [CrossRef] [PubMed]
  284. Lee, J.-C.; Won, S.-J.; Chao, C.-L.; Wu, F.-L.; Liu, H.-S.; Ling, P.; Lin, C.-N.; Su, C.-L. Morusin induces apoptosis and suppresses NF-κB activity in human colorectal cancer HT-29 cells. Biochem. Biophys. Res. Commun. 2008, 372, 236–242. [Google Scholar] [CrossRef] [PubMed]
  285. Saud, S.M.; Li, W.; Gray, Z.; Matter, M.S.; Colburn, N.H.; Young, M.R.; Kim, Y.S. Diallyl disulfide (DADS), a constituent of garlic, inactivates NF-κB and prevents colitis-induced colorectal cancer by inhibiting GSK-3β. Cancer Prev. Res. 2016, 9, 607–615. [Google Scholar] [CrossRef]
  286. Son, Y.; An, Y.; Jung, J.; Shin, S.; Park, I.; Gwak, J.; Ju, B.G.; Chung, Y.-H.; Na, M.; Oh, S. Protopine isolated from Nandina domestica induces apoptosis and autophagy in colon cancer cells by stabilizing p53. Phytother. Res. 2019, 33, 1689–1696. [Google Scholar] [CrossRef]
  287. 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, e72882. [Google Scholar] [CrossRef]
  288. Chowdhury, F.A.; Hossain, M.K.; Mostofa, A.G.M.; Akbor, M.M.; Bin Sayeed, M.S. Therapeutic potential of thymoquinone in glioblastoma treatment: Targeting major gliomagenesis signaling pathways. BioMed Res. Int. 2018, 2018, 4010629. [Google Scholar] [CrossRef] [Green Version]
  289. Yin, J.; Wang, C.; Mody, A.; Bao, L.; Hung, S.-H.; Svoronos, S.A.; Tseng, Y. The Effect of Z-Ligustilide on the Mobility of Human Glioblastoma T98G Cells. PLoS ONE 2013, 8, e66598. [Google Scholar] [CrossRef]
  290. Wu, N.; Wu, G.-c.; Hu, R.; Li, M.; Feng, H. Ginsenoside Rh2 inhibits glioma cell proliferation by targeting microRNA-128. Acta Pharmacol. Sin. 2011, 32, 345–353. [Google Scholar] [CrossRef] [Green Version]
  291. Yang, Y.; Jia, G.; Wang, Q.; Wang, R.; Deng, D.; Xue, L.; Shao, N.; Zhang, Y.; Xia, X.; Zhi, F. Tubeimoside-1 induces glioma apoptosis through regulation of Bax/Bcl-2 and the ROS/Cytochrome C/Caspase-3 pathway. OncoTargets Ther. 2015, 8, 303–311. [Google Scholar] [CrossRef] [Green Version]
  292. Pudełek, M.; Catapano, J.; Kochanowski, P.; Mrowiec, K.; Janik-Olchawa, N.; Czyż, J.; Ryszawy, D. Therapeutic potential of monoterpene α-thujone, the main compound of Thuja occidentalis L. essential oil, against malignant glioblastoma multiforme cells in vitro. Fitoterapia 2019, 134, 172–181. [Google Scholar] [CrossRef]
  293. Zhang, L.; Wang, H.; Zhu, J.; Xu, J.; Ding, K. Mollugin induces tumor cell apoptosis and autophagy via the PI3K/AKT/mTOR/p70S6K and ERK signaling pathways. Biochem. Biophys. Res. Commun. 2014, 450, 247–254. [Google Scholar] [CrossRef]
  294. Sales, L.; Pezuk, J.A.; Borges, K.S.; Brassesco, M.S.; Scrideli, C.A.; Tone, L.G.; dos Santos, M.H.; Ionta, M.; de Oliveira, J.C. Anticancer activity of 7-epiclusianone, a benzophenone from Garcinia brasiliensis, in glioblastoma. BMC Complement. Altern. Med. 2015, 15, 393. [Google Scholar] [CrossRef] [Green Version]
  295. Wang, K.; Fu, X.-T.; Li, Y.; Hou, Y.-J.; Yang, M.-F.; Sun, J.-Y.; Yi, S.-Y.; Fan, C.-D.; Fu, X.-Y.; Zhai, J.; et al. Induction of S-Phase Arrest in Human Glioma Cells by Selenocysteine, a Natural Selenium-Containing Agent Via Triggering Reactive Oxygen Species-Mediated DNA Damage and Modulating MAPKs and AKT Pathways. Neurochem. Res. 2016, 41, 1439–1447. [Google Scholar] [CrossRef]
  296. Khan, M.; Yi, F.; Rasul, A.; Li, T.; Wang, N.; Gao, H.; Gao, R.; Ma, T. Alantolactone induces apoptosis in glioblastoma cells via GSH depletion, ROS generation, and mitochondrial dysfunction. IUBMB Life 2012, 64, 783–794. [Google Scholar] [CrossRef]
  297. Wang, C.-N.; Shiao, Y.-J.; Lin, Y.-L.; Chen, C.-F. Nepalolide A inhibits the expression of inducible nitric oxide synthase by modulating the degradation of IκB-α and IκB-β in C6 glioma cells and rat primary astrocytes. Br. J. Pharmacol. 1999, 128, 345–356. [Google Scholar] [CrossRef] [Green Version]
  298. Zhou, L.; Tang, H.; Wang, F.; Ou, S.; Wu, T.; Fang, Y.; Xu, J.; Guo, K. Cyclovirobuxine D inhibits cell proliferation and migration and induces apoptosis in human glioblastoma multiforme and low-grade glioma. Oncol. Rep. 2020, 43, 807–816. [Google Scholar] [CrossRef]
  299. Guimarães, L.P.T.P.; da Graça Rocha, G.; Queiroz, R.M.; Martins, C.A.; Takiya, C.M.; Gattass, C.R. Pomolic acid induces apoptosis and inhibits multidrug resistance protein MRP1 and migration in glioblastoma cells. Oncol. Rep. 2017, 38, 2525–2534. [Google Scholar] [CrossRef] [Green Version]
  300. Wang, Z.-S.; Luo, P.; Dai, S.-H.; Liu, Z.-B.; Zheng, X.-R.; Chen, T. Salvianolic acid B induces apoptosis in human glioma U87 cells through p38-mediated ROS generation. Cell. Mol. Neurobiol. 2013, 33, 921–928. [Google Scholar] [CrossRef]
  301. Fujiwara, Y.; Shiraishi, D.; Yoshitomi, M.; Ikeda, T.; Mizuta, H.; Takeya, M.; Komohara, Y. Soyasapogenols contained in soybeans suppress tumour progression by regulating macrophage differentiation into the protumoural phenotype. J. Funct. Foods 2015, 19, 594–605. [Google Scholar] [CrossRef]
  302. Giacomelli, C.; Daniele, S.; Natali, L.; Iofrida, C.; Flamini, G.; Braca, A.; Trincavelli, M.L.; Martini, C. Carnosol controls the human glioblastoma stemness features through the epithelial-mesenchymal transition modulation and the induction of cancer stem cell apoptosis. Sci. Rep. 2017, 7, 15174. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  303. Giacomelli, C.; Natali, L.; Trincavelli, M.L.; Daniele, S.; Bertoli, A.; Flamini, G.; Braca, A.; Martini, C. New insights into the anticancer activity of carnosol: p53 reactivation in the U87MG human glioblastoma cell line. Int. J. Biochem. Cell Biol. 2016, 74, 95–108. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  304. Lowe, H.I.C.; Toyang, N.J.; Watson, C.T.; Ayeah, K.N.; Bryant, J. HLBT-100: A highly potent anti-cancer flavanone from Tillandsia recurvata (L.) L. Cancer Cell Int. 2017, 17, 38. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  305. Chen, Z.; Pan, X.; Georgakilas, A.G.; Chen, P.; Hu, H.; Yang, Y.; Tian, S.; Xia, L.; Zhang, J.; Cai, X.; et al. Tetramethylpyrazine (TMP) protects cerebral neurocytes and inhibits glioma by down regulating chemokine receptor CXCR4 expression. Cancer Lett. 2013, 336, 281–289. [Google Scholar] [CrossRef] [PubMed]
  306. Chen, H.-B.; Zhou, L.-Z.; Mei, L.; Shi, X.-J.; Wang, X.-S.; Li, Q.-L.; Huang, L. Gambogenic acid-induced time- and dose-dependent growth inhibition and apoptosis involving Akt pathway inactivation in U251 glioblastoma cells. J. Nat. Med. 2012, 66, 62–69. [Google Scholar] [CrossRef] [PubMed]
  307. Clark, P.A.; Bhattacharya, S.; Elmayan, A.; Darjatmoko, S.R.; Thuro, B.A.; Yan, M.B.; van Ginkel, P.R.; Polans, A.S.; Kuo, J.S. Resveratrol targeting of AKT and p53 in glioblastoma and glioblastoma stem-like cells to suppress growth and infiltration. J. Neurosurg. 2017, 126, 1448–1460. [Google Scholar] [CrossRef] [Green Version]
  308. Liu, B.; Gao, Y.-Q.; Wang, X.-M.; Wang, Y.-C.; Fu, L.-Q. Germacrone inhibits the proliferation of glioma cells by promoting apoptosis and inducing cell cycle arrest. Mol. Med. Rep. 2014, 10, 1046–1050. [Google Scholar] [CrossRef] [Green Version]
  309. Chen, L.; Jiang, Z.; Ma, H.; Ning, L.; Chen, H.; Li, L.; Qi, H. Volatile Oil of Acori Graminei Rhizoma-Induced Apoptosis and Autophagy are dependent on p53 Status in Human Glioma Cells. Sci. Rep. 2016, 6, 21148. [Google Scholar] [CrossRef]
  310. Miao, J.; Jiang, Y.; Wang, D.; Zhou, J.; Fan, C.; Jiao, F.; Liu, B.; Zhang, J.; Wang, Y.; Zhang, Q. Trichosanthin suppresses the proliferation of glioma cells by inhibiting LGR5 expression and the Wnt/β-catenin signaling pathway. Oncol. Rep. 2015, 34, 2845–2852. [Google Scholar] [CrossRef] [Green Version]
  311. Nan, Y.-N.; Zhu, J.-Y.; Tan, Y.; Zhang, Q.; Jia, W.; Hua, Q. Staurosporine induced apoptosis rapidly downregulates TDP-43 in glioma cells. Asian Pac. J. Cancer Prev. 2014, 15, 3575–3579. [Google Scholar] [CrossRef] [Green Version]
  312. Guerram, M.; Jiang, Z.-Z.; Sun, L.; Zhu, X.; Zhang, L.-Y. Antineoplastic effects of deoxypodophyllotoxin, a potent cytotoxic agent of plant origin, on glioblastoma U-87 MG and SF126 cells. Pharmacol. Rep. 2015, 67, 245–252. [Google Scholar] [CrossRef]
  313. Wang, Y.; Tang, H.; Zhang, Y.; Li, J.; Li, B.; Gao, Z.; Wang, X.; Cheng, G.; Fei, Z. Saponin B, a novel cytostatic compound purified from Anemone taipaiensis, induces apoptosis in a human glioblastoma cell line. Int. J. Mol. Med. 2013, 32, 1077–1084. [Google Scholar] [CrossRef] [Green Version]
  314. Shu, G.; Mi, X.; Cai, J.; Zhang, X.; Yin, W.; Yang, X.; Li, Y.; Chen, L.; Deng, X. Brucine, an alkaloid from seeds of Strychnos nux-vomica Linn., represses hepatocellular carcinoma cell migration and metastasis: The role of hypoxia inducible factor 1 pathway. Toxicol. Lett. 2013, 222, 91–101. [Google Scholar] [CrossRef]
  315. Hsueh, K.-C.; Lin, C.-L.; Tung, J.-N.; Yang, S.-F.; Hsieh, Y.-H. Nimbolide induced apoptosis by activating ERK-mediated inhibition of c-IAP1 expression in human hepatocellular carcinoma cells. Environ. Toxicol. 2018, 33, 913–922. [Google Scholar] [CrossRef]
  316. Yu, X.; Wang, Y.; Tao, S.; Sun, S. Geniposide plays anti-tumor effects by down-regulation of microRNA-224 in HepG2 and Huh7 cell lines. Exp. Mol. Pathol. 2020, 112, 104349. [Google Scholar] [CrossRef]
  317. Xie, X.; Zhu, H.; Yang, H.; Huang, W.; Wu, Y.; Wang, Y.; Luo, Y.; Wang, D.; Shao, G. Solamargine triggers hepatoma cell death through apoptosis. Oncol. Lett. 2015, 10, 168–174. [Google Scholar] [CrossRef] [Green Version]
  318. Chen, W.; Hou, J.; Yin, Y.; Jang, J.; Zheng, Z.; Fan, H.; Zou, G. α-Bisabolol induces dose- and time-dependent apoptosis in HepG2 cells via a Fas- and mitochondrial-related pathway, involves p53 and NFκB. Biochem. Pharmacol. 2010, 80, 247–254. [Google Scholar] [CrossRef]
  319. Wang, X.; Sun, D.; Tai, J.; Wang, L. Ganoderic acid A inhibits proliferation and invasion, and promotes apoptosis in human hepatocellular carcinoma cells. Mol. Med. Rep. 2017, 16, 3894–3900. [Google Scholar] [CrossRef] [Green Version]
  320. Law, B.Y.K.; Mok, S.W.F.; Chan, W.K.; Xu, S.W.; Wu, A.G.; Yao, X.J.; Wang, J.R.; Liu, L.; Wong, V.K.W. Hernandezine, a novel AMPK activator induces autophagic cell death in drug-resistant cancers. Oncotarget 2016, 7, 8090–8104. [Google Scholar] [CrossRef] [Green Version]
  321. Li, J.; Wang, H.; Wang, L.; Tan, R.; Zhu, M.; Zhong, X.; Zhang, Y.; Chen, B.; Wang, L. Decursin inhibits the growth of HepG2 hepatocellular carcinoma cells via Hippo/YAP signaling pathway. Phytother. Res. 2018, 32, 2456–2465. [Google Scholar] [CrossRef]
  322. Rodenak-Kladniew, B.; Castro, A.; Stärkel, P.; de Saeger, C.; de Bravo, M.G.; Crespo, R. Linalool induces cell cycle arrest and apoptosis in HepG2 cells through oxidative stress generation and modulation of Ras/MAPK and Akt/mTOR pathways. Life Sci. 2018, 199, 48–59. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  323. Wu, C.-M.; Yang, C.-W.; Lee, Y.-Z.; Chuang, T.-H.; Wu, P.-L.; Chao, Y.-S.; Lee, S.-J. Tylophorine arrests carcinoma cells at G1 phase by downregulating cyclin A2 expression. Biochem. Biophys. Res. Commun. 2009, 386, 140–145. [Google Scholar] [CrossRef] [PubMed]
  324. Ma, Z.-J.; Lu, L.; Yang, J.-J.; Wang, X.-X.; Su, G.; Wang, Z.-l.; Chen, G.-h.; Sun, H.-m.; Wang, M.-y.; Yang, Y. Lariciresinol induces apoptosis in HepG2 cells via mitochondrial-mediated apoptosis pathway. Eur. J. Pharmacol. 2018, 821, 1–10. [Google Scholar] [CrossRef] [PubMed]
  325. Li, N.-N.; Meng, X.-S.; Bao, Y.-R.; Wang, S.; Li, T.-J. Evidence for the involvement of COX-2/VEGF and PTEN/Pl3K/AKT pathway the mechanism of oroxin B treated liver cancer. Pharmacogn. Mag. 2018, 14, 207–213. [Google Scholar] [PubMed]
  326. Zhang, Z.; Liu, T.; Yu, M.; Li, K.; Li, W. The plant alkaloid tetrandrine inhibits metastasis via autophagy-dependent Wnt/β-catenin and metastatic tumor antigen 1 signaling in human liver cancer cells. J. Exp. Clin. Cancer Res. 2018, 37, 7. [Google Scholar] [CrossRef] [Green Version]
  327. Thirusangu, P.; Vigneshwaran, V.; Avin, B.V.; Rakesh, H.; Vikas, H.M.; Prabhakar, B.T. Scutellarein antagonizes the tumorigenesis by modulating cytokine VEGF mediated neoangiogenesis and DFF-40 actuated nucleosomal degradation. Biochem. Biophys. Res. Commun. 2017, 484, 85–92. [Google Scholar] [CrossRef]
  328. Zhu, Y.; Pan, Y.; Zhang, G.; Wu, Y.; Zhong, W.; Chu, C.; Qian, Y.; Zhu, G. Chelerythrine inhibits human hepatocellular carcinoma metastasis in vitro. Biol. Pharm. Bull. 2018, 41, 36–46. [Google Scholar] [CrossRef] [Green Version]
  329. Hu, Y.-W.; Liu, C.-Y.; Du, C.-M.; Zhang, J.; Wu, W.-Q.; Gu, Z.-L. Induction of apoptosis in human hepatocarcinoma SMMC-7721 cells in vitro by flavonoids from Astragalus complanatus. J. Ethnopharmacol. 2009, 123, 293–301. [Google Scholar] [CrossRef]
  330. Wu, L.; Li, J.; Liu, T.; Li, S.; Feng, J.; Yu, Q.; Zhang, J.; Chen, J.; Zhou, Y.; Ji, J.; et al. Quercetin shows anti-tumor effect in hepatocellular carcinoma LM3 cells by abrogating JAK2/STAT3 signaling pathway. Cancer Med. 2019, 8, 4806–4820. [Google Scholar] [CrossRef]
  331. Chan, K.T.; Meng, F.Y.; Li, Q.; Ho, C.Y.; Lam, T.S.; To, Y.; Lee, W.H.; Li, M.; Chu, K.H.; Toh, M. Cucurbitacin B induces apoptosis and S phase cell cycle arrest in BEL-7402 human hepatocellular carcinoma cells and is effective via oral administration. Cancer Lett. 2010, 294, 118–124. [Google Scholar] [CrossRef]
  332. Lin, Z.-Y.; Wu, C.-C.; Chuang, Y.-H.; Chuang, W.-L. Anti-cancer mechanisms of clinically acceptable colchicine concentrations on hepatocellular carcinoma. Life Sci. 2013, 93, 323–328. [Google Scholar] [CrossRef]
  333. Vijayalakshmi, A.; Sindhu, G. Umbelliferone arrest cell cycle at G0/G1 phase and induces apoptosis in human oral carcinoma (KB) cells possibly via oxidative DNA damage. Biomed. Pharmacother. 2017, 92, 661–671. [Google Scholar] [CrossRef]
  334. Si, L.; Zheng, L.; Xu, L.; Yin, L.; Han, X.; Qi, Y.; Xu, Y.; Wang, C.; Peng, J. Dioscin suppresses human laryngeal cancer cells growth via induction of cell-cycle arrest and MAPK-mediated mitochondrial-derived apoptosis and inhibition of tumor invasion. Eur. J. Pharmacol. 2016, 774, 105–117. [Google Scholar] [CrossRef]
  335. Huang, T.-T.; Liu, F.-G.; Wei, C.-F.; Lu, C.-C.; Chen, C.-C.; Lin, H.-C.; Ojcius, D.M.; Lai, H.-C. Activation of Multiple Apoptotic Pathways in Human Nasopharyngeal Carcinoma Cells by the Prenylated Isoflavone, Osajin. PLoS ONE 2011, 6, e18308. [Google Scholar] [CrossRef]
  336. Zhang, G.; Xu, Y.; Zhou, H.-F. Esculetin Inhibits Proliferation, Invasion, and Migration of Laryngeal Cancer In Vitro and In Vivo by Inhibiting Janus Kinas (JAK)-Signal Transducer and Activator of Transcription-3 (STAT3) Activation. Med. Sci. Monit. 2019, 25, 7853–7863. [Google Scholar] [CrossRef]
  337. Cheng, M.-F.; Lin, C.-S.; Chen, Y.-H.; Sung, P.-J.; Lin, S.-R.; Tong, Y.-W.; Weng, C.-F. Inhibitory growth of oral squamous cell carcinoma cancer via bacterial prodigiosin. Mar. Drugs 2017, 15, 224. [Google Scholar] [CrossRef] [Green Version]
  338. Deng, Q.; Yu, X.; Xiao, L.; Hu, Z.; Luo, X.; Tao, Y.; Yang, L.; Liu, X.; Chen, H.; Ding, Z.; et al. Neoalbaconol induces energy depletion and multiple cell death in cancer cells by targeting PDK1-PI3-K/Akt signaling pathway. Cell Death Dis. 2013, 4, e804. [Google Scholar] [CrossRef] [Green Version]
  339. Kowshik, J.; Nivetha, R.; Ranjani, S.; Venkatesan, P.; Selvamuthukumar, S.; Veeravarmal, V.; Nagini, S. Astaxanthin inhibits hallmarks of cancer by targeting the PI3K/NF-κΒ/STAT3 signalling axis in oral squamous cell carcinoma models. IUBMB Life 2019, 71, 1595–1610. [Google Scholar] [CrossRef]
  340. Yeh, C.-M.; Hsieh, M.-J.; Yang, J.-S.; Yang, S.-F.; Chuang, Y.-T.; Su, S.-C.; Liang, M.-Y.; Chen, M.-K.; Lin, C.-W. Geraniin inhibits oral cancer cell migration by suppressing matrix metalloproteinase-2 activation through the FAK/Src and ERK pathways. Environ. Toxicol. 2019, 34, 1085–1093. [Google Scholar] [CrossRef]
  341. He, J.; Wei, W.; Yang, Q.; Wang, Y. Phillygenin Exerts In Vitro and In Vivo Antitumor Effects in Drug-Resistant Human Esophageal Cancer Cells by Inducing Mitochondrial-Mediated Apoptosis, ROS Generation, and Inhibition of the Nuclear Factor kappa B NF-κB Signalling Pathway. Med. Sci. Monit. 2019, 25, 739–745. [Google Scholar] [CrossRef]
  342. Wu, G.; Chen, G.; Zhou, J.; Zhu, H.; Chu, J.; Zhang, F. Liriodenine enhances radiosensitivity in esophageal cancer ECA-109 cells by inducing apoptosis and G2/M arrest. Oncol. Lett. 2018, 16, 5020–5026. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  343. Lin, C.-W.; Bai, L.-Y.; Su, J.-H.; Chiu, C.-F.; Lin, W.-Y.; Huang, W.-T.; Shih, M.-C.; Huang, Y.-T.; Hu, J.-L.; Weng, J.-R. Ilimaquinone Induces Apoptosis and Autophagy in Human Oral Squamous Cell Carcinoma Cells. Biomedicines 2020, 8, 296. [Google Scholar] [CrossRef] [PubMed]
  344. Rizo, W.F.; Ferreira, L.E.; Colnaghi, V.; Martins, J.S.; Franchi, L.P.; Takahashi, C.S.; Beleboni, R.O.; Marins, M.; Pereira, P.S.; Fachin, A.L. Cytotoxicity and genotoxicity of coronaridine from Tabernaemontana catharinensis A. DC in a human laryngeal epithelial carcinoma cell line (Hep-2). Genet. Mol. Biol. 2013, 36, 105–110. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  345. Herrmann, R.; Skaf, J.; Roller, J.; Polednik, C.; Holzgrabe, U.; Schmidt, M. Anticancer effects of NSC-631570 (Ukrain) in head and neck cancer cells: In vitro analysis of growth, invasion, angiogenesis and gene expression. Oncol. Rep. 2020, 43, 282–295. [Google Scholar] [CrossRef] [PubMed]
  346. Chinni, S.R.; Li, Y.; Upadhyay, S.; Koppolu, P.K.; Sarkar, F.H. Indole-3-carbinol (I3C) induced cell growth inhibition, G1 cell cycle arrest and apoptosis in prostate cancer cells. Oncogene 2001, 20, 2927–2936. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  347. Banerjee, S.; Li, Y.; Wang, Z.; Sarkar, F.H. Multi-targeted therapy of cancer by genistein. Cancer Lett. 2008, 269, 226–242. [Google Scholar] [CrossRef] [Green Version]
  348. Chiyomaru, T.; Yamamura, S.; Fukuhara, S.; Yoshino, H.; Kinoshita, T.; Majid, S.; Saini, S.; Chang, I.; Tanaka, Y.; Enokida, H.; et al. Genistein inhibits prostate cancer cell growth by targeting miR-34a and oncogenic HOTAIR. PLoS ONE 2013, 8, e70372. [Google Scholar] [CrossRef]
  349. Bell, C.; Hawthorne, S. Ellagic acid, pomegranate and prostate cancer—A mini review. J. Pharm. Pharmacol. 2008, 60, 139–144. [Google Scholar] [CrossRef] [Green Version]
  350. Eskandari, E.; Heidarian, E.; Amini, S.; Saffari-Chaleshtori, J. Evaluating the effects of ellagic acid on pSTAT3, pAKT, and pERK1/2 signaling pathways in prostate cancer PC3 cells. J. Cancer Res. Ther. 2016, 12, 1266–1271. [Google Scholar]
  351. Lee, H.-S.; Safe, S.; Lee, S.-O. Inactivation of the orphan nuclear receptor NR4A1 contributes to apoptosis induction by fangchinoline in pancreatic cancer cells. Toxicol. Appl. Pharmacol. 2017, 332, 32–39. [Google Scholar] [CrossRef]
  352. Zhou, J.; Feng, J.-H.; Fang, L. A novel monoterpenoid indole alkaloid with anticancer activity from Melodinus khasianus. Bioorg. Med. Chem. Lett. 2017, 27, 893–896. [Google Scholar] [CrossRef]
  353. Dilshara, M.G.; Jayasooriya, R.G.P.T.; Choi, Y.H.; Kim, G.-Y. Camptothecin induces c-Myc- and Sp1-mediated hTERT expression in LNCaP cells: Involvement of reactive oxygen species and PI3K/Akt. Food Chem. Toxicol. 2019, 127, 53–60. [Google Scholar] [CrossRef]
  354. Forestier-Román, I.S.; López-Rivas, A.; Sánchez-Vázquez, M.M.; Rohena-Rivera, K.; Nieves-Burgos, G.; Ortiz-Zuazaga, H.; Torres-Ramos, C.A.; Martínez-Ferrer, M. Andrographolide induces DNA damage in prostate cancer cells. Oncotarget 2019, 10, 1085–1101. [Google Scholar] [CrossRef] [Green Version]
  355. Nie, C.; Zhou, J.; Qin, X.; Shi, X.; Zeng, Q.; Liu, J.; Yan, S.; Zhang, L. Diosgenin-induced autophagy and apoptosis in a human prostate cancer cell line. Mol. Med. Rep. 2016, 14, 4349–4359. [Google Scholar] [CrossRef] [Green Version]
  356. Agarwal, S.; Amin, K.S.; Jagadeesh, S.; Baishay, G.; Rao, P.G.; Barua, N.C.; Bhattacharya, S.; Banerjee, P.P. Mahanine restores RASSF1A expression by down-regulating DNMT1 and DNMT3B in prostate cancer cells. Mol. Cancer 2013, 12, 99. [Google Scholar] [CrossRef]
  357. Shang, X.J.; Yao, G.; Ge, J.P.; Sun, Y.; Teng, W.H.; Huang, Y.F. Procyanidin Induces Apoptosis and Necrosis of Prostate Cancer Cell Line PC-3 in a Mitochondrion-Dependent Manner. J. Androl. 2009, 30, 122–126. [Google Scholar] [CrossRef]
  358. Bommareddy, A.; McGlynn, D.; Lewis, M.; Lockus, L.; Seward, J.; Hong, K.L.; VanWert, A.L.; Dwivedi, C. Akt/survivin pathway inhibition enhances the apoptotic cell death-induced by alpha-santalol in human prostate cancer cells. Fitoterapia 2020, 143, 104552. [Google Scholar] [CrossRef]
  359. Jiang, C.; Masood, M.; Rasul, A.; Wei, W.; Wang, Y.; Ali, M.; Mustaqeem, M.; Li, J.; Li, X. Altholactone Inhibits NF-κB and STAT3 Activation and Induces Reactive Oxygen Species-Mediated Apoptosis in Prostate Cancer DU145 Cells. Molecules 2017, 22, 240. [Google Scholar] [CrossRef] [Green Version]
  360. Liew, S.Y.; Looi, C.Y.; Paydar, M.; Cheah, F.K.; Leong, K.H.; Wong, W.F.; Mustafa, M.R.; Litaudon, M.; Awang, K. Subditine, a New Monoterpenoid Indole Alkaloid from Bark of Nauclea subdita (Korth.) Steud. Induces Apoptosis in Human Prostate Cancer Cells. PLoS ONE 2014, 9, e87286. [Google Scholar] [CrossRef]
  361. Yu, Z.; Chen, Y.; Liang, C. Eriocalyxin B induces apoptosis and autophagy involving akt/mammalian target of rapamycin (mTOR) pathway in prostate cancer cells. Med. Sci. Monit. Int. Med. J. Exp. Clin. Res. 2019, 25, 8534–8543. [Google Scholar] [CrossRef]
  362. Wang, Y.; Tsai, M.-L.; Chiou, L.-Y.; Ho, C.-T.; Pan, M.-H. Antitumor Activity of Garcinol in Human Prostate Cancer Cells and Xenograft Mice. J. Agric. Food Chem. 2015, 63, 9047–9052. [Google Scholar] [CrossRef] [PubMed]
  363. Ghasemi, S.; Lorigooini, Z.; Wibowo, J.; Amini-khoei, H. Tricin isolated from Allium atroviolaceum potentiated the effect of docetaxel on PC3 cell proliferation: Role of miR-21. Nat. Prod. Res. 2019, 33, 1828–1831. [Google Scholar] [CrossRef] [PubMed]
  364. Xu, F.; Li, Q.; Wang, Z.; Cao, X. Sinomenine inhibits proliferation, migration, invasion and promotes apoptosis of prostate cancer cells by regulation of miR-23a. Biomed. Pharmacother. 2019, 112, 108592. [Google Scholar] [CrossRef] [PubMed]
  365. Wu, X.-H.; Zhou, H.-J.; Lee, J. Dihydroartemisinin inhibits angiogenesis induced by multiple myeloma RPMI8226 cells under hypoxic conditions via downregulation of vascular endothelial growth factor expression and suppression of vascular endothelial growth factor secretion. Anti-Cancer Drugs 2006, 17, 839–848. [Google Scholar] [CrossRef] [PubMed]
  366. Liao, N.-C.; Shih, Y.-L.; Chou, J.-S.; Chen, K.-W.; Chen, Y.-L.; Lee, M.-H.; Peng, S.-F.; Leu, S.-J.; Chung, J.-G. Cardamonin induces cell cycle arrest, apoptosis and alters apoptosis associated gene expression in WEHI-3 mouse leukemia cells. Am. J. Chin. Med. 2019, 47, 635–656. [Google Scholar] [CrossRef]
  367. Martins, C.; Doran, C.; Silva, I.C.; Miranda, C.; Rueff, J.; Rodrigues, A.S. Myristicin from nutmeg induces apoptosis via the mitochondrial pathway and down regulates genes of the DNA damage response pathways in human leukaemia K562 cells. Chem.-Biol. Interact. 2014, 218, 1–9. [Google Scholar] [CrossRef]
  368. Wang, Y.; Zhu, X.-F.; Xiao, Z.-J.; Wang, H.-H.; Zhou, J.-M.; Mei, Y.-P.; Deng, R.; Jiang, W.-Q.; Liu, Z.-C. Inducement effect of Meisoindigo on apoptosis of leukemia cell line HL-60 and its mechanism. Ai Zheng Chin. J. Cancer 2005, 24, 1464–1468. [Google Scholar]
  369. Kluza, J.; Mazinghien, R.; Degardin, K.; Lansiaux, A.; Bailly, C. Induction of apoptosis by the plant alkaloid sampangine in human HL-60 leukemia cells is mediated by reactive oxygen species. Eur. J. Pharmacol. 2005, 525, 32–40. [Google Scholar] [CrossRef]
  370. Saeed, M.; Jacob, S.; Sandjo, L.P.; Sugimoto, Y.; Khalid, H.E.; Opatz, T.; Thines, E.; Efferth, T. Cytotoxicity of the Sesquiterpene Lactones Neoambrosin and Damsin from Ambrosia maritima Against Multidrug-Resistant Cancer Cells. Front. Pharmacol. 2015, 6, 267. [Google Scholar] [CrossRef]
  371. Sepporta, M.V.; Mazza, T.; Morozzi, G.; Fabiani, R. Pinoresinol Inhibits Proliferation and Induces Differentiation on Human HL60 Leukemia Cells. Nutr. Cancer 2013, 65, 1208–1218. [Google Scholar] [CrossRef]
  372. Gaboriaud-Kolar, N.; Myrianthopoulos, V.; Vougogiannopoulou, K.; Gerolymatos, P.; Horne, D.A.; Jove, R.; Mikros, E.; Nam, S.; Skaltsounis, A.-L. Natural-based indirubins display potent cytotoxicity toward wild-type and t315i-resistant leukemia cell lines. J. Nat. Prod. 2016, 79, 2464–2471. [Google Scholar] [CrossRef]
  373. Hoffmann, R.; von Schwarzenberg, K.; López-Antón, N.; Rudy, A.; Wanner, G.; Dirsch, V.M.; Vollmar, A.M. Helenalin bypasses Bcl-2-mediated cell death resistance by inhibiting NF-κB and promoting reactive oxygen species generation. Biochem. Pharmacol. 2011, 82, 453–463. [Google Scholar] [CrossRef] [Green Version]
  374. Karmahapatra, S.; Kientz, C.; Shetty, S.; Yalowich, J.C.; Rakotondraibe, L.H. Capsicodendrin from Cinnamosma fragrans Exhibits Antiproliferative and Cytotoxic Activity in Human Leukemia Cells: Modulation by Glutathione. J. Nat. Prod. 2018, 81, 625–629. [Google Scholar] [CrossRef]
  375. Li, C.; Dong, L.; Su, R.; Bi, Y.; Qing, Y.; Deng, X.; Zhou, Y.; Hu, C.; Yu, M.; Huang, H.; et al. Homoharringtonine exhibits potent anti-tumor effect and modulates DNA epigenome in acute myeloid leukemia by targeting SP1/TET1/5hmC. Haematolgica 2020, 105, 148–160. [Google Scholar] [CrossRef] [Green Version]
  376. Sung, B.; Ahn, K.S.; Aggarwal, B.B. Noscapine, a benzylisoquinoline alkaloid, sensitizes leukemic cells to chemotherapeutic agents and cytokines by modulating the NF-κB signaling pathway. Cancer Res. 2010, 70, 3259–3268. [Google Scholar] [CrossRef] [Green Version]
  377. Yusenko, M.V.; Trentmann, A.; Andersson, M.K.; Ghani, L.A.; Jakobs, A.; Arteaga Paz, M.-F.A.; Mikesch, J.-H.; von Kries, J.P.; Stenman, G.; Klempnauer, K.-H. Monensin, a novel potent MYB inhibitor, suppresses proliferation of acute myeloid leukemia and adenoid cystic carcinoma cells. Cancer Lett. 2020, 479, 61–70. [Google Scholar] [CrossRef]
  378. Issa, M.E.; Berndt, S.; Carpentier, G.; Pezzuto, J.M.; Cuendet, M. Bruceantin inhibits multiple myeloma cancer stem cell proliferation. Cancer Biol. Ther. 2016, 17, 966–975. [Google Scholar] [CrossRef] [Green Version]
  379. Alachkar, H.; Santhanam, R.; Harb, J.G.; Lucas, D.M.; Oaks, J.J.; Hickey, C.J.; Pan, L.; Kinghorn, A.D.; Caligiuri, M.A.; Perrotti, D.; et al. Silvestrol exhibits significant in vivo and in vitro antileukemic activities and inhibits FLT3 and miR-155 expressions in acute myeloid leukemia. J. Hematol. Oncol. 2013, 6, 21. [Google Scholar] [CrossRef] [Green Version]
  380. Li, Y.; Gong, L.; Qi, R.; Sun, Q.; Xia, X.; He, H.; Ren, J.; Zhu, O.; Zhuo, D. Paeoniflorin suppresses pancreatic cancer cell growth by upregulating HTRA3 expression. Drug Des. Dev. Ther. 2017, 11, 2481–2491. [Google Scholar] [CrossRef] [Green Version]
  381. Shan, B.; Wang, L.; Lu, A.; Liu, X.; Sang, M.; Meng, F.; Cao, Q.; Ji, X. The flavonoid Baohuoside-I inhibits cell growth and downregulates survivin and cyclin D1 expression in esophageal carcinoma via β-catenin-dependent signaling. Oncol. Rep. 2011, 26, 1149–1156. [Google Scholar] [CrossRef]
  382. Kim, J.Y.; Lee, S.G.; Chung, J.-Y.; Kim, Y.-J.; Park, J.-E.; Koh, H.; Han, M.S.; Park, Y.C.; Yoo, Y.H.; Kim, J.-M. Ellipticine induces apoptosis in human endometrial cancer cells: The potential involvement of reactive oxygen species and mitogen-activated protein kinases. Toxicology 2011, 289, 91–102. [Google Scholar] [CrossRef] [PubMed]
  383. Yang, H.; Liu, C.; Zhang, Y.-q.; Ge, L.-t.; Chen, J.; Jia, X.-q.; Gu, R.-x.; Sun, Y.; Sun, W.-d. Ilexgenin A induces B16-F10 melanoma cell G1/S arrest in vitro and reduces tumor growth in vivo. Int. Immunopharmacol. 2015, 24, 423–431. [Google Scholar] [CrossRef] [PubMed]
  384. Saisomboon, S.; Kariya, R.; Vaeteewoottacharn, K.; Wongkham, S.; Sawanyawisuth, K.; Okada, S. Antitumor effects of flavopiridol, a cyclin-dependent kinase inhibitor, on human cholangiocarcinoma in vitro and in an in vivo xenograft model. Heliyon 2019, 5, e01675. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  385. Khumkhrong, P.; Piboonprai, K.; Chaichompoo, W.; Pimtong, W.; Khongkow, M.; Namdee, K.; Jantimaporn, A.; Japrung, D.; Asawapirom, U.; Suksamrarn, A.; et al. Crinamine Induces Apoptosis and Inhibits Proliferation, Migration, and Angiogenesis in Cervical Cancer SiHa Cells. Biomolecules 2019, 9, 494. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  386. Du, J.; Tang, B.; Wang, J.; Sui, H.; Jin, X.; Wang, L.; Wang, Z. Antiproliferative effect of alpinetin in BxPC-3 pancreatic cancer cells. Int. J. Mol. Med. 2012, 29, 607–612. [Google Scholar] [CrossRef] [Green Version]
  387. Liang, W.; Lai, Y.; Zhu, M.; Huang, S.; Feng, W.; Gu, X. Combretastatin A4 Regulates Proliferation, Migration, Invasion, and Apoptosis of Thyroid Cancer Cells via PI3K/Akt Signaling Pathway. Med. Sci. Monit. Int. Med. J. Exp. Clin. Res. 2016, 22, 4911–4917. [Google Scholar] [CrossRef]
  388. Appadurai, P.; Rathinasamy, K. Indicine N-oxide binds to tubulin at a distinct site and inhibits the assembly of microtubules: A mechanism for its cytotoxic activity. Toxicol. Lett. 2014, 225, 66–77. [Google Scholar] [CrossRef]
  389. Zheng, Q.; Li, Q.; Zhao, G.; Zhang, J.; Yuan, H.; Gong, D.; Guo, Y.; Liu, X.; Li, K.; Lin, P. Alkannin induces cytotoxic autophagy and apoptosis by promoting ROS-mediated mitochondrial dysfunction and activation of JNK pathway. Biochem. Pharmacol. 2020, 180, 114167. [Google Scholar] [CrossRef]
  390. Lee, J.-G.; Kim, J.-H.; Ahn, J.-H.; Lee, K.-T.; Baek, N.-I.; Choi, J.-H. Jaceosidin, isolated from dietary mugwort (Artemisia princeps), induces G2/M cell cycle arrest by inactivating cdc25C-cdc2 via ATM-Chk1/2 activation. Food Chem. Toxicol. 2013, 55, 214–221. [Google Scholar] [CrossRef]
  391. Kim, T.W.; Lee, S.Y.; Kim, M.; Cheon, C.; Ko, S.-G. Kaempferol induces autophagic cell death via IRE1-JNK-CHOP pathway and inhibition of G9a in gastric cancer cells. Cell Death Dis. 2018, 9, 875. [Google Scholar] [CrossRef] [Green Version]
  392. Fukuda, H.; Nakamura, S.; Chisaki, Y.; Takada, T.; Toda, Y.; Murata, H.; Itoh, K.; Yano, Y.; Takata, K.; Ashihara, E. Daphnetin inhibits invasion and migration of LM8 murine osteosarcoma cells by decreasing RhoA and Cdc42 expression. Biochem. Biophys. Res. Commun. 2016, 471, 63–67. [Google Scholar] [CrossRef]
  393. Hossan, M.S.; Chan, Z.-Y.; Collins, H.M.; Shipton, F.N.; Butler, M.S.; Rahmatullah, M.; Lee, J.B.; Gershkovich, P.; Kagan, L.; Khoo, T.-J.; et al. Cardiac glycoside cerberin exerts anticancer activity through PI3K/AKT/mTOR signal transduction inhibition. Cancer Lett. 2019, 453, 57–73. [Google Scholar] [CrossRef]
  394. Deng, X.; Sheng, J.; Liu, H.; Wang, N.; Dai, C.; Wang, Z.; Zhang, J.; Zhao, J.; Dai, E. Cinobufagin Promotes Cell Cycle Arrest and Apoptosis to Block Human Esophageal Squamous Cell Carcinoma Cells Growth via the p73 Signalling Pathway. Biol. Pharm. Bull. 2019, 42, 1500–1509. [Google Scholar] [CrossRef] [Green Version]
  395. Ma, J.; Wang, L.; Li, J.; Zhang, G.; Tao, H.; Li, X.; Sun, D.; Hu, Y. Swainsonine Inhibits Invasion and the EMT Process in Esophageal Carcinoma Cells by Targeting Twist1. Oncol. Res. 2018, 26, 1207–1213. [Google Scholar] [CrossRef]
  396. Lu, C.-H.; Chen, S.-H.; Chang, Y.-S.; Liu, Y.-W.; Wu, J.-Y.; Lim, Y.-P.; Yu, H.-I.; Lee, Y.-R. Honokiol, a potential therapeutic agent, induces cell cycle arrest and program cell death in vitro and in vivo in human thyroid cancer cells. Pharmacol. Res. 2017, 115, 288–298. [Google Scholar] [CrossRef]
  397. Dong, Y.; Cao, A.; Shi, J.; Yin, P.; Wang, L.; Ji, G.; Xie, J.; Wu, D. Tangeretin, a citrus polymethoxyflavonoid, induces apoptosis of human gastric cancer AGS cells through extrinsic and intrinsic signaling pathways. Oncol. Rep. 2014, 31, 1788–1794. [Google Scholar] [CrossRef]
  398. Rabi, T.; Catapano, C.V. Aphanin, a triterpenoid from Amoora rohituka inhibits K-Ras mutant activity and STAT3 in pancreatic carcinoma cells. Tumor Biol. 2016, 37, 12455–12464. [Google Scholar] [CrossRef]
  399. Pelinson, L.P.; Assmann, C.E.; Palma, T.V.; da Cruz, I.B.M.; Pillat, M.M.; Mânica, A.; Stefanello, N.; Weis, G.C.C.; de Oliveira Alves, A.; de Andrade, C.M. Antiproliferative and apoptotic effects of caffeic acid on SK-Mel-28 human melanoma cancer cells. Mol. Biol. Rep. 2019, 46, 2085–2092. [Google Scholar] [CrossRef]
  400. Mansingh, D.P.; OJ, S.; Sali, V.K.; Vasanthi, H.R. [6]-Gingerol–induced cell cycle arrest, reactive oxygen species generation, and disruption of mitochondrial membrane potential are associated with apoptosis in human gastric cancer (AGS) cells. J. Biochem. Mol. Toxicol. 2018, 32, e22206. [Google Scholar] [CrossRef]
  401. Yong, W.K.; Abd Malek, S.N. Xanthohumol Induces Growth Inhibition and Apoptosis in Ca Ski Human Cervical Cancer Cells. Evid.-Based Complement. Altern. Med. 2015, 2015, 921306. [Google Scholar] [CrossRef] [Green Version]
  402. An, F.; Wang, S.; Tian, Q.; Zhu, D. Effects of orientin and vitexin from Trollius chinensis on the growth and apoptosis of esophageal cancer EC-109 cells. Oncol. Lett. 2015, 10, 2627–2633. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  403. Yang, L.; Yu, Y.; Zhang, Q.; Li, X.; Zhang, C.; Mao, T.; Liu, S.; Tian, Z. Anti-gastric cancer effect of Salidroside through elevating miR-99a expression. Artif. Cells Nanomed. Biotechnol. 2019, 47, 3500–3510. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  404. Júnior, P.L.d.S.; Câmara, D.A.D.; Costa, A.S.; Ruiz, J.L.M.; Levy, D.; Azevedo, R.A.; Pasqualoto, K.F.M.; de Oliveira, C.F.; de Melo, T.C.; Pessoa, N.D.S.; et al. Apoptotic effect of eugenol envolves G2/M phase abrogation accompanied by mitochondrial damage and clastogenic effect on cancer cell in vitro. Phytomedicine 2016, 23, 725–735. [Google Scholar] [CrossRef] [PubMed]
  405. Kuete, V.; Sandjo, L.P.; Seukep, J.A.; Zeino, M.; Mbaveng, A.T.; Ngadjui, B.; Efferth, T. Cytotoxic compounds from the fruits of Uapaca togoensis towards multifactorial drug-resistant cancer cells. Planta Med. 2015, 81, 32–38. [Google Scholar] [CrossRef] [PubMed]
  406. Li, C.; Wang, Y.; Wang, C.; Yi, X.; Li, M.; He, X. Anticancer activities of harmine by inducing a pro-death autophagy and apoptosis in human gastric cancer cells. Phytomedicine 2017, 28, 10–18. [Google Scholar] [CrossRef]
  407. He, L.; Wu, Y.; Lin, L.; Wang, J.; Wu, Y.; Chen, Y.; Yi, Z.; Liu, M.; Pang, X. Hispidulin, a small flavonoid molecule, suppresses the angiogenesis and growth of human pancreatic cancer by targeting vascular endothelial growth factor receptor 2-mediated PI3K/Akt/mTOR signaling pathway. Cancer Sci. 2011, 102, 219–225. [Google Scholar] [CrossRef]
  408. Edler, M.C.; Fernandez, A.M.; Lassota, P.; Ireland, C.M.; Barrows, L.R. Inhibition of tubulin polymerization by vitilevuamide, a bicyclic marine peptide, at a site distinct from colchicine, the vinca alkaloids, and dolastatin 10. Biochem. Pharmacol. 2002, 63, 707–715. [Google Scholar] [CrossRef]
  409. Zou, P.; Xia, Y.; Ji, J.; Chen, W.; Zhang, J.; Chen, X.; Rajamanickam, V.; Chen, G.; Wang, Z.; Chen, L.; et al. Piperlongumine as a direct TrxR1 inhibitor with suppressive activity against gastric cancer. Cancer Lett. 2016, 375, 114–126. [Google Scholar] [CrossRef]
  410. Tian, B.; Xiao, Y.; Ma, J.; Ou, W.; Wang, H.; Wu, J.; Tang, J.; Zhang, B.; Liao, X.; Yang, D.; et al. Parthenolide Inhibits Angiogenesis in Esophageal Squamous Cell Carcinoma Through Suppression of VEGF. OncoTargets Ther. 2020, 13, 7447–7458. [Google Scholar] [CrossRef]
  411. Zhao, Z.; Jia, Q.; Wu, M.-S.; Xie, X.; Wang, Y.; Song, G.; Zou, C.-Y.; Tang, Q.; Lu, J.; Huang, G.; et al. Degalactotigonin, a natural compound from Solanum nigrum L., inhibits growth and metastasis of osteosarcoma through GSK3β inactivation—Mediated repression of the Hedgehog/Gli1 pathway. Clin. Cancer Res. 2018, 24, 130–144. [Google Scholar] [CrossRef] [Green Version]
  412. Kashyap, D.; Tuli, H.S.; Yerer, M.B.; Sharma, A.; Sak, K.; Srivastava, S.; Pandey, A.; Garg, V.K.; Sethi, G.; Bishayee, A. Natural Product-Based Nanoformulations for Cancer Therapy: Opportunities and Challenges. In Seminars in Cancer Biology; Academic Press: Cambridge, MA, USA, 2019. [Google Scholar]
  413. Srinivasan, M.; Rajabi, M.; Mousa, S.A. Nanobiomaterials in Cancer Therapy. In Nanobiomaterials in Cancer Therapy; Elsevier: Amsterdam, The Netherlands, 2016; pp. 57–89. [Google Scholar]
  414. Mousa, D.S.; El-Far, A.H.; Saddiq, A.A.; Sudha, T.; Mousa, S.A. Nanoformulated Bioactive Compounds Derived from Different Natural Products Combat Pancreatic Cancer Cell Proliferation. Int. J. Nanomed. 2020, 15, 2259–2268. [Google Scholar] [CrossRef] [Green Version]
  415. Andima, M.; Costabile, G.; Isert, L.; Ndakala, A.J.; Derese, S.; Merkel, O.M. Evaluation of β-Sitosterol loaded PLGA and PEG-PLA nanoparticles for effective treatment of breast cancer: Preparation, physicochemical characterization, and antitumor activity. Pharmaceutics 2018, 10, 232. [Google Scholar] [CrossRef] [Green Version]
  416. Feng, J.; Xu, M.; Wang, J.; Zhou, S.; Liu, Y.; Liu, S.; Huang, Y.; Chen, Y.; Chen, L.; Song, Q.; et al. Sequential delivery of nanoformulated α-mangostin and triptolide overcomes permeation obstacles and improves therapeutic effects in pancreatic cancer. Biomaterials 2020, 241, 119907. [Google Scholar] [CrossRef]
  417. Liu, Y.; Gao, D.; Zhang, X.; Liu, Z.; Dai, K.; Ji, B.; Wang, Q.; Luo, L. Antitumor drug effect of betulinic acid mediated by polyethylene glycol modified liposomes. Mater. Sci. Eng. C 2016, 64, 124–132. [Google Scholar] [CrossRef]
  418. Liu, Y.; Zhao, L.; Shen, G.; Chang, R.; Zhang, Y.; Yan, X. Coordination self-assembly of natural flavonoids into robust nanoparticles for enhanced in vitro chemo and photothermal cancer therapy. Colloids Surf. A Physicochem. Eng. Asp. 2020, 598, 124805. [Google Scholar] [CrossRef]
  419. Ahmadi, E.; Zarghami, N.; Jafarabadi, M.A.; Alizadeh, L.; Khojastehfard, M.; Yamchi, M.R.; Salehi, R. Enhanced anticancer potency by combination chemotherapy of HT-29 cells with biodegradable, pH-sensitive nanoparticles for co-delivery of hydroxytyrosol and doxorubicin. J. Drug Deliv. Sci. Technol. 2019, 51, 721–735. [Google Scholar] [CrossRef]
  420. Abdelaziz, H.M.; Elzoghby, A.O.; Helmy, M.W.; Samaha, M.W.; Fang, J.-Y.; Freag, M.S. Liquid crystalline assembly for potential combinatorial chemo–herbal drug delivery to lung cancer cells. Int. J. Nanomed. 2019, 14, 499–517. [Google Scholar] [CrossRef] [Green Version]
  421. Xia, Q.; Ling, L.; Ismail, M.; Du, Y.; He, W.; Zhou, W.; Yao, C.; Li, X. Paclitaxel encapsulated in artesunate-phospholipid liposomes for combinatorial delivery. J. Drug Deliv. Sci. Technol. 2019, 51, 372–382. [Google Scholar] [CrossRef]
  422. Wang, S.; Shao, M.; Zhong, Z.; Wang, A.; Cao, J.; Lu, Y.; Wang, Y.; Zhang, J. Co-delivery of gambogic acid and TRAIL plasmid by hyaluronic acid grafted PEI-PLGA nanoparticles for the treatment of triple negative breast cancer. Drug Deliv. 2017, 24, 1791–1800. [Google Scholar] [CrossRef] [Green Version]
  423. Bian, Y.; Guo, D. Targeted Therapy for Hepatocellular Carcinoma: Co-Delivery of Sorafenib and Curcumin Using Lactosylated pH-Responsive Nanoparticles. Drug Des. Dev. Ther. 2020, 14, 647. [Google Scholar] [CrossRef] [Green Version]
  424. Wang, D.; Zhou, J.; Chen, R.; Shi, R.; Zhao, G.; Xia, G.; Li, R.; Liu, Z.; Tian, J.; Wang, H.; et al. Controllable synthesis of dual-MOFs nanostructures for pH-responsive artemisinin delivery, magnetic resonance and optical dual-model imaging-guided chemo/photothermal combinational cancer therapy. Biomaterials 2016, 100, 27–40. [Google Scholar] [CrossRef] [PubMed]
  425. Wang, J.; Muhammad, N.; Li, T.; Wang, H.; Liu, Y.; Liu, B.; Zhan, H. Hyaluronic Acid-Coated Camptothecin Nanocrystals for Targeted Drug Delivery to Enhance Anticancer Efficacy. Mol. Pharm. 2020, 17, 2411–2425. [Google Scholar] [CrossRef] [PubMed]
  426. Gupta, L.; Sharma, A.K.; Gothwal, A.; Khan, M.S.; Khinchi, M.P.; Qayum, A.; Singh, S.K.; Gupta, U. Dendrimer encapsulated and conjugated delivery of berberine: A novel approach mitigating toxicity and improving in vivo pharmacokinetics. Int. J. Pharm. 2017, 528, 88–99. [Google Scholar] [CrossRef] [PubMed]
  427. Ding, J.; Liang, T.; Min, Q.; Jiang, L.-P.; Zhu, J.-J. “Stealth and Fully-Laden” Drug Carriers: Self-Assembled Nanogels Encapsulated with Epigallocatechin Gallate and siRNA for Drug-Resistant Breast Cancer Therapy. ACS Appl. Mater. Interfaces 2018, 10, 9938–9948. [Google Scholar] [CrossRef]
  428. Amaral, R.; dos Santos, S.A.; Andrade, L.N.; Severino, P.; Carvalho, A.A. Natural Products as Treatment against Cancer: A Historical and Current Vision. Clin. Oncol. 2019, 4, 1562. [Google Scholar]
  429. Mita, A.; Lockhart, A.C.; Chen, T.-L.; Bochinski, K.; Curtright, J.; Cooper, W.; Hammond, L.; Rothenberg, M.; Rowinsky, E.; Sharma, S. A phase I pharmacokinetic (PK) trial of XAA296A (Discodermolide) administered every 3 wks to adult patients with advanced solid malignancies. J. Clin. Oncol. 2004, 22, 2025. [Google Scholar] [CrossRef]
Figure 1. Curcumin inhibits cancer progression by regulating many signal pathways. (I) Akt/PI3K/mTOR signaling pathway. PTEN inhibits Akt activation by PI3K, mTOR phosphorylates p70S6K1 (S6K1), and 4E-BP1 which activate cell growth and survival pathways. Curcumin inhibits the Akt/PI3K/mTOR pathway by increasing PTEN expression through decreased miR-21 levels, and by inhibiting PI3K activity by upregulating miR-192-5p, miR-206, and miR-203; (II) MAPK signaling pathway. Signal cascades lead to activation of the MEKs, which subsequently activate the ERK1/2, p38, and JNK cascades and initiate the transcription of genes. Curcumin activates the p38 MAPK pathway by upregulating miR-378, which increases p21/27, cleaves caspase-3,9, and decreases Bcl-2 and MMP2/9. (III) Wnt/β-catenin pathway. The wnt molecule binds to both Frizzled and LRP5/6 receptors to cause the Axin/APC/GSK3β complex to dissociate. As a result, β-catenin is phosphorylated and translocated to the nucleus where it binds TCF/LEF co-transcription factors, which stimulate the transcription of Wnt-response genes. Curcumin inhibits Wnt/ β -catenin signaling through inhibition of lncRNA H19, lincROR, miR-130, 21, and upregulation of miR-192-5p and miR-33b; and (IV) JAK/STAT signaling. This pathway is activated when a ligand binds to a receptor, which communicates signals downstream STATs, whereas STATs are transcription factors that regulate gene expression; (V) p53 signaling pathway. The activation of MDM2 by AKT may inhibit the antitumor activity of p53. Curcumin inhibits lncRNA H19, miR-1246, and miR-19 to increase p53’s anti-tumor activity; (VI) NF-kB signaling pathway. This signaling cascade results in the phosphorylation of IkBα, which is then degraded by the proteasome. This allows the NF-kB/p65/p50 complex to translocate to the nucleus in order to facilitate transcription. This figure was reproduced from [32] (licensed under creative commons license).
Figure 1. Curcumin inhibits cancer progression by regulating many signal pathways. (I) Akt/PI3K/mTOR signaling pathway. PTEN inhibits Akt activation by PI3K, mTOR phosphorylates p70S6K1 (S6K1), and 4E-BP1 which activate cell growth and survival pathways. Curcumin inhibits the Akt/PI3K/mTOR pathway by increasing PTEN expression through decreased miR-21 levels, and by inhibiting PI3K activity by upregulating miR-192-5p, miR-206, and miR-203; (II) MAPK signaling pathway. Signal cascades lead to activation of the MEKs, which subsequently activate the ERK1/2, p38, and JNK cascades and initiate the transcription of genes. Curcumin activates the p38 MAPK pathway by upregulating miR-378, which increases p21/27, cleaves caspase-3,9, and decreases Bcl-2 and MMP2/9. (III) Wnt/β-catenin pathway. The wnt molecule binds to both Frizzled and LRP5/6 receptors to cause the Axin/APC/GSK3β complex to dissociate. As a result, β-catenin is phosphorylated and translocated to the nucleus where it binds TCF/LEF co-transcription factors, which stimulate the transcription of Wnt-response genes. Curcumin inhibits Wnt/ β -catenin signaling through inhibition of lncRNA H19, lincROR, miR-130, 21, and upregulation of miR-192-5p and miR-33b; and (IV) JAK/STAT signaling. This pathway is activated when a ligand binds to a receptor, which communicates signals downstream STATs, whereas STATs are transcription factors that regulate gene expression; (V) p53 signaling pathway. The activation of MDM2 by AKT may inhibit the antitumor activity of p53. Curcumin inhibits lncRNA H19, miR-1246, and miR-19 to increase p53’s anti-tumor activity; (VI) NF-kB signaling pathway. This signaling cascade results in the phosphorylation of IkBα, which is then degraded by the proteasome. This allows the NF-kB/p65/p50 complex to translocate to the nucleus in order to facilitate transcription. This figure was reproduced from [32] (licensed under creative commons license).
Molecules 27 08367 g001
Figure 2. The mechanism of chemoresistance in cancer cells may include a variety of molecular mechanisms, such as regulating drug influx and efflux through ABC transporters, inhibiting cell death, altering drug targets, regulating epigenetic factors, inactivating chemotherapeutic agents, inactivating tumor suppressor genes, modifying DNA repair processes, and modulating growth factor signaling (adapted from [33], License Number: 5433490420433, License date: 21 November 2022).
Figure 2. The mechanism of chemoresistance in cancer cells may include a variety of molecular mechanisms, such as regulating drug influx and efflux through ABC transporters, inhibiting cell death, altering drug targets, regulating epigenetic factors, inactivating chemotherapeutic agents, inactivating tumor suppressor genes, modifying DNA repair processes, and modulating growth factor signaling (adapted from [33], License Number: 5433490420433, License date: 21 November 2022).
Molecules 27 08367 g002
Figure 3. Several aspects of the development of cancer and the potential role of natural products in cancer prevention and therapy (adapted from [46], licensed under creative common attribution license).
Figure 3. Several aspects of the development of cancer and the potential role of natural products in cancer prevention and therapy (adapted from [46], licensed under creative common attribution license).
Molecules 27 08367 g003
Figure 4. Approved marketed anticancer drugs from 1950 to 2014 which are directly obtained or derived from natural products or imitate them (adopted from [428], licensed under creative common attribution license).
Figure 4. Approved marketed anticancer drugs from 1950 to 2014 which are directly obtained or derived from natural products or imitate them (adopted from [428], licensed under creative common attribution license).
Molecules 27 08367 g004
Table 1. Some of the important natural products in managing different types of cancers.
Table 1. Some of the important natural products in managing different types of cancers.
SourcesBioactive ConstituentsChemical FormulaMolecular Weight
(g/mol)
MechanismsStudy ModelConcentrationsOther CancersReferences
Lung cancer
Fascaplysinopsis BergquistFascaplysinC18H11ClN2O306.7↓CDK-4, ↑ROS, ↓mTOR, ↓4EBP1, ↓p70S6K1A549, H1299, PC-9 and NCI-H526 cell linesIC50: 0.57–1.15 µMLiver, ovarian, and colon cancer[195]
Chelidonium majusChelidonineC20H19NO5353.4↓cyclin B1, ↑p21, ↓MMP, ↑ROS/AMPK, ↓p-p70S6K, ↓EGFR, G2/M phase arrestA549, PC9, H460, H358 cells and H1975 xenograft in miceIC50: 2.58–20 µMBreast, liver, head, and neck cancer[196]
Cephaelis IpecacuanhaEmetineC29H40N2O4480.6↓MMP-2/-9, ↓ERK1/2, and ↑p38A549 and H1299 cell lineIC50: 0.243 and 1 µMBreast, colon, prostate, and pancreatic cancer[197]
Citrus limonD-limoneneC10H16136.23↑Bax, ↓Bcl-2, ↑cleaved PARP, ↑Atg5H1299 and A549 cells xenograft in Balb/c mice0.5–0.75 mMSkin, liver, breast, and kidney cancer[198]
Artemisia annuaSesquiterpene (AMDT)C17H26O2262.39↑caspase 3- 9, G1/S phase arrest95-D, HO8910, QGY, and HeLa cellsIC50: 52.44–73.3 µMGastric and ovarian cancer[199]
ArtemisininC15H22O5282.33↓VEGF-C, ↓p38 MAPKLLC cells and C57BL/6 mice5–20 µMBreast, cervical, gastric, liver, and bladder cancer[200]
ArtemetherC16H26O5298.37↓Bcl-2, ↓cIAP1, ↓cIAP2, ↓CDK1, ↓CDK2, ↓CDK6, ↓cyclin A2, ↓cyclin B1, ↓cyclin D1, ↑p16A549, NCI-H1299 cells40–80 µMBrain, breast, and gastric cancer[201]
Betula platyphyllaBetulinC30H50O2442.7↑caspase 3- 9 and 8, ↑ Bax, ↑Cyto- C, ↑Bak, ↓PARPA549, HepG2, Hela, MCF-7 cellsIC50: 10–15 µg/mLGastric, renal, colon, and melanoma[202]
Betulinic AcidC30H48O3456.7↓cyclin A2, ↓Sp1, G2/M phase arrestA549, Hela, H1299 xenograft mice and Scgb1a1-rtTA/TetO-KrasG12Dmice≤40 µMColon, breast, prostate, nasopharyngeal, and cervical cancer[203,204]
GarciniaGambogic acidC38H44O8628.7↑Caspase-3, ↓Bcl2, ↓pI3K, DLL1, ↓DLL3, ↓ DLL4, ↓Jagged1, ↓Jagged2SPC-A-1 and A549 cells0.5–1 µmol/LLiver and colorectal[205]
Laminaria japonicaFucoxanthinC42H58O6658.9↑p21, ↑p53, ↑puma, ↑Fas, ↓Bcl-2, G0/G1 phase arrestSPC-A1, A549, H460, H1299IC50: <82 µMSkin, gastric, breast, and colon cancer[206]
Chinese trichosanthesTrichosanthin247 Amino acids27 kDa↑Caspase-3 and 9, ↑E-cadherin, ↓N-cadherin, ↓Bcl-2, ↑ Bax, ↑DR4, ↑DR5NCI-H1299, NCI-H1975, and A549 cellsIC20: 39.5, 61.1 and 49.3 µg/mLCervical, gastric, breast, and nasopharyngeal[207]
HedyotisUrsolic acidC30H48O3456.7↓MMP2, ↓MMP 9, ↑E-cadherin, ↓N-cadherinH1975 cellsIC50: 29.7 nMBreast, prostate, cervical, colorectal, and liver cancer[208]
Oleanolic acidC30H48O3456.7↑ROS, ↑Ca2+, ↓Akt, ↓JACk2, ↓STAT3A549 cellsIC50: 9.66 µMGallbladder, prostate, and gastric cancer[209]
Scutellaria barbataWogoninC16H12O5284.26↑caspase-3, 8 and 9, ↑ROSA427, BEAS-2B and A459 cells50 µMBreast, cervical, and nasopharyngeal carcinoma[210]
LuteolinC15H10O6286.24↑miR-34a-5p, ↑p21, ↑p53, ↓MDM4, ↑caspase-3 and 9A549, and H460 xenograft modelIC50: 40 µMBreast, ovarian, gastric, and prostate cancer[211]
Polygonum cuspidatumPolydatinC20H22O8390.4↓NLRP3, ↓ASC, ↓p-NF-κβ p65A549 and H1299 cells50 µMBreast, hepatic, renal, and ovarian cancer[212]
Rheum officinaleEmodinC15H10O5270.24↑PPARγ, ↑IGFBP1, ↓Sp1H1975, and A549 cells xenograft mice50 µMRenal and cervical cancer[213]
Rhizoma zedoariaeβ-elemeneC15H24204.35↑caspase-3, ↑ROS, ↑Cyto- C, ↑Bad, ↓Bcl-2A549/DDP cells20 and 40 µg/mLLiver, gastric, and bladder cancer[214]
ChansuBufalinC24H34O4386.5↑caspase-3, ↓Akt, ↓ ERK ½, ↓VEGF, ↓cMyc, ↓NF-κβ, ↓p38 MAPK, G1/S phase arrestA549, NCI-H4602.5–10 µMProstate, liver, and ovarian cancer[215,216]
Platycodon grandiflorumPlatycodin-DC57H92O281225.3↑Atg-3, ↑Atg-7, ↑Beclin-1, ↑LC3-II, ↓ p-Akt, ↓p-p70S6K, ↓p-4EBP1, ↓ ERK ½NCI-H460 and A549 cells5–30 µmol/LColon, leukemia, and breast cancer[217]
Salvia miltiorrhizaTanshinone IIA (TSIIA)C19H18O3294.34↑Cyto- C, ↓ Bax, ↓MMP, ↑caspase-3 and 9A549 cellsIC50: 14.5 µMProstate, glioma, leukemia, and hepatoma[218]
Tripterygium wilfordii HookCelastrolC29H38O4450.6↓Bcl-2, ↑ Bax, ↓Akt, ↑Cyto- C, ↓PARP, ↑Fas, ↑FasLA549 cellsIC50: 2.12 µMProstate, brain, and liver cancer[219]
Curcuma longaCurcuminC21H20O6368.4↑ROS, ↓ΔΨm, ↓SOD1, ↑caspase-3 and 9, ↑ Bax, G2/M phase arrestA549 and SPC-A1 cells10–25 µMPancreatic, gastric, prostate, and colorectal cancer[220,221]
Cinnamomum cassiaCinnamaldehydeC9H8O132.16↑caspase-3, ↑PARP, ↓Bcl-2,↑Bax,↓Bcl-xL, ↓HIF-1α, ↓β-catenin, ↓MMPs, ↑E-cadherinNCI-H1299, YTMLC, A549 xenograft modelIC50: 10.5, 32, 41 µg/mLoral squamous cell carcinoma and colorectal cancer[222]
Coix lacryma-jobi L.CP-1 polysacchariden.d.n.d.↑caspase-3 and 9, ↓ΔΨm, S phase arrestA549 cells10–300 µg/mLBreast and colon cancer[223]
Ophiopogon japonicusOphiopogonin BC39H62O12722.9↑E-cadherin, ↓N-cadherin, ↓linc00668, ↑miR-432-5pA549, 293T and THP-1 cells5 and 10 µmol/LColon, breast, and gastric cancer[224]
Tetrastigma hemsleyanumRadix Tetrastigma Hemsleyani FlavoneC27H30O10514.5↓MMP 2 and 9, ↓TIMP-1, ↑TIMP-2A549 cells0.5–10 mg/mLLiver, leukemia, and stomach cancer[225]
Panax notoginsengTrilinoleinC57H98O6879.4↓Bcl-2, ↑ Bax, ↑ROS, ↓PARP, ↓Akt, ↑caspase-3, ↑Cyto- CA549, MKN-45 and A498 cells25–100 µg/mLLiver, colon, and gastric cancer[226]
Fucus vesiculosusFucoidanC7H14O7S242.25↑ROS, ↑ATF4, ↑CHOP, ↑TLR4A549, CL1-5 and LLC-1 xenograft C57BL/6 mice100–800 µg/mLBreast, liver, and colon cancer[227]
Breast cancer
Rabdosia rubescensOridoninC20H28O6364.4↓Jagged2, ↓Notch1-44T1 cells and xenograft BALB/c mice0.1–10 mmol/LGastric, esophageal, prostate, and pancreatic cancer[228]
Panax notoginsengPanaxadiol saponinsC30H52O3460.73↓NLR, ↓MPO, ↓G-CSF, ↓PU.1, ↓C/EBPα, ↓EP1, ↓GATA 1 and 24T1-Luc and xenograft BALB/c mice0.1–40 µMLung, liver, and colon cancer[229]
Ganoderma lucidumGanodermanontriolC30H48O4472.7↓u-PA, ↓CDC20, ↓survivinMDA-MB-231 cellsIC50: 11.6 µMGastric, prostate, and colon cancer[230]
Polysaccharidesn.d.n.d.↑SOD, ↑CAT, ↑GPx, ↓IL-1β, ↓IL-6, ↓TNF-αMDA-MB-231 cells, Wistar rats0.5–3 mg/mLMelanoma, cervical, and colorectal cancer[231,232]
Maclura pomiferaPomiferinC25H24O6420.5↑CANX, ↑BCAP31, ↑Mn-SOD,↑ACOX1,↓BMP-7 ↓ID 2 and 3MCF-7 and MCF-10AIC50: 5.2 µMLiver, cholangiocarcinoma, and colon cancer[233]
ChansuBufadienolidesC24H34O2354.52↓STAT3/Mcl-1, ↓Bcl-xL, ↓PARP, ↑caspase-3MDA-MB-231, MCF-10A and MCF-7 cells0.1–0.5 µMLiver, leukemia, lung, and gastric cancer[234]
Cimicifuga foetidaActeinC37H56O11676.8↓VEGF, ↓pERK, ↓pJNK, ↓CD34, ↓CXCR4HMEC-1, 4T1 cells xenograft in miceIC50: 0.065 µMGastric, bladder, and lung cancer[235]
Astragalus membranaceusFormononetinC16H12O4268.26↓MMP 2, ↓MMP9, ↓PI3K, ↓Akt, ↑TIMP-1 and 2MDA-MB-231 and 4T1 cells2.5–160 µmol/LCervical, bladder, and gastric cancer[236]
Astragulus Polysaccharidesn.d.n.d.↓Bcl-2, ↑ Bax, ↑NO, ↑TNF-α, G1/S phase arrestMCF-7 cells100–1000 µg/mLLiver, lung, and gastric cancer[237]
Narcissus L. bulbNarciclasineC14H13NO7307.25↑AMPK-ULK1, ↓PRAS40MDA-MB-231, MCF-7, MCF-10A, BT483, HCC1937 and xenograft in miceIC50: 5–100 nMBrain, liver, and hematological cancers[238]
Cimicifuga racemosaKHF16C37H58O10662.85↓XIAP, Mcl-1, ↓Survivin, ↓Cyclin B1/D1, ↓NF-kβMDA-MB-231, MDA-MB-468 and SW527IC50: <10 µMGastric, bladder, and lung cancer[239]
Pueraria lobataPuerarinC21H20O9416.4↓NFκβ, ↑AMPK, ↑ACC, ↑GSK-3β, ↓CREBMCF-7/Adr cells20–100 µMBladder, liver, and lung cancer[240]
Gardenia jasminoides EllisGenipinC11H14O5226.23↓Bcl-2, ↑ Bax, ↑caspase-3, ↑JNK, ↑p38 MAPKMDA-MB-231 cellsIC50: 327 µMGastric, multiple myeloma, and bladder cancer[241]
Tetradium ruticarpumEvodiamineC19H17N3O303.4↓u-PA, ↓MMP 9, Bcl-2, ↑ Bax, ↓CD1, ↓CDK6, G0/G1 arrestMDA-MB-231 cells and xenograft BALB/c miceIC50: 90 µMBladder, colorectal, pancreatic, and melanoma[242]
Euphorbia prolifera Buch-HamMyrsinol diterpene
(J196-10-1)
n.d.n.d.↓MDR, ↓Pgp efflux, ↑ATP hydrolysisMCF-7/Adr cells0.39–50 µMGastric, colon and lung cancer[243]
Solanum nigrum L.α-SolanineC45H73NO15868.1↑Bax, ↓Bcl-2, ↓Ψm4T1 cells and xenograft BALB/c miceIC50: 17 µMPancreatic, esophageal, and prostate cancer[244]
Melonidus suaveolensMelosuavine IC41H42N4O6686.79↑caspase 3, ↓Bcl-2, ↑p53BT549 cellsIC50: 0.89 µMLung, colon, and prostate cancer[245]
Cnidium monnieri (L.) Cusson.OstholeC15H16O3244.28↑caspase 9 and 3, ↓PARP, ↑p53 and p23, ↓Cdk2 and CD1MDA-MB 435 cells20–100 µmol/LLiver and lung cancer[246]
Ophiopogon japonicusOphiopogonin DC44H70O16855↑Caspase 9 and 8, ↓cyclin B1 and G2/M arrestMCF-7 cells12.5–50 µmol/LProstate and laryngeal carcinoma[247]
Panax ginseng C. A. Mey.Ginsenoside Rg5C42H70O12767↑p53, p21 and p15, ↓CD1, CE2 and Cdk4, ↑caspase 6,7,8 and 9, ↓PARPMCF-7 cells25–100 µMLeukemia, gastric, and cervical cancer[248]
Epimedium brevicornumIcariside IIC27H30O10514.5↑Cl-Caspase-9,8,7,3, ↑ Cl-PARP, ↑ Bax, ↑Bcl-xL, ↑BimL, ↑Fas, ↓ FasL, ↑FADD, ↓MMP, ↑Cyto- C, ↑AIFMCF-7 and MDA-MB-231 cellsIC50: 72.73 and 97.14 µMLung, melanoma, epidermoid, and prostate cancer[249]
Ovarian cancer
Tripterygium wilfordii HookTriptolideC20H24O6360.4↓ MMP7 and MMP19, ↑EcadherinSKOV3, A2780 cells and SKVO3 xenograft mice1.5–150 nMColon, renal, and cervical cancer[250]
Panax ginsengGinsenoside 20(S)-Rg3C42H72O13785↑ Caspase-3 and 9, ↓ PI3K/Akt and IAPHO-8910 cells25–100 µg/mLLung, prostate, and breast cancer[251]
Syzygium aromaticumKumatakeninC17H14O6314.29↑ Caspase-3,8 and 9, ↓MCP-1, ↓RANTES, ↓IL-10, ↓VEGF, ↓MMP 2 and 9SKOV3 and A2780 cells<100 µg/mLBreast, liver, colon, and gastric cancer[252]
Pueraria mirificaDaidzeinC15H10O4254.24↑Bcl-2, ↑cym-c, ↑cleaved caspase-3/9, G2/M cell arrest, ↓pCdc25c, ↓pCdc2, ↓cyclin B1, ↓MMP-2/9SKOV3 cell lineIC50: 20 µMBreast, colon, bladder, and pancreatic cancer[253]
Thelypteris torresiana (Gaud)ProtoapigenoneC15H10O6286.24↓p-Cdk2, ↓Cdk2, ↓p-Cyclin B1, ↓Cyclin B1, ↑p-Cdc25C, ↓Bcl-xL, ↓Bcl-2, ↑caspase-3MDAH-2774 and SKOV3 cellsIC50: 0.69 and 0.78 µMLung, breast, and prostate cancer[254]
Epimedium brevicornumIcariinC33H40O15676.7↑Caspase-3, ↓miR-21, ↑PTEN and RECK, ↓Bcl-2A2780 cells13–100 µMEsophageal, prostate, and liver cancer[255]
IcaritinC21H20O6368.4↑caspase 3- 9, ↑p53 and ↓Akt/mTOR pathwayOV2002, C13*, A2780cp and PDXs in NOD/SCID mice10–50 µMEndometrial, bladder, colorectal, and prostate cancer[256]
Gundelia tournefortiiStigmasterolC29H48O412.7↑ROS, ↑caspase 3 and 9, Bax, ↑BAk, ↑cym-c, ↓VEGFA, ↓MMP 2,9 and 14OV90 and ES2 cells5–20 µg/mLGastric, skin, liver, and lung cancer[257]
Vitex Agnus-castus L.CasticinC19H18O8374.3↑FOXO3a, ↓FoxM1, ↓survivin, ↓PLK1, p27KLP1SKOV3 and A2780 cells2.5–10 µmol/LGastric, gallbladder, cervical, and melanoma[258]
Carya cathayensis SargJugloneC10H6O3174.15↑ROS, ↑p21, ↑Bax, ↑Bad, ↑Cyto c, ↓CDK2, cdc25A, CHK1, and cyclin A, ↓Bcl-2 and Bcl-xL, ↓Cyclin A, S-phase cell cycle arrestIshikawa cellsIC50: 20.81 µMGlioma, lung, and leukemia[259]
Scutellaria baicalensis GeorgiBaicalin and BaicaleinC21H18O11, C15H10O5446.4, 270.24↓VEGF, HIF-1α, cMyc, and NFκBOVCAR-3, IOSE-364, and CP-70 cells5–160 µMBurkett lymphoma, colorectal, pancreatic, prostate, and osteosarcoma[260]
Potamogeton crispus L.Luteolin-3’-O-β-D-glucopyranosideC21H20O11448.4↓MMP-2, ↓MMP-9, G1/S phase arrestES-2 cells15–240 µg/mLColon and breast cancer[261]
Artemisia annuaartesunateC19H28O8384.4↑ROS, ↑p21, ↓CDKs, ↓Rb, ↓E2F-1, ↓CDC25C, G2/M cell cycle arrestHEY, IGROV-1, OVCAR8, and OVCAR3 cells, and ID8 xenograft in C57BL/6 miceIC50: 0.51–31.89 µMLeukemia, pancreatic, and breast cancer[262]
Arctium lappaArctigeninC21H24O6372.4↓STAT3, ↓survivin, ↓iNOS, ↑caspase3OVCAR3 and SKOV3 cellsIC50: 10 µMBreast, colon, and lung cancer[263]
Asparagus officinalis L.Asparanin AC39H64O13740.9↓Bak/Bcl-xl, ↑ROS, ↑Cyto c, ↓Δψm, ↑caspases, G0/G1 cell cycle arrest, ↓PI3K/Akt/mTORIshikawa cells and xenograft BALB/c miceIC50: 9.34 µMLiver and pancreatic cancer[264]
Salvia miltiorrhizaCryptotanshinoneC19H20O3296.4↑caspase3 and 9, ↑Bax/Bcl-2, ↓MMP 2 and 9A2780 cells5–30 µMLeukemia, prostate, and colon cancer[265]
Colon cancer
Asiatic MoonseedDauricineC38H44N2O6624.8↓cyclin D1, ↓COX2, ↓cMyc, ↓survivin, ↓Bcl-2, ↓IAP1, ↓MMP 9, ↓ICAM1, ↓VEGF, ↓NFκβHCT116, HCT8, SW480 and SW260 cells5–20 µMPancreatic and renal cell carcinoma[266]
Withania somniferaWithaferin AC28H38O6470.6↑ROS, ↓ Bcl-2/Bax, ↑caspase3 and 9, ↓ΔΨmRKO and HCT116 cells0.1–10 µMLung, breast, and pancreatic cancer[267]
Piper nigrumPiperineC17H19NO3285.34↓wnt/β-catenin pathwaySW480 and SW480-pBAR/Renilla, HCT116, DLD1, RKO30–100 µMLung, liver, breast, and brain cancer[268]
Crocus sativusCrocetinC20H24O4328.4↑p53, ↑PIDD, ↑Bax, ↑FAS, ↑caspase-3, -8 and -9HCT116 and HT29 cell lines100 µMProstate, breast, pancreatic, and gastric cancer[269]
Cynanchum paniculatumAntofineC23H25NO3363.4↓proliferation, ↑cytotoxicity, G2/M cycle arrestCol2 and A549 cellsIC50: <9 ng/mLBreast, lung, and renal carcinoma[270]
Zingiber zerumbetZerumboneC15H22O218.33↑ROS, ↓ Bcl-2/Bax, ↑caspase-3/-8/-9, ↓ΔΨm, G2/M cycle arrestSW480 cell lineIC50: 102 µMOral, breast, lung, and prostate cancer[271]
Hymenocallis littoralisPancratistatinC14H15NO8325.27↑LC3II, ↑beclin-1, ↑Bax, ↓cyclin B1, ↓cdc25c, G2/M cycle arrestSW948, DLD1, HTC15, and HT29 cellsIC50: 15–25 µMLymphoma, breast, liver, skin, and teratocarcinoma[272]
Saussurea lappaCostunolideC15H20O2232.32↓Survivin, ↓β-catenin, ↓galectin-3, ↓cyclin D1, G2/M cycle arrestSW480, L-Wnt3a cells0.5–5 µMBreast, prostate, and ovarian cancer[273]
Rehmannia glutinosaCatalpolC15H22O10362.33↓VEGF, ↓EGFR2, ↓HIF-1α, ↓IL-1β, ↓IL-6 and 8, ↓iNOSCT26 cells and xenograft in mice<40 µMGastric, lung, and liver cancer[274]
Laminaria japonicaLJGPn.d.n.d.↓CDK2, ↓PCNA, ↓E2F-1, ↓cyclin E, ↓cyclin D1, ↓PARP, ↑p27, ↑caspase 9, ↓Bcl-2, G1 phase arrestHT-29, HepG2 and AGS cellsIC50: 100 µg/mLLung, liver, and cervical carcinoma[275]
Rosmarinus officinalisRosmarinic acidC18H16O8360.3↑E-cadherin, ↓N-cadherin, ↓twist, ↓vimentin, ↓MMP 2 and 9, ↓ICAM-1, ↓ITGβ1CT26 and HCT116 cells50–200 µMBreast, gastric, leukemia, and cervical cancer[276]
Coptis chinensisBerberineC20H18NO4+336.4↑p21, ↓PARP, ↑caspase 8, ↓VEGF, ↓COX2, ↓Bcl-2, G2/M phase arrestSW480 cells0.5–50 µMProstate, cervical, esophageal, thyroid, and gastric cancer[277]
Sanguinaria canadensisSanguinarineC20H14NO4+332.3↑ROS, ↓MMP, ↑caspase-3, 8 and 9, ↓Bcl-2, ↓XIAP, ↑Egr-1HCT-116 cell line0.3–1.2 µMBreast, prostate, cervical, and pancreatic cancer[278]
Nauclea orientalisNaucleaoral A and BC20H20N2O3336.38↑cytotoxicityHela and KB cellsIC50: 4.0 and 7.8 µg/mLCervical, bladder, and pancreatic cancer[279]
Glycyrrhiza uralensisLicoricidinC26H32O5424.5↑caspase-3, 8 and 9, ↓CDK1, ↑AMPK, ↓Akt/mTOR, G1/S phase arrestSW480 cells and xenograft BALB/c miceIC50: 7.2 µMGastric, lung, prostate, and osteosarcoma[280]
Carpobrotus edulisRutinC27H30O16610.5↑caspase 3, G0/G1 phase arrestHCT116 and HaCaT cells1000 µM
IC50: 679 µM
Breast, lung, and cervical cancer[281]
Scutellaria baicalensisBaicalin and BaicaleinC21H18O11, C15H10O5446.4, 270.24↓hTERT, ↓MAPK, ↓ERK, ↑p38HT-29, SW480 cells and HCT-116 xenograft NSG mice10–150 µMBreast, prostate, and pancreatic cancer[282]
Nicotiana glaucaScopoletinC10H8O4192.17↓ERK1, ↓VEGF-A, ↓FGF-2HUVEC, CCD18-Co and HCT116 Xenograft miceIC50: 0.06 µMBreast, lung, and skin cancer[283]
Morus australisMorusinC25H24O6420.5↓ PI3K/Akt, ↓PDK1, ↓XIAP, ↓cMyc, ↓NFκβ, ↑caspase-3, 8 and 9, G1 phase arrestHT-29 cellsIC50: 6.1–12.7 µMBreast, ovarian, and prostate cancer[284]
Allium sativumDiallyl disulphideC6H10S2146.3↓GSK3β, ↓NFκβSW480 cells and AOM/DSS mouse model2.5–40 µMLung, breast, and gastric cancer[285]
Nandina domesticaProtopineC20H19NO5353.4↑p53, ↑p21, ↑Bax, ↑caspase-3 and 7, ↑LC3-IIHCT116 cells10–40 µMProstate, breast, ovarian, and head and neck cancer[286]
Brain cancer
Nigella sativaThymoquinoneC10H12O2164.2↑LC3-II, ↑p62, ↓MMP 2 and 9, ↓FAC, ↓Nf-kβ, ↓ERK, ↓Akt, ↓mTORT98MG and U87MG cells10–40 µM
IC50: 10.3 µM and 8.3 µM
Breast, liver, colon, and lung cancer[287,288]
Radix Angelica sinensisZ-ligustilideC12H14O2190.24↓RhoA, ↓Cdc42, ↓Rac1T98MG cells2.5–25 µMBreast, prostate, and colon cancer[289]
Panax ginseng C. A. MeyGinsenoside Rh2C36H62O8622.9↑miR128, ↑E2F3a, ↑caspase-3T98MG, A172, and U251 cells12 µg/mLBreast, ovarian, colon, and prostate cancer[290]
Bolbostemma paniculatumTubeimoside-1C65H102O291347.5↓Bcl-2, ↑ Bax, ↑caspase-3, ↑Cyto- C, ↑ROSU251 and U87 cells10–50 µg/mLGastric, liver, ovarian, and lung cancer[291]
Thuja occidentalisα-/β-ThujoneC10H16O152.23↓VEGF, ↓Ang-4, ↓CD31U87-MG and C6 cells xenograft in miceLD50: 400 and 300 µg/mLMelanoma, breast, lung, and colon cancer[292]
Rubia cordifolia L.MolluginC17H16O4284.31↓ Akt, ↓ P70S6K, ↓ mTOR, ↓ ERK ½, ↑JNK, ↑p38U87MG, U251 and MKN45 cells10–40 µMBreast, colon, ovarian, and lung cancer[293]
Garcinia brasiliensis7-epiclusianoneC33H42O4502.7↓cyclin A, ↑caspase-3, S and G2/M phase arrestU251MG and U138MG cell linesIC50: 23 and 18.52 µMColon, melanoma, breast, lung, and ovarian cancer[294]
Escherichia coliSelenocysteineC3H6NO2Se167.06↑ROS, ↑p21waf1/cip1, ↑p53, ↓Akt, ↑p38MAPK, ↑JNK, ↑ERK, S phase cycle arrestU251 and U87 cell lines5–20 µMBreast, lung, and prostate cancer[295]
Anula heleniumAlantolactoneC15H20O2232.32↑ROS, ↓GSH, ↓Bcl-2, ↑ Bax, ↑p53, ↑Cyto- C, ↑caspase-3/-9, ↓ΔΨm, ↓Nf-kβU87, U373 and LN229 cellsIC50: 33–36 µMLung, breast, liver, and pancreatic cancer[296]
Carpesium nepalenseNepalolide AC20H28O6364.4↓IkB-α, ↓IkB-β, ↓iNOS, ↓NF-kβC6 cell line2–10 µMBrain cancer[297]
Buxus microphyllaCyclovirobuxine DC26H46N2O402.7↑ Bax, ↓Bcl-2, ↑caspase-3, S and G0/G1 phase arrestT98G and Hs683 cell lines15–240 µmol/LGastric, colon, breast, and prostate cancer[298]
Euscaphis japonicaPomolic acidC30H48O4472.7↑caspase-3 and 9, ↑ROS, ↓MRP1A172, GBM-1 and U87 cellsIC50: 8.82, 9.72 and 11.09 µg/mLLeukemia, melanoma, gastric and uterine cancer[299]
DanshenSalvianolic Acid BC36H30O16718.6↑ROS, ↑p53, ↑p38 MAPKU87 cell line1–100 µMHNSCC, breast, and colon cancer[300]
Glycine maxSoyasapogenol BC30H50O3458.7↓STAT3, ↓M2 polarization, ↑M1, ↑IL-12U373-MG, SaOS2, and LM8 cells xenograft in mice1–100 µMGastric, breast, colon, and renal cancer[301]
Rosmarinus officinalisCarnosolC20H26O4330.4↑p53, ↓twist, ↓Zeb1, ↓slug, ↑miR-200cU87MG, T98G, U373 MGIC50: <40 µMLymphoma, lung osteosarcoma, and gastric cancer[302,303]
Tillandsia recurvataHLBT-100C19H20O8376.4↑caspase-3, and 7, G1 cell cycle arrest, ↓angiogenesisNCI60 cell lines (U87-cells)GI50 values: <0.100 µM
IC50 for U87: 0.054 µM
Breast, prostate, leukemia, and melanoma[304]
Ligusticum chuanxiong hortTetramethylpyrazineC8H12N2136.19↓CXCR4C6 cell line100 µMBreast, liver, colon, and lung cancer[305]
Garcinia hanburyi Hook. fGambogenic acidC38H46O8630.8↓cyclin E, ↓cyclin D1, ↓EGFR, ↓Akt, ↓GSK3β, Go/G1 phase arrestU251 cell line0.75–6 µMGastric, breast, and lung cancer[306]
Vitis viniferaResveratrolC14H12O3228.24↓Akt, ↑p53U87 and patient derived (22,33 and 44 GSC) cells5–100 µMColon, gastric, and breast cancer[307]
Curcuma aromatica Salisb.GermacroneC15H22O218.33↑p53, ↑ Bax, ↓Bcl-2, ↑p21, ↓CD1, ↓CDK2, G1 phase arrestU87 and U251 cells50–250 µmol/LBreast, prostate, and liver cancer[308]
Acori Graminei RhizomaVolatile Oil (VOA)n.d.n.d.↑caspase-3, 8 and 9, ↑ Bax/Bcl-2, ↑LC3-II/I, ↑atg5, ↑beclin1, ↓p62U87, U251, 3T3 and A172 cells25–250 µg/mLPancreatic and breast cancer[309]
Trichosanthes kirilowii Maxim.Trichosanthin247 Amino acids27 kDa↓LGR5, ↓β-catenin, ↓pGSK-3βSer9, ↓cMyc, ↓CD1U87 and U251 cellsIC50: 40 and 51.6 µMLeukemia and cervical cancer[310]
Streptomyces staurosporeusStaurosporineC28H26N4O3466.5↑caspase-3, ↓TDP-43U87 cell lineIC50: 5µMLung, breast, prostate, and colon cancer[311]
Dysosma versipellisDeoxypodophyllotoxinC22H22O7398.4↓Cdc2, ↓CB1, ↓Cdc25C, ↑caspase-8 and 9, ↓Bcl-2, ↓Bcl-xL, G2/M phase arrestU-87 MG and SF126 cellsIC50: 15.06 and 13.95 nMLung, breast, prostate, and colon cancer[312]
Anemone taipaiensisSaponin BC48H78O17927.1↑Fas-l, ↑caspase-3, ↓Bcl-2, G1/S phase arrestU87MG cellsIC50: 6.7 µmol/LLeukemia and breast cancer[313]
Liver cancer
Strychnos nux-vomica LinnBrucineC23H26N2O4394.5↓HIF-1, ↓MMP-2, ↓FN, ↓LOX, ↓CDSMMC-7721, HepG2, and HCC in Male Kunming mice20–150 µMColon and breast cancer[314]
Azadirachta indicaNimbolideC27H30O7466.5↑caspase-3,7 and 9, ↑ Bax, ↓Bcl-2, ↓Mcl-1, ↓XIAP, ↓ c-IAP1, ↓c-IAP2, G2/M phase arrestPLC/PRF/5 and Huh-7 cells xenograft Balb/c mice1–5 µMPancreatic, breast, lung, and colon cancer[315]
Gardenia jasminoides ellisGeniposideC17H24O10388.4↓miR-224, ↓wnt/βcatenin, ↓AktHepG2 and Huh7 cells100–500 µMBrain, oral, skin, and colon cancer[316]
Solanum nigrum L.SolamargineC45H73NO15868.1↓pcna, ↓Ki67↓Bcl-2, ↑ Bax, ↑caspase-3 and 9, G2/M phase arrestSMMC-7721 and HepG2 cellsIC50: 12.17 and 20 µMColon, lung, prostate, and breast cancer[317]
Matricaria recutitaα- bisabololC15H26O222.37↑caspase-3,-8 and -9, ↑Fas, ↓Bcl-2, ↑p53, ↑Nf-kβ,HepG2, ECa109, PC-3 and Hela cells1–20 µMColon, brain, endometrial, and prostate cancer[318]
Ganoderma lucidumGanoderic acid AC30H44O7516.7↓cyclin D1, ↑p21, ↑cleaved caspase 3, G0/G1 phase arrestHepG2 and SMMC-7721 cellsIC50: 187.6 and 158.9 µmol/LProstate, breast, lung, and meningioma[319]
Thalictrum glandulosissimumHernandezineC39H44N2O7652.8↑AMPK, ↑Atg-7HepG2 and Hep3B cellsIC50: 7.42 and 6.71 µMLung, prostate, breast, and cervical cancer[320]
Angelica gigas NakaiDecursinC19H20O5328.4↑LATS1, ↑βTRCP, ↑p-YAP, ↑cleaved caspase 3, ↑cleaved PARP, G1 phase cell cycle arrestHepG2, Huh-7 cells and tumor xenograft in mice5–80 µMGastric, lung, prostate, and lymphoma[321]
Lavandula officinalisLinaloolC10H18O154.25↓cyclin A, ↓CDK4, ↑p21, ↑p27, ↑ROS, ↑caspase-3, ↓Ras, ↓Akt, ↓mTOR, G0/G1 phase arrestHepG2 cell line0–2.5 mMLeukemia, breast, prostate, and ovarian cancer[322]
Tylophora indicaTylophorineC24H27NO4393.5↓cyclin A2, G1 phase cell cycle arrestHepG2, HONE-1, and NUGC-3 cells2 µMBreast, stomach, nasopharyngeal, and colon cancer[323]
Patrinia scabra BungeLariciresinolC20H24O6360.4↓ΔΨm, ↑Cyto- C, ↑caspase-3 and 9, ↑PARP, ↓Bcl-2/BaxHepG2 cellsIC50: 208 µg/mLLeukemia, breast, and prostate cancer[324]
Oroxylum indicumOroxin BC27H30O15594.5↑PTEN, ↓COX-2, ↓VEGF, ↓p-Akt, ↓PI3KSMMC-77210.34–1.68 µMBreast, lung, and lymphoma[325]
Stephania tetrandraTetrandrineC38H42N2O6622.7↓wnt/β-catenin, ↓MTA1, ↑E-cadherin, ↑occludin, ↓VimentinHuh7, Hep3B and HCCLM9 xenograft Balb/c mice0.5–4 µMColon, esophageal, and pancreatic cancer[326]
Scutellaria laterifloraScutellareinC15H10O6286.24↓HIF-1α, ↓Flt-1, ↓VEGFA, ↓MMP 2 and 9, ↑caspase-3HepG2, MCF-7, EAC, A549 and liver carcinoma and ascites lymphoma model in miceIC50: <13.8 µMGastric, colon, lung, and fibrosarcoma[327]
Toddalia asiatica LamChelerythrineC21H18NO4+348.4↓MMP 2 and 9, ↓p-FAK ↓PI3K, ↓Akt, ↓mTOR, ↓c-JNK, ↓ERKHep3B Cell line0.625–5 µMBreast, prostate, renal, and lung cancer[328]
Astragalus complanatus R.Br.(FAC) flavonoidsn.d.n.d.↑caspase-3 and 8, ↑ Bax, ↑p21, ↑p27, ↓CDK1, CDK4, ↓cyclin B1, ↓cyclin D1, G0/G1 and S phase arrestSMMC-7721 and HepG2 cellsIC50: 48 and 53 µg/mLBreast and nasopharyngeal[329]
Quercus IbericaQuercetinC15H10O7302.23↑E-cadherin, ↓MMP9, ↑LC3, ↓Vimentin, ↓Jak2, ↓STAT3,LM3 cells and xenograft mice model20–200 µMBreast, colon, ovarian, and pancreatic cancer[330]
Trichosanthis RadixCucurbitacin BC32H46O8558.7↓cdc2, ↓cyclin D1, ↓c-Raf, S phase arrestBEL-7402 cells and xenograft miceIC50: 0.32 µMBrest, pancreatic, and laryngeal cancer[331]
Colchium speciosumColchicineC22H25NO6399.4↑AKAP12, ↑TGFB2, ↑MX1, ↓APOH, ↑GDF15, ↑IL32HCC24 and HCC38/KMUH, F28 and F59/KMUH cell lines and Balb/c-nu xenograft2 and 6 ng/mLThyroid, oropharyngeal, and breast cancer[332]
Head and Neck cancer
Wilkstroemia elliptica Merr.UmbelliferoneC9H6O3162.14↑ROS, ↓MMP, G0/G1 cycle arrestHOC KB cellsIC50: 200 µMRenal prostate, lung, and breast cancer[333]
Dioscorea nipponicaDioscinC45H72O16869↑p53, ↓Cyclin A, ↓CDK2, ↓p-ERK, ↓Bcl-2, ↑p-JNK, ↑p-p38, ↑Bax, ↑cleaved caspase-3/-9NP69, Hep-2 and TU-212 cellsIC50: <5 µg/mLGastric, breast, lung, ovarian, and colorectal cancer[334]
Maclura pomiferaOsajinC25H24O5404.5↑Bax, ↓Bcl-2, ↑Fasl, ↑cym-c, ↓GRP78, ↑caspase-3,4,8 and -9TW076, TW04 and CG-1 cellsIC50: 5 µMKidney, prostate, breast, and colon cancer[335]
Cichorium intybusEsculetinC9H6O4178.14↓pJAK1/2, ↑ROS, ↓STAT3, G1/S cycle arrestTU-212, M4e, Hep-2 xenograft in miceIC50: 2.969, 12.88 and 1.958 µMPancreatic, prostate, colon, and lung cancer[336]
Serratia marcescensProdigiosinC20H25N3O323.4↓Cyclin D1, ↑beclin-1, ↓mTOR, ↓PI3K/Akt, G0/G1 phase arrestOECM1 and SAS cell linesIC50: 1.59 and 3.25 µMBreast, gastric, colon, and hematopoietic cancer[337]
Albatrellus confluensNeoalbaconolC22H34O3346.5↓PDK1, ↓PI3K/AKT HK-2, ↑RIP1, ↑RIP3NP69, k562,MCF-7 and A549 and C666-1 xenograft in miceIC50: ≤18 µMLung, breast, colon, and gastric cancer[338]
Haematococcus pluvilisAstaxanthinC40H52O4596.8↓PI3k/Akt, ↓STAT3, ↓Nf-kβ, ↓miR-21, ↓HOTAIRSCC131 and SCC4 cellsIC50: 720 and 700 µMPancreatic, breast, colon, and melanoma[339]
Geranium thunbergiiGeraniinC41H28O27952.6↓MMP2, ↓Fak, ↓Src, ↓ERK ½SCC-9 and SCC-14 cells20–80 µMBrain, breast, colon, ovarian, bladder, and osteosarcoma[340]
Forsythia suspensaPhillygeninC21H24O6372.4↑Caspase-3/-9, ↑Bax, ↑ROS, ↓Bcl-2, G2/M phase arrest, ↓Nf-kβSH-1-V1 cell line and xenograft in miceIC50: 6 µMLung, liver, and pancreatic cancer[341]
Enicosanthellum phalcrumLiriodenineC17H9NO3275.26↑Bax, ↑Caspase-3, ↓Bcl-2, G2/M phase arrestECA-109 cell line0.1–20 µMBreast, prostate, and gastric cancer[342]
Halichondria sp.IlimaquinoneC22H30O4358.5↑ROS, ↑LC3B-II, ↑Atg5, ↓p-pAkt, ↓p-p38, ↓HIF-1α, ↓Mcl-1, ↓Bcl-2SCC4 and SCC2095 cellsIC50: <9 µMProstate, lung, and colon cancer[343]
Tabernaemontana catharinensisCoronaridineC21H26N2O2338.4↑apoptosis, ↑cytotoxicityHep-2 cell lineIC50: 54.47 µg/mLLeukemia, breast, colon, and gastric cancer[344]
Chelidonium majusUkrainC66H75N6O18PS1303.4↓EGFR, ↓AKT2, ↓STAT3, JAK1, ↓β-catenin, ↑CYP1A1, ↑CYP1B1FaDu, HlaC78 cellsEC50: <11 µg/mLLung and prostate cancer[345]
Prostate Cancer
CruciferaeIndole-3-carbinolC9H9NO147.17↓CDK6, ↑Bax, ↓Bcl-2, ↑p21, ↑p27, G1 phase arrestPC-3 cells30–100 µMMelanoma, colon, breast, and endometrial[346]
Genista tinctoriaGenisteinC15H10O5270.24↓HOTAIR, miR-34a↓NFκβ, ↓AktPC3 and Du145 cells25 µMLung and breast cancer[347,348]
Punica granatumEllagic acidC14H6O8302.19↑IL-6, ↓STAT3, ↓Akt, ↓ERKLNCaP and PC-3 cells20–100 µMBreast and ovarian cancer[349,350]
Stephania tetrandraFangchinolineC37H40N2O6608.7↓NR4A1, ↓survivin, ↑ROS, ↑caspase 3 and 8MiaPaCa-2 and Panc-1 cellsIC50: 11.1 and 17 µMGastric, bladder, breast, and colon cancer[351]
Melodinus khasianusKhasuanine AC21H24N2O3352.43↑Caspase-3, ↑p53, ↓Bcl-2PC3 cell lineIC50: 0.45 µMBreast, lung, and colon cancer[352]
Camptotheca acuminateCamptothecinC20H16N2O4348.4↑ROS, ↑c-Myc, ↑sp1, ↑PI3k/Akt, ↑hTERT, G2/M cycle arrestLNCaP cells1–5 µMLeukemia, breast, colon, and lung cancer[353]
Andrographis paniculataAndrographolideC20H30O5350.4↓MMP11, ↑γH2AX, ↑ Caspase-3, and 7, G2/M phase arrestPC3, LNCaP, and 22RV1 SCID orthotopic modelGI50: 26.2, 28.1 and 24.2 µMGlioblastoma, renal, colon, and ovarian cancer[354]
Dioscorea nipponicaDiosgeninC27H42O3414.6↓Bcl-2, ↓beclin-1, ↑caspase-9, ↓ PI3K, ↓Akt, ↓mTOR,DU145 cell lineIC50: 6.757 µg/mLBreast, liver, and gastric cancer[355]
Murraya koenigiiMahanineC23H25NO2347.4↓DNMT1, ↓DNMT3B, ↓PDK1, ↓Akt, ↑RASSF1ALNCaP and PC-3 cells10–20 µMColon, brain, and lung cancer[356]
Quercus petraeaProcyanidinC30H26O13594.5↓ΔΨm, ↑ apoptosis, and necrosisPC-3 cell line100–300 µg/mLBreast, lung, stomach, and colon cancer[357]
Santalum albumα-santalolC15H24O220.35↓Survivin, ↓p-Akt, ↑ Caspase-3, ↑cleaved PARPLNCaP and PC-3 cells20 and 40 µMBreast, colon, and skin cancer[358]
Goniothalamus spp.AltholactoneC13H12O4232.23↓p65, ↓NF-kβ, ↓STAT3, ↓survivin, ↓Bcl-2, ↑Bax, S phase arrestDU145, PC3, and LNCap cells20 and 40 µM
IC50: 38.5 µM
Colon, leukemia, and bladder cancer[359]
Nauclea subditaSubditineC20H15N3O2330.1237↑ROS, ↓MMP, ↑cym-c, ↑caspase-3,7 and 9, ↓Bcl-2LNCaP and PC-3 cellsIC50: <14 µMBreast, lung, and colon cancer[360]
Isodon eriocalyxEriocalyxin BC20H24O5344.4↑cleaved caspase 3 and 8, ↑cleaved PARP, ↑LC3B-II, ↓Akt/mTORPC3 and 22RV1 cellsIC50: <4 µMBreast, colon, lung, and bladder cancer[361]
Garcinia indicaGarcinolC38H50O6602.8↑ Bax/Bcl2, ↑caspase-3 and 9, ↓PARP, ↑PI3K, ↑Akt, ↑mTORPC-3 cells and xenograft mice30 µMBreast, lung, leukemia, and bladder cancer[362]
Allium atroviolaceumTricinC17H14O7330.29↓MiR-21PC-3 cell lineIC50: 117.5 µMBreast, colon, and liver cancer[363]
Sinomenium acutumSinomenineC19H23NO4329.4↓miR-23a, ↓CD1, ↓CDk4, ↓Bcl-2, ↓MMP-2 and 9, ↓PI3K/Akt, ↓JAK/STATPC-3 cells0.25–1 mMGastric, ovarian, breast, and lung cancer[364]
Hematological Cancer
Artemisia annuaDihydroartemisininC15H24O5284.35↓VEGF, ↓ERK1/2RPMI18226 MM cellsIC50: 30.24 µmol/LBrain, melanoma, and ovarian[365]
Artemisia absinthumCardamoninC16H14O4270.28↑ROS, ↑Ca2+, ↑ Caspase-3, -8 and-9, ↓Bcl-2, ↑Bax, ↑ cyto-c, ↑AIF, ↑GRP78, ↑FasL, ↑DAP,WEHI-3 cell line2–10 µMGastric, nasopharyngeal, prostate, and colon cancer[366]
Myristica fragransMyristicinC11H12O3192.21↑Caspase-3, ↑Cyto- C, ↓PARP, ↓ERCC1, ↓RAD50, 51, ↓ATMK562 cellsIC50: 368 µMStomach, lung, and ovarian cancer[367]
Ganggui luhui wanMeisoindigoC17H12N2O2276.29↑Caspase-3,8 and 9, ↑Bax, ↑Cyto- C, ↑PARP, ↓Bcl-2,HL60 cell line20 µmol/LColon, breast, and lung cancer[368]
Cananga odorataSampangineC15H8N2O232.24↑ROS, ↓ΔΨm, ↑Caspase-3, G1 phase arrestHL60 cell line1–20 µMLung, head, and neck cancer[369]
Ambrosia maritomaDamsinC15H20O3248.32↓c-Src, ↓AKT, STAT5, ↓NFkβCEM/ADR5000, CCRF CEM cellsIC50: 4.8 µMColon, breast, and lung cancer[370]
Forsythia suspensaPinoresinolC20H22O6358.4↑p21 WAF1/Cip1, G0/G1 phase arrestHL60, HL60R, and K562 cellsIC50: 8 and 32 µMBreast, prostate, and colon cancer[371]
Danggui longhui wanIndirubinsC16H10N2O2262.26↓c-Src, ↓Abl kinase, ↓Gsk3βKCL-22 and T315I mutant KCL-22 cellsIC50: 0.3–0.9 µMColon, breast, renal, and pancreatic cancer[372]
Arnica chamissonisHelenalinC15H18O4262.3↓NFkβ, ↑ROSJurkat T (J16) and Neo jurkat or Bcl-2 jurkat cells2–20 µMBreast, prostate, and colon cancer[373]
Cinnamosma fragransCapsicodendrinC34H48O10616.7DNA damage and proapoptoticK562 and HL-60 cellsIC50: <0.5 µMBreast, lung, and colon cancer[374]
Cephalotaxus fortuneiHomoharringtonineC29H39NO9545.6↓SP1/TET1/5hmC/FLT3/MYCMA9.3ITD, MA9.3RAS, MONOMAC 6, Kasumi-1 and xenograft in miceIC50: 9.2–36.7 nMColon, breast, and gastric cancer[375]
Papaver somniferumNoscapineC22H23NO7413.4↓Bcl-2, ↓XIAP, ↓Mcl-1, ↓Bcl-xl, ↑Bax, ↓TRAF1, ↓IKK, ↓NFkβ,KBM-5 and U266 cellsIC50: 84.4 and 155 µmol/LColon, breast, lung, and ovarian cancer[376]
Streptomyces cinnamonensisMonensinC36H62O11670.9↓MyB, ↓MyB-NFIBHEK-Luc, HEK-MyB-Luc, and MyB-NFIB ACC cells0.3–3 µMProstate, ovarian, colon, and pancreatic cancer[377]
Brucea antidysenterieaBruceantinC28H36O11548.6↑NOTCH1, ↑HES1, G1 phase arrestMM-CSCs cellsIC50: 77 nMBreast, colon, osteosarcoma, and lung cancer[378]
Aglaia foveolataSilvestrolC34H38O13654.7↓FLT3, ↓miR155, ↓ p65Mv4-11, THP-1, AML blasts from patients and Mv4-11 xenograft murine modelIC50: 2.7, 3.8 and 12 nMLiver, colon, and melanoma[379]
Miscellaneous Cancer
Cynanchum otophyllumPaeoniflorinC23H28O11480.5↑HTRA3, ↑BaxCapan-1 and MIAPaCa-2 cellsIC50: 862.7, and 489.5 µMPancreatic, glioma, lung, and colon cancer[380]
Periploca sepium bungeBaohuoside-IC27H30O10514.5↓β-catenin, ↓survivin, ↓cyclin D1Eca109 cells and Eca109-Luc xenograft in mice12.5–50 µg/mL
IC50:24.8 µg/mL
Esophageal, breast, and leukemia[381]
Ochrosia ellipticaEllipticineC17H14N2246.31↑AIF, ↑Cyto- C, ↑ROS, ↑caspase-3, 7 and 9, ↑ERK, ↑JNK, G2/M cycle arrestRL95-2 cells1–10 µMEndometrial, ovarian, thyroid, and breast cancer[382]
Ilex hainanensisIlexgenin AC30H46O6502.7↓IL-6, ↑TNF-αG0/G1 cycle arrestRAW 264.7 and B16-F10 xenograft in miceIC50: 66.57 and 27.34 µMMelanoma, colon, breast, and cervical cancer[383]
Dysoxylum binectariferumFlavopiridolC21H20ClNO5401.8↑caspase-3, 8 and 9, ↓Mcl-1 ↑cyclin-B1, G2/M cycle arrestKKU-055, -100, -214 cells and KKU-213 xenograft in Balb/c miceIC50: 40–213 nMOsteosarcoma, leukemia, lung, and breast cancer[384]
Crinum asiaticumCrinamineC17H19NO4301.34↓Snail1, ↓vimentin, ↓VEGF-A, ↓CDK4, ↓RHOA, ↓PLK1, ↓BCL2L1, ↓Akt1SiHa and C33a cell linesIC50: 23.52 and 60.89 µMCervical, prostate, gastric, and colon cancer[385]
Alpinia katsumadaiAlpinetinC16H14O4270.28↓Bcl-2, ↓Bcl-xL, ↑ Bax, ↑Cyto- C, ↑XIAP, ↑caspase-3, 8 and 9BxPC-3, PANC1, and AsPC-1 cells20–80 µg/mLHepatoma, leukemia, colon, lung, and breast cancer[386]
Combretum caffrumCombretastatin A4C18H20O5316.3↓N-cadherin, ↓Vimentin, ↓slug, ↓snail1, ↓zeb1, ↓p-PI3K, ↓p-AktTPC1 cell line5 and 10 µMThyroid, breast, colon, and lung cancer[387]
Heliotropium Indicum LinnIndicine N-oxideC15H25NO6315.36↑p53, cleave PARP, cleave DNAHela, MCF-7, PC3, and SiHa cell linesIC50: 46–91 µMColon, breast, prostate, and head and oral cancer[388]
Arnebia euchromaAlkanninC16H16O5288.29↑ROS, ↓ΔΨm, ↑p38 MAPK, ↑JNKMDA-MB-231, HCT116, A549, Huh7, HepG2, MCF cellsIC50: 5–12 µMBreast, colon, lung, liver, and glioma[389]
Artemisia princepsJaceosidinC17H14O7330.29↑pCdc25C, ↑p21, ↑ROS, ↓ERk ½, ↑ATM-Chk1/2, G2/M phase arrestHES, HESC, Hec1, and KLE cell lineIC50: 52.68–147.14 µMEndometrial, ovarian, glioblastoma, and oral cancer[390]
Coccinia grandisKaempferolC15H10O6286.24↑LC3-II, ↓p62, ↓G9a, ↑IRE1-JNK-CHOPAGS, NCI-N87, MKN-74, SNU-216, and SNU-638IC50: 50 µMGastric, colon, breast, prostate, and pancreatic cancer[391]
Daphne odoraDaphnetinC9H6O4178.14↓RhoA, ↓Cdc42LM8 OS cells1–30 µMOsteosarcoma, breast, lung, and colon cancer[392]
Cerbera odollamCerberinC32H48O9576.7↓cMyc, ↑Caspase 3 and 7, ↑ROS ↓Bcl-2, ↓Mcl-1, ↓STAT-3, ↓PLK-1, G2/M phase arrestPANC-1, MIA Paca2, A549, HepG2, HT29 cells and mice modelGI50: <90 nMPancreatic, liver, colon, and breast cancer[393]
Asiatic toadCinobufaginC26H34O6442.5↓Bcl-2, ↓CB1, ↑p21, ↓CDC2, ↑puma, ↑caspase 3, G2/M phase arrestEC-109, Kyse-150, and Kyse-520 cellsIC50: 0.91, 0.66 and 0.62 µMGastric, liver, colon, and melanoma[394]
Astragalus membranaceusSwainsonineC8H15NO3173.21↑E-cadherin,↓N-cadherin, ↓Vimentin, ↓slug, ↓snail1, ↓zeb1, ↓p-PI3K, ↓p-Akt, ↓Twist1EC9706 and 293T cells10–100 µg/mLEsophageal, colon, liver, and glioma[395]
Magnolia officinalisHonokiolC18H18O2266.3↓cyclin D1, ↓CDK2 and 4, ↓Akt/mTOR, ↓p38, ↓ERK, G0/G1 phase arrestWRO, SW579 cells, and ARO cells xenograft in Balb/c miceIC50: 37.7, 19.9, and 36.3 µMThyroid, leukemia, prostate, and colon cancer[396]
PolymethoxyflavoneTangeritinC20H20O7372.4↓ MMP, ↑ Caspase-3, -8 and -9, ↑ Bax, ↑ Bid, ↑ tBid, ↑ p53, ↑ p21/cip1, ↑ Fas and ↑ FasLAGS cell line5–240 µMGastric, colon, and lung cancer[397]
Amoora rohitukaAphaninC35H54O4538.80↓K-Ras, ↓p-AKT, ↓cMyc, ↓cyclin D1, ↓STAT3, G0/G1 cycle arrestHPAF-II, BxPC3, and HPAC cellsIC50: 12.80, 15.68 and 17.26 µMPancreatic, lung, colon, liver, and skin cancer[398]
Eucalyptus globulusCaffeic acidC9H8O4180.16↑ Caspase-1, 3, and 8, G0/G1 cycle arrestSK-Mel-28 cells25–200 µMMelanoma, breast, colon, gastric, and ovarian cancer[399]
Zingiber officinale6-gingerolC17H26O4294.4↑ROS, ↑Bax/Bcl-2, ↑Caspase-3, 9, ↑cym-c, G2/M phase arrestAGS cell lineIC50: 250 µMLiver, lung, breast, and retinoblastoma[400]
Humulus lupulusXanthohumolC21H22O5354.4↑caspase-3, 8 and 9, ↑PARP, ↑p53, ↑AIF, ↓Bcl2, ↓XIAP, S phase cycle arrestCa Ski cell lineIC50: 20 µMCervical, breast, and mammary adenocarcinoma[401]
Trollius chinensisOrientin and VitexinC21H20O11, C21H20O10448.4,
432.4
↑p53, and ↓Bcl2EC-109 cells5–80 µMEsophageal, breast, colon, and lung cancer[402]
Rhodiola roseaSalidrosideC14H20O7300.3↓p-MAPK, ↓p-ERK, ↓p-PI3K, ↓p-AKT, ↓miR99a, ↑p21, ↓Bcl-2GES-1, NU-216 and MGC803 cells0.8–8000 µMGastric and bladder cancer[403]
Syzgium aromaticumEugenolC10H12O2164.2↑ROS, ↑Bax, ↓PCNA, ↓ΔΨm, G2/M cycle arrestSIHA, SK-MEL-28, A2058, MCF-7, MDA-MB-231IC50: <23 µMCervical, breast, colon, and lung cancer[404]
Uapaca togoensisArborinineC16H15NO4285.29↓MMP, ↑ROS, G0/G1 and S phase arrestCCRF-SEM, MDA-MB-231, HCT119, U87 MG, HepG2 cellsIC50: <10 µMLeukemia, brain, breast, and pancreatic cancer[405]
Peganum harmalaHarmineC13H12N2O212.25↑beclin-1, ↑LC3-II, ↓Bcl-2, ↑Bax, ↓Akt/mTOR, ↓p70S6kMGC-803 and SGC-7901 cellsIC50: <59 µMGastric, breast, ovarian, and pancreatic cancer[406]
Artemisia vestitaHispidulinC16H12O6300.26↓VEGFR2, ↓Akt, ↓mTOR, ↓S6 kinase, ↓PI3K, ↓Bcl-2PANC-28, BxPC-3, HUVEC, PANC-1 xenograft BALB/c miceIC50: 20–200 µmol/LPancreatic, glioblastoma, and ovarian cancer[407]
Didemnum cuculiferumVitilevuamideC77H114N14O21S1603.9↓tubulin polymerization, G2/M phase arrestA498, HCT116, A5249, and SK-MEL-5 cells and p388 xenograft in miceLC50: 6–311 nMKidney, colon, prostate, ovarian, and pancreatic cancer[408]
Piper longumPiperlongumineC17H19NO5317.34↑p-elF2α, ↓ΔΨm, ↑ATF4, ↑CHOP, ↓TrxR1SGC-7901, BGC-823, KATO III cell lines0.625–20 µMGastric, breast, lung, and colon cancer[409]
Chrysanthemum partheniumParthenolideC15H20O3248.32↓NFκβ, ↓VEGF, ↓c-jun, ↓c-FosKYSE510, Het-1A, and Eca109 xenograft BALB/c miceIC50: 13.3, 21.54, and 10.3 µMEsophageal carcinoma, colon, and prostate cancer[410]
Solanum nigrum L.DegalactotigoninC50H82O221035.2↓Hedgehog Gli1, ↓GSK3βU2OS, U2OS/MTX, ZOS-M, and ZOS xenograft in miceIC50: 12.91 –31.46 µmol/LOsteosarcoma, colon, and pancreatic cancer[411]
Note: Bold font chemical compounds represent the compounds with good anticancer potential against the selected cancer model (lower IC50 values).
Table 2. Natural products undergoing clinical trials for the prevention/treatment of various cancers.
Table 2. Natural products undergoing clinical trials for the prevention/treatment of various cancers.
CancerBioactivesClass/FamilyCombination
With Other Drugs
Clinical trial StatusResultsPurposeReferences
LungCurcuminPhenolGefitinib/ErlotinibUnknownNot reportedSafety and TolerabilityNCT02321293 *
--LovazaOngoing-PreventionNCT03598309 *
EpigallocatechinCatechinmLDG--TreatmentNCT02577393 *
SulforaphaneOrganosulfur --PreventionNCT03232138 *
GossypolPhenolic aldehydeDTX, CISUnknownNot reportedTreatmentNCT01977209*
BreastATRARetinoidAnastrozoleOngoing--NCT04113863 *
CurcuminPhenol ---NCT03980509 *
--PaclitaxelCompletedNot posted-NCT03072992 *
GenisteinIsoflavoneGemcitabine-EfficaciousTreatment/PreventionNCT00244933 *
7-hydroxystaurosporineStaurosporine derivative -Not postedTreatmentNCT00001444 *
EmodinResin UnknownNot postedObservationalNCT01287468 *
CognutrinFatty acid
+
Phenols
CompletedEfficaciousTreatmentNCT01823991 *
Cervical/OvarianCurcuminPhenolCyclophosphamide, Lansoprazole, Aspirin, Vitamin D, PembrolizumabOngoing--NCT03192059 *
OPT-821SaponinAntigen-KLH conjugate VaccineCompletedEfficaciousTreatmentNCT00857545 *
ColonCurcuminPhenol5-FUOngoing--NCT02724202 *
--IrinotecanCompletedNot postedToxicity and PharmacokineticsNCT01859858 *
--Avastin/FOLFIRICompletedNot reportedTreatmentNCT02439385 *
--FOLFOXCompleted--NCT01490996 *
--CelecoxibUnknown--NCT00295035 *
AndrographolideDiterenoidCapecitabineTerminatedLow accrual rateTreatmentNCT01993472 *
BerberineBenzylisoquinoline alkaloid OngoingNot postedPreventionNCT03281096 *
AquaminMulti-mineral complexCalcium carbonateActive--NCT02647671 *
Cyanidin-3-glucosideAnthocyaninCurcuminUnknown-PreventionNCT01948661 *
Ellagic AcidPhenol Completed-TreatmentNCT01916239 *
SilymarinFlavonolignan CompletedUnknown-NCT03130634 *
BrainChlorogenic acidPhenol -Not postedTreatmentNCT02728349 *
Perillyl alcoholTerpenes Ongoing-TreatmentNCT02704858 *
Head and NeckCombretastatin A4 PhosphatePhenol derivative Completed-TreatmentNCT00060242 *
β-CaroteneTerpenoidα-Tocopherol--PreventionNCT00169845 *
CapsaicinCapsaicinoidRadiation therapy--Supportive careNCT00003610 *
LiverSilybinFlavonoid --TreatmentNCT01129570 *
ProstateCurcuminPhenol --SupportiveNCT01917890 *
--TaxotereTerminatedNo Significant outcome-NCT02095717 *
LycopeneCarotenoidVit D3 and E, Selenium, green tea extractCompletedNot postedTreatmentNCT00844792 *
Polyphenon EpolyphenolEGCG-EfficaciousTreatment/preventionNCT00676780 *
CholecalciferolVitamin D -Effective-NCT00524680 *
HematologicalBioperineAlkaloidCurcumin---NCT00113841 *
PlitidepsinDepsipeptideBortezomib and dexamethasone-Not posted-NCT02100657 *
Bryostatin 1polyketide ---NCT00003171 *
HomoharringtonineAlkaloid ---NCT00006364 *
Skin cancerIngenol MebutateDiterpene ester ---NCT01325688 *
Advanced solid tumorsElisidepsinDepsipeptideErlotinib-Not posted-NCT00884845 *
DiscodermolidePolyketide -unknown-[429]
TongueLuteolinFlavonoid UnknownNot postedTreatmentNCT03288298 *
GI TumorsResveratrolPhenol CompletedNot postedTreatmentNCT01476592 *
PancreaticEtoposidePodophyllotoxin derivativeGemcitabine---NCT00202800 *
* represent the clinical trial number obtained from https://clinicaltrials.gov (accessed on 1 June 2022).
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Naeem, A.; Hu, P.; Yang, M.; Zhang, J.; Liu, Y.; Zhu, W.; Zheng, Q. Natural Products as Anticancer Agents: Current Status and Future Perspectives. Molecules 2022, 27, 8367. https://doi.org/10.3390/molecules27238367

AMA Style

Naeem A, Hu P, Yang M, Zhang J, Liu Y, Zhu W, Zheng Q. Natural Products as Anticancer Agents: Current Status and Future Perspectives. Molecules. 2022; 27(23):8367. https://doi.org/10.3390/molecules27238367

Chicago/Turabian Style

Naeem, Abid, Pengyi Hu, Ming Yang, Jing Zhang, Yali Liu, Weifeng Zhu, and Qin Zheng. 2022. "Natural Products as Anticancer Agents: Current Status and Future Perspectives" Molecules 27, no. 23: 8367. https://doi.org/10.3390/molecules27238367

Article Metrics

Back to TopTop