Next Article in Journal
Call the Eckols: Present and Future Potential Cancer Therapies
Previous Article in Journal
Antiviral Activity and Mechanisms of Seaweeds Bioactive Compounds on Enveloped Viruses—A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Antitumor Potential of Immunomodulatory Natural Products

1
Bio-Organic Chemistry Unit, Institute of Biomolecular Chemistry-CNR, Via Campi Flegrei 34, 80078 Pozzuoli, Italy
2
Department of Biology, University of Naples Federico II, Via Cinthia–Bld. 7, 80126 Napoli, Italy
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Mar. Drugs 2022, 20(6), 386; https://doi.org/10.3390/md20060386
Submission received: 14 April 2022 / Revised: 23 May 2022 / Accepted: 24 May 2022 / Published: 8 June 2022

Abstract

:
Cancer is one of the leading causes of death globally. Anticancer drugs aim to block tumor growth by killing cancerous cells in order to prevent tumor progression and metastasis. Efficient anticancer drugs should also minimize general toxicity towards organs and healthy cells. Tumor growth can also be successfully restrained by targeting and modulating immune response. Cancer immunotherapy is assuming a growing relevance in the fight against cancer and has recently aroused much interest for its wider safety and the capability to complement conventional chemotherapeutic approaches. Natural products are a traditional source of molecules with relevant potential in the pharmacological field. The huge structural diversity of metabolites with low molecular weight (small molecules) from terrestrial and marine organisms has provided lead compounds for the discovery of many modern anticancer drugs. Many natural products combine chemo-protective and immunomodulant activity, thus offering the potential to be used alone or in association with conventional cancer therapy. In this review, we report the natural products known to possess antitumor properties by interaction with immune system, as well as discuss the possible immunomodulatory mechanisms of these molecules.

1. Introduction

Natural products (NPs) have historically been the active constituents of traditional medicines and represent a major source of modern therapeutic agents [1]. According to the World Health Organization (WHO), 75% of humanity relies on natural remedies for health care [2].
Currently, significant revival of interest in NPs as a source for novel drugs is occurring for their matchless structural variety and for the usual small molecular weight that make these molecules particularly suitable for pharmacological development [3]. According to Newman and Cragg, from 1981 to 9/2019, almost 33% of the new approved drugs were NP or natural derivate (ND), and over 22% were nature-inspired synthetic compounds [4]. In cancer research, 40 of the 75 small molecules (53.3%) identified from 1946 to 1980 are NP or ND [4].
Tumor is one of the primary cause of death. It is responsible for approximately 7.6 million deaths in the world every year, with the dramatic prediction of 13.1 million by 2030. NPs have inspired or provided a large fraction of anticancer drugs. Today, more than 60% of anticancer drugs in clinical use originate from NPs derived from plants, marine organisms, and microorganisms [5]. Furthermore, NPs are often molecules that are inaccessible by alternative ways, such as paclitaxel (Taxol), a complex diterpenoid compound that was first isolated from Taxus brevifolia [6] and approved by the Food and Drug Administration (FDA) in 1992. Another example is Trabectedin (ET-743) (Yondelis®, approved in 2015), one of the few marine natural products commercially available, whose tetrahydroisoquinoline core is prepared by semi-synthesis from cyanosafracin B produced by cultures of the bacterium Pseudomonas fluorescens [7,8].
Conventional anticancer chemotherapy has been historically supposed to act through direct killing of tumor cells. Cytotoxic drugs interfere with basic cell functions leading to tumor cell death [9]. However, most of the anticancer drugs showed toxic effect to both cancer and healthy cells. In recent years, the importance of the immune system to combat tumor has been discovered and the interest on the exploitation of natural immunomodulatory substances in combination with usual cancer treatments increases, with the aim to ameliorate the immunological reaction against tumors and decrease the chemotherapy suppressive impact. The defense against cancerous cells comprises a dynamic and orchestrated interplay of innate and acquired immune response. Immunoediting is the phenomenon that regulates the evolution of tumors by the immune system. This dynamic process consists of three phases and starts with immunosurveillance followed by tumor progression and escape. In the first stage, a crucial inflammatory response generated by the immune system is necessary to recognize and eliminate the early originated cancer cells [10]. The main immune cells recruited for tumor defense are the antigen presenting cells (APC), such as dendritic cells (DCs) that play a pivotal role in detecting tumor cells and coordinating tumor eradication (Figure 1). DCs display antigen complexed with major histocompatibility complexes (MHC) and co-stimulatory molecules on their surfaces that allow the interaction with T cells. The release of chemical messengers or inflammatory cytokines such as interferon γ (IFN-γ) and interleukin 12 (IL-12) prime T helper 1 (Th1) response until triggering of cytotoxic T lymphocytes (CTLs). Nevertheless, cancer cells can escape immune control by several strategies shifting into equilibrium condition between cancer expansion and apoptosis. Among the immunotherapy plans, immune checkpoint therapy has taken on great relevance. The PD-1/PD-L1 checkpoint represents a crucial immunosuppressive arbiter in tumor immune microenvironment (TME) and is mainly related with IFN-γ signaling. Proinflammatory cytokines as INF-γ induce PD-L1 expression, an essential process to control the immunological functions. The physiological role of this dynamic process is to prevent an immune response from being too dangerous for healthy cells. Checkpoint inhibitors targeting PD-1/PD-L1 axes are antibodies blocking the “off” signal on T cells, thus favoring T cell activation against tumors. The better results in clinical practice were observed with checkpoint proteins block in melanoma and non-small-cell lung cancer (NSCLC) [11]. Nevertheless, only a fraction of the patients responds to these therapies and there are off-target side effects due to the treatment [12,13].
Although the field of immunotherapy is still growing, the discovery of immunomodulatory NPs able to stimulate the innate immunity cells represent the new frontier for tumor therapy. Zhong et al. have recently reviewed natural products from herbal medicines with immunomodulatory potential as immune checkpoints inhibitors [14]. Other studies have demonstrated that also some conventional chemotherapeutic agents can also affect the immune system by different mechanisms. Some cytotoxic drugs determined immunogenic cell death of cancer cells, resulting in the upregulation of damage-associated molecular patterns (DAMPS) triggering phagocytic activity of debris and promoting DC maturation and initiation of antitumor reaction [9]. The same Paclitaxel has been reported to induce DC activation (through TLR4 and MyD88) in mice [15,16] and recently different proofs indicated that antitumor drugs, as anthracyclines, induced an immunogenic apoptosis favoring the engulfment of apoptotic debris by DCs with cytotoxic CD8+ T cells activation [17]. For these reasons, alternative therapies to cytotoxicity that use the immune system to defeat and eliminate tumor cells are spreading.
Many authors have previously reviewed naturally occurring immunomodulators with antitumor activity [18,19,20,21,22,23]. Immunomodulatory potential and mechanism of action of natural compounds have also been discussed for metabolites from plants and microorganisms specifically [17,18]. However, a comprehensive report is still missing. In this review, we focused on NPs with immunomodulatory and antitumor properties from several sources. Special attention is paid to the discovery and the mechanism of action of immunomodulatory anticancer compounds from marine organisms.

2. Terrestrial Environment

Table 1 collects immunomodulatory natural products with antitumor properties derived from terrestrial source herein reported.

2.1. Plant Compounds

In plants, antitumor and immunomodulating substances are abundant. Numerous phytochemicals displayed promising activity with few side effects in the therapy of several tumors [24].

2.1.1. Terpenes

Many studies reported the potential features and utilization of extracts of terrestrial plants for the development of drugs for several diseases including tumors.
Andrographis paniculata (Acanthaceae), known as “kalmegh”, is a medical plant of Ayurveda (Indian subcontinent) that is extensively cultivated in Asia. From leaves and stems of this plant, active phytochemicals, useful as anti-inflammatory [25], antiviral [26], anticancer [27], and immunostimulatory [28], were extracted. The organic extracts of A. paniculata and a few isolated metabolites displayed growth inhibition and differentiation activity on murine leukemia cells [27]. Andrographolide (1, Figure 2), the most abundant diterpenoid of the A. paniculata showed cytotoxic activity against KB (human epidermoid carcinoma) and P388 (lymphocytic leukemia cells) [28]. Andrographolides also stimulated antigen specific and non-specific immune response in mice [28]. Kumar et al. [29] confirmed the immunomodulatory activity of A. paniculata extracts and their constituents in human immune cells and antitumor activity in human tumor cells. However, between the diterpene compounds that have been isolated from the extract, andrographolide (1), 14-deoxyandrographolide (2), and 14-deoxy-11,12-didehydroandrographolide (3), only 1 significantly inhibited the proliferation of cancer cells with GI50 (concentration required to inhibit the 50% growth) values between 10 and 28 µM [29].
Among active terpenes, zerumbone (4, Figure 2), a sesquiterpenoidic derivative found in the rhizomes of Zingiberaceae plants, showed different pharmacological properties like antioxidant, antibacterial, antipyretic, anti-inflammatory, immunomodulatory, as well as anti-neoplastic [30,31]. This compound exerted its antitumor activity significantly suppressing expansion and metastasis by modulating nuclear factor kappa of activated B cells (NF-kB), protein kinase B (PKB), interleukin 6/janus kinase 2/signal transducer, activator of transcription 3 (IL-6/JAK2/STAT3), and their downstream target proteins. The immunomodulatory activities of zerumbone is suggested to be due to the α, β-unsaturated carbonyl-based part of the molecule [32]. Various in vivo and in vitro experiments proved 4 as a powerful antitumor metabolite [33], showing its impact on the mechanisms of tumor signaling. Studies to investigate the immunomodulatory properties of zerumbone have been mainly focused on mitogen-activated protein kinase (MAPK) and NF-kB pathways, NO production, and inflammation [34]. Zerumbone determined the suppression of different pro-inflammatory mediators as nitric oxide (NO), cyclooxygenase-2 (COX2), prostaglandin E2 (PGE2), and inducible nitric oxide synthase (iNOS) in RAW264.7 macrophages [35]. According to Eguchi et al. [36], this molecule highlighted immunosuppressive activity inhibiting activator protein 1 (AP-1) and NF-kB in the THP-1 cellular line. The effect of Zerumbone on human promyelocytic leukemia cells (HL-60) was reported to be due to the inhibition of 12-O-tetradecanoylphorbol-13-acetate (TPA)-induced super oxide anion generation from NADPH oxidase. Enhanced arrest of G2/M cell cycle was evident with a decreased cyclin B1/CDK1 ratio [37]. This compound was also evaluated on human tumor cells of ovary (Coav-3), breast (MCF-7), and HeLa cells (IC50 value of 20.30, 24.30, and 27.7 μM respectively) in addition to hamster ovaria normal cells [38]. Enhanced level of CASP3 was evident in zerumbone-treated HeLa, associated to cellular morphological features of apoptosis [38]. In addition, Yodkeeree et al. proved that increase of TNF-related apoptosis-inducing ligand (TRAIL) caused apoptosis of HCT116 cells (colon adenocarcinoma) [39]. To date, several pre-clinical trials seem to support zerumbone as promising immunomodulatory and preventive chemotherapeutic molecule. However, clinical studies on this compound have not been reported and further investigations are necessary to establish the therapeutic potential of zerumbone.
The diterpenoid triptolide (5, Figure 2) [40], isolated from Chinese Tripterygium Hook F (TWHF), is reported to inhibit proliferation of mice and human T lymphocytes [41,42] via the block of interleukin 2 (IL-2) production [42]. This compound inhibited several inflammatory mediators such as interleukin 1 (IL-1), tumor necrosis factor α (TNFα), IL-6, interleukin 8 (IL-8), and PGE2. Downregulation of IL-8 expression was connected to the nuclear block of transcriptional activation of NF-KB [43]. Triptolide induces apoptosis of cancer cells and sensitize solid tumor cells toward TNFα-dependent apoptosis, by inhibition of NF-kB [44]. NF-kB is a protein transcription factor that binds DNA, regulates the expression of many biologically important genes such as those encoding inflammatory cytokines, and plays a role in the control of apoptosis. It has been described that triptolide inhibited transcriptional activation of NF-kB but not the DNA binding of NF-kB [42]. At concentrations between 5 and 20 ng/mL, this compound reduced cell viability by 40–70% in several solid tumor cell lines [44]. Triptolide can also induce modifications of the cell cycle checkpoints that enhances apoptosis in cancer cells by blocking p21-mediated growth arrest [45]. Amino acid derivatives of triptolide have been patented as immune modulators and antitumor compounds [46]. Phase II clinical studies of a triptolide analog in patients with refractory pancreatic cancer are available (A Phase II, International Open Label Trial of Minnelide™ in Patients With Refractory Pancreatic Cancer (MinPAC). Available online: https://clinicaltrials.gov/ct2/show/NCT03117920?term=triptolide&cond=cancer&draw=2&rank=2 (accessed on 13 April 2022)).
Carotenoids, tetraterpenoid organic pigments typically found in plants or other photosynthetic organisms such as algae and some species of bacteria, are characterized by different immunomodulatory activities. β-Carotene (6, Figure 2) is the main carotenoid isolated from dietary plants. This substance displayed immune-activating properties by triggering of natural killer cells (NK), increasing the number of leukocytes, CD4/CD8 ratio and upregulating MHC I proteins [47]. The anticancer property of lutein (7, Figure 2), characterized by two hydroxyl groups in the beta carotene end rings, was assessed in murine mammary tumor model. The molecule showed stimulatory impact on IFN-γ mRNA expression, downregulating interleukin 10 (IL-10) in splenocytes [48,49] along with apoptosis induction. According to World Cancer Research Fund/American Institute for Cancer Research (WCRF/AICR), foods containing carotenoids are considered protective against lung tumors [50,51]. The co-administration of carotenoids with traditional cytotoxic agents synergistically enhanced their cytotoxicity in Caco-2 cells, probably by inhibiting the function of the ATP-binding cassette (ABC) transporters. Although carotenoids showed a low cytotoxicity in cells with Multi Drug Resistance (MDR) with IC50 values between 100 and 200 μM, they can increase the cytotoxicity of chemotherapeutic drugs in human MDR1 expressing cells [52]. This observation revealed that carotenoids can be used in combination with anticancer drugs to enhance the efficacy of traditional drugs and reverse multidrug resistance (MDR) in cancer cells.
Other phytochemicals with a remarkable anticancer activity are phytosterols. β-sitosterol (8, Figure 2) favored apoptosis of colon tumor HT-29 and prostate cancer LNCaP cells at concentration of 16 µM and these effects were concomitant with reduction of cellular sphingomyelins and increase of ceramides amounts [53]. Mixtures of sterols and sterolins enhanced the cellular responsiveness of T lymphocytes both in vitro and in vivo experiments [54]. Dietary phytosterols have been recognized as immunomodulatory and chemopreventive compounds. Bouic [54] described the role of phytosterols (mixture of β-sitosterol 8 and its glucoside derivative, Figure 2) on human lymphocyte, along with production of a cytokine profile suggestive of a selective effect on Th1 (IL-2 and Tumor Necrosis Factor γ). Further investigation on β-sitosterol glucosides highlighted enhanced NK cell-dependent cytotoxicity against transformed cell lines [54]. Phytosterols also impact on macrophage activities; β-sitosterol has been reported to reduce NO release by phorbol ester from RAW264.7 macrophages, potentially correlated with the impairment of iNOS levels and NF-kB activation [55]. β-sitosterol also suppressed P388D1/MAB macrophages growth and inflammatory PGE2 secretion [56].

2.1.2. Phenolic Compounds

Natural phenolic acids are well known for the antioxidant, anti-inflammatory and antimutagenic activities, as well as apoptosis inducted cytotoxic properties, and a suppressor effect on cancer attack and angiogenesis [45]. In addition to their phenolic antioxidants (PhOH) capability of transferring electrons and protect against pathogens, flavonoids represent a relevant class playing a crucial role in the regulation of diverse cell activities including production of chemokines for attraction of immune effectors into the tumor environment [15].
Many studies reported that the extract of the plant Scutellaria baicalensis controls cancer proliferation by cell cycle regulation. The low toxic effect toward non-malignant cells made S. baicalensis appealing as a source of novel antitumor drugs. S. baicalensis showed a cytostatic activity on different tumor cell lines in vitro [57] and in vivo in mouse cancer models [58]. The major constituents of S. baicalensis radix are flavonoid derivatives as baicalin, baicalein, wogonin, and wogonin-7-0-glucoronide (Figure 3). Among these flavones, wogonin (9), baicalein (10), and baicalin (11) inhibited the proliferation in many human tumor cell lines between 20 and 200 μM, based on the tumor cells evaluated. A detailed summary of these studies is discussed by Min Li-Weber [59]. Significant decrease in cyclin D1 protein level (regulatory protein of the cell cycle) in wogonin-treated breast tumor cell lines and in baicalein-treated prostate tumor cell lines has been also reported [60]. However, only for baicalein-treated lung cancer cell lines, a concomitant reduction of cyclin D1 and B1 levels was noticed [60]. In flavonoids-treated leukemia cell lines, the cell cycle was reduced at G2/M state [61]. In the case of wogonin or baicalein, the percentage of cells in G1 phase increased while in S phase decreased [62,63]. Lamer-Zarawska et al. [60] highlighted that the antitumor effect of flavones found in root of the S. baicalensis (wogonin, baicalein, and baicalin) was related to reduction of high-level reactive oxygen species (ROS), reduction of inflammatory reactions, and NF-kB activation. Flavonoids as wogonin, baicalein, and baicalin have been shown also to protect tissues from chronic inflammation associated to cancer, both in in vitro and in vivo models [59,64,65,66]. The anti-inflammatory effect of the S. baicalensis flavones is due to the inhibition of cytokines [67] and NO production via down-regulation of several inflammation-associated genes [64], such as iNOS [68], cyclooxygenase (COX) [68,69], and lipooxigenase (LOX) [68]. NO, a highly reactive free radical and its synthase NOS are ubiquitous in malignant tumors and are known to exert both pro- and antitumor effects [64,68,70,71]. Recently, Mengyun Ke et al. [72] showed that baicalein and baicalin flavonoids stimulated the T cell-mediated immune response against tumors through reduction of PD-L1 expression in hepatocellular carcinoma (HCC) and these effects were mediated by STAT3 activity inhibition. Therefore, baicalein and baicalin decreased STAT3 activity, downregulated IFN-γ-induced PD-L1 expression, and subsequently restored T cell sensitivity to kill tumor cells. These findings provide novel insight into the anticancer effects of baicalein and baicalin through which tumor growth was inhibited by PD-L1 downregulation, suggesting that these flavonoids have great potential for clinical treatment.
Fruit of Emblica (Phyllanthus emblica Linnaeus), a euphorbiaceous plant broadly present in subtropical and tropical zones of China, India, Indonesia, and the Malay Peninsula, exhibited antioxidant [73], hypolipidemic [73], and hypoglycemic activities [74], and antimicrobial [75] and anti-inflammatory properties. Recently, in vitro analysis also showed that this fruit was able to mitigate the immunosuppressant activity of chromium in rat lymphocytes [76]. Emblica extracts presented peculiar cytotoxicity to different tumor cell lines in vitro and in vivo [77,78]. Liu et al. [79] investigated immunomodulatory and antitumor activity of six phenolic molecules—geraniin (12), kaempferol 3-β-d-glucopyranoside (13), quercetin 3-β-d-gluco-pyranoside (15), isocorilagin (17), quercetin (16), and kaempferol (14)—found in Emblica fruit on mice splenocytes, human breast tumor lines (MCF-7), and human embryonic lung fibroblast (HELF) [79]. Significant stimulatory activities were detected for geraniin (12) and isocorilagin (17), which displayed major cytotoxicity toward MCF-7. Moreover, it was reported that isocorilagin showed potent cytotoxicity toward HELF, whereas geraniin, quercetin, kaempferol, and the related glycosides displayed minor effect to HELF cells [79].
Persimmon (Diospyros kaki L.) leaves, an Asian traditional medicine agent, containing mainly flavonoids, have been also described for their beneficial pharmacological effects including antitumor [80], hypoglycemic [81], antioxidant [82], and anti-inflammatory [82]. Li Chen et al. [83] identified four main flavone and flavonol glycosides, rutin (18), quercetin (16), kaempferol (14), and myricetin (19), described to have anticancer effect to leukemia, colorectal cancer, neuroblastoma, melanotic melanoma [84], and human prostate cancer cell line [85]. In this study, they found that persimmon leaves promoted serum interleukin 18 (IL-18) amount and NK cell cytotoxicity, suggesting activation of macrophages, and also favoring the NK cell-mediated antitumor response [83].
Analogously, tea plant Camellia sinensis exhibited multiple health benefits [86,87,88] and epidemiologic analysis with association between green tea consuming with chemoprotection to tumor [89]. Epigallocatechin-3-gallate (EGCG) (20) represents the main active ingredient in green tea [90,91]. Several antitumor mechanisms were reported, including apoptosis activation, initiation of cellular growth block, modification of cellular life regulatory proteins, induction of killer caspases, and NF-kB pathway inhibition [89,90]. EGCG attenuated molecular pathways implicated in tumor advance [92] and in modulation of the immune system in murine cancers [93,94,95,96]. Rawangkan et al. [97] hypothesized that EGCG inhibited PD-L1, a checkpoint molecule, increasing anticancer immune response [97].
Resveratrol (21, Figure 3), a non-flavonoid polyphenolic substance present in grapes and red wine, modulates lipoprotein metabolism [98,99], eicosanoid production [100,101,102], lipid oxidation [103], platelet aggregation [100], and inhibits cyclooxygenase-2 (COX-2) [104]. Recently, it was also demonstrated that 21 arrested cell processes related to cancer induction, in vivo expansion and in vitro development of pre-neoplastic lesions in mice mammary glands [105]. In several studies, resveratrol has demonstrated to impede expansion of breast, oral, liver, prostate, and colon tumor cell lines [106,107,108,109]. This activity has been linked to the ability of resveratrol to block citokyne production (IFN-γ, IL-2, TNF-α and IL-12) and the activation of transcription factor NF-kB [110]. This compound also suppressed IL-6, IL-10, and interleukin 1 (IL-1) secretion in human PBMCs with tumor promoter activity [111]. Several clinical trials to investigate potential of resveratrol in cancer disease are available (Resveratrol Clinical Trials. Available online: https://clinicaltrials.gov/ct2/results?cond=Cancer&term=resveratrol&cntry=&state=&city=&dist= (access on 13 April 2022)).

2.2. Therapeutic Antitumor Activity and Natural Compounds from Spices

Natural spices have been the subject of many recent studies on the discovery of antitumor agents. In 2009, Majdalawieh and Carr investigated in vitro immunomodulatory and anticancer activities of black pepper (Piper nigrum) and cardamom (Elettaria cardamomum) [112]. These spices possess great potential in the prevention and treatment of different pathologies, including tumors [24,113,114]. The authors suggested that the anticancer activity of black pepper and cardamom extracts could be due to activation of NK cells [112]. However, the immunomodulatory influence of spices on the secretion of main cytokines by splenocytes and macrophages were not deeply analyzed. In this in vitro analysis, the production of T helper 2 cytokines (Th2), interleukin 4 (IL-4), and IL-10 was hindered by black pepper extracts and favored by cardamom extracts. On the contrary, the secretion of Th1 cytokine IFNγ was increased and decreased by black pepper and cardamom aqueous extracts, respectively, suggesting that these extracts contained compounds capable to promote proliferative signaling pathways in splenocytes. Piperine (22, Figure 4), an alkaloid of black pepper, enhanced murine splenocyte proliferation [115], induced anti-proliferative effects on human colon tumor cells [116], and in vivo antitumor activity [117,118,119,120,121]. Piperine alone did not elicit any cytotoxicity; 50% loss in cell was observed when the alkaloid at 1 μg/mL was added to cadmium [115]. Similarly, eugenol (23, Figure 4), also contained in cardamom, enhanced in vitro lymphocyte expansion [122] and inhibited in vivo tumor formation [123,124,125,126].
Curcumin, a polyphenol isolated from the plant Curcuma longa, called turmeric (24, Figure 4), is a substance widely analyzed for the immunomodulatory and antitumor activities [127,128]. Various preclinical investigations showed antitumor activity of curcumin by modulation of tumor mediators, such as NF-kB, COX-2, LOX, and PKC [129,130]. Curcumin inhibited the accumulation of myeloid-derived suppressor cells (MDSC) and their differentiation and interaction with tumor cells, and induced MDSC differentiation [126,128]. These properties are potential strategies for cancer prevention and therapy. Furthermore, curcumin decreased intra-tumor IL-6 secretion and metastasis production in a breast tumor model [131]. In RAW 264.7 murine macrophages, curcumin reduced the lipopolysaccharides (LPS)-induced inflammatory signals NO, PGE2, and IL-6 by NF-kB deactivation [132]. Moreover, curcumin inhibited the expression and functionality of indoleamine-2,3 dioxygenase (IDO), a main immunosuppressive enzyme in cancer immunology and reduced inflammation-mediated PDL-1 expression [133,134]. Several clinical studies are available on the application of curcumin in oncological diseases (alone or in combination with chemotherapeutic drugs) (Curcumin Clinical Trials. Available online: https://clinicaltrials.gov/ct2/results?cond=Cancer&term=curcumin&cntry=&state=&city=&dist= (accessed on 13 April 2022)).
A main hurdle for the development of curcumin in cancer treatment is its scarce natural availability [135]. To study the role of oxidative activation in protein adduction of curcumin, a stable and oxidizable alkynyl-tagged analog was prepared, the 3′-O-Alkynyl-5′-O-methoxycurcumin, very similar to curcumin in the autoxidation potency (1.6 ± 0.4 µM/min v.s. 4.0 ± 0.4 µM/min) and IC50 for NF-ĸB inhibition (24 v.s. 18 µM) [127].
Although black pepper, cardamom, and curcumin are promising as immunomodulatory and chemopreventive molecules, their molecular mechanisms are still under investigation.

2.3. Macromycetes

Macrofungi are renowned source of macro- and small molecules characterized by different immunomodulatory and antitumor activities, including polysaccharides, glycopeptide, proteoglycans, proteins, and terpenoids [136]. In particular, β-(1→3)-d-glucans along with their peptide analogs and fungal immunomodulatory proteins (FIPs) play an important role in immunomodulatory and anticancer activity. The bioactivity of these molecules is linked to their potential impact toward immune effector cells as hematopoietic stem cells, lymphocytes, macrophages, T cells, DCs, and NK cells involved in the innate and adaptive immunity [136,137]. Even if a considerable quantity of investigations on biological mechanisms of macrofungi macromolecules were published, few works investigated on the anticancer and immunomodulatory effects of terpenoids isolated from this source. Triterpenoids (Figure 5) such as ganoderic acids (25), ganoderenic acids (26), ganodermic acids (27), applanoxidic acids (28), ganoderols (29) lucidone (30), ganodermanontriol (31), and ganodermanondiol (32) have been isolated exclusively from macrofungi and higher fungi (Basidiomycetes) [138]. These molecules show highly oxidized lanostane structures with anti-infectives, cytotoxic, and immunomodulatory activities [100]. Studies on Ganoderma lucidum underlined that the anticancer property of some terpenoids was comparable to that of β-d-glucans in the capability to induce NF-kB pathway and regulate Ras/Erk, c-Myc, cAMP response element-binding protein (CREB) protein and MAPK [139].

2.4. Other Sources

The honeybees secrete the propolis, long used in traditional medicine. Propolis show antiviral, antibacterial, antifungal, anti-inflammatory, hepatoprotective properties, and cytotoxic activity against malignant tumor cell lines [140,141]. This substance also posseses immunomodulatory activity [141,142] and indirect antitumor activity by increasing the host’s defense against neoplastic growth via macrophage activation [143]. Discrepancy in bioactivity of propolis of diverse geographic areas suggested differences of the chemical constituents due to plant variety [144]. It has been reported that propolis from Australia attenuated skin inflammation through immunosuppressant property and reduction of lipid peroxidation in the setting of chronic skin inflammation [145]. It can also suppress the classic photo-immune reaction by cutting down the inflammatory cytokines IL-6 and IL-12 and by overexpression of the anti-inflammatory IL-10 [145].
Caffeic acid phenethyl ester (CAPE) (33, Figure 6) is one of chemical compounds identified from propolis [146] with anti-inflammatory and antitumor properties [147,148,149]. The anticancer activity of CAPE is due to the impairment of DNA synthesis [150], block of transduction signaling growth [151], induction of apoptosis through intrinsic apoptotic pathway and priming of antiangiogenic activity [152,153]. CAPE strongly inhibits T cell expansion [154] and reduces the binding and DNA transcriptional functions of NF-kB and nuclear factor of activated T-cells (NFAT), key transcriptional factors for T cell activation. CAPE inhibits DCs maturation after LPS engagement [155,156]. It has been shown that this compound was able to inhibit the release of IL-2 in a concentration-dependent manner (with an IC50 of approximately 1 µM) [154]. CAPE was cytotoxic to cancer or virally transformed cells [155,157,158] but not to normal cells. Lee et al. [159] found that the molecule exhibited strong antitumor effects in oral cancer cells [159].
A further interesting compound present in propolis is artepillin C (hydroxycinnamic acid derivative) (34, Figure 5), which is reported to kill tumor cells indirectly by amplifying T cell-mediated cytotoxicity and inhibiting NF-kB activity in macrophages [160]. The influence of artepillin C on the reactivity of inflammatory cells was demostrated in vitro model on the macrophage cell line RAW 264.7 and an IC50 of 8.5 (7.8–9.2) μM was estimated. Nevertheless, immunogenic properties of artepillin C were less studied than that of CAPE [160].
Table 1. Terrestrial occurring immunomodulators with antitumor activity.
Table 1. Terrestrial occurring immunomodulators with antitumor activity.
MoleculeType of CompoundSourceTumorImmuno System’s RoleReferences
Andrographolide (1)TerpeneAndrographis paniculataHuman epidermal carcinoma (KB, ED50 1.5 µg/mL); lymphocytic leukemia (P388, ED50 1.0 µg/mL)Stimulate antigen specific and non-specific immune responses in mice[29]
Triptolide (5)TerpeneChinese Tripterygium Hook F (TWHF)Solid tumor cellsApoptosis induced by TNFα, inhibition of NF-kB[44]
Zerumbone (4)TerpeneZingiberaceaeHuman cancer cell lines of the ovary (Coav-3) breast (MCF-7) promyelocytic leukemia (HL-60) and colon adenocarcinoma HCT116Immunosuppressive effects via inhibition of AP-1 and NF-kB[36,37,38,39]
β-Carotene (6); Lutein (7)CarotenoidsPlantLung human cancer,
mammary tumor bearing mice model
Stimulate NK cell activities, increase the number of leukocyte immune cells, CD4/CD8 ratio, and surface expression of MHC I molecules
Stimulation effect on IFN-γ mRNA expression;
suppression of IL-10 in splenocytes
[47,48,49]
β-Sitosterol (8)TerpenePlantHuman cancer cell line of the colon (HT-29) and prostate (LNCaP)stimulated blood lymphocyte proliferation in vitro; enhanced lytic and cytotoxic activities of NK cells[53,161]
Wogonin (9), Baicalein (10), Baicalin (11)FlavonesScutellaria baicalenisBreast, prostate, and lung human cancerActivation of NF-kB factor;
cell cycle regulation
[60]
Geraniin (12),Phenolic compoundsPhyllanthus emblica LinnaeusHuman cancer cell line of the breast (MCF-7) and embryonic fibroblast (HELF)Promoted the level of serum IL-18 and NK cell cytotoxicity, suggesting stimulation of macrophages, thereby upregulating the NK cell-mediated antitumor immune response[79]
Kaempferol 3-β−d-glucopyranoside (13)
Kaempferol (14)
Quercetin 3-β-d-glucopyranoside (15)
Quercetin (16)
Isocorilagin (17)
Ruitin (18)Phenolic compoundsDiospyros kaki L.Leukemia, colorectal, neuroblastoma, melanotic melanoma and prostate human cancer.Increase of IL-18;
Upregulation of NK cells
[83,84,85]
Myricetin (19)
Epigallocatechin-3-gallate (20)Phenolic compoundCamelia sinesisHuman lung cancer cell linesInduction of apoptosis and suppression of NF-kB pathway
Inhibition of PD-L1
[89,90,91,97]
Resveratrol (21)Phennolic compoundGrapes and red wineBrest, oral, liver, prostate and colon human cancerInhibition of citokyne production, (IFN-γ, IL-2, TNF-α and IL-12);
block the activation of transcription factor NF-kB
[104,105,106,110]
Piperine (22)AlkaloidPiper nigrumHuman colon cancer cell linesCytotoxic activity of NK cells;
suppression of the relase of Th2 cytokines IL-4 and IL-10;
enhance murine splenocyte proliferation
[112,115,116,117,121]
Eugenol (23)Phenolic compoundCardamomInhibition of tumor
formation in vivo
Cytotoxic activity of NK cells;
suppression of the relase of Th2 cytokines IL-4 and IL-10
[112,123,124,125,126]
Curcumin (24)Phenolic compoundCurcuma longaHuman breast cancerModulation of NF-kB; reduction of IL-6;
inhibit inflammation-mediated PD-L1 expression
[129,130,131,133,134]
Ganoderic acids (25)Triterpenoid compoundsMacromycetes Act on immune effecter cells such as hematpoietic stem cells, lymphocytes, macrophages, T cells, DCs, and NK cells
Activation NF-kB pathway and modulate Ras/Erk, c-myc, CREB protein and MAPK
[139]
Ganoderenic acids (26)
Ganodermic acids P2 (27)
Applanoxidic acid (28)
Ganoderol A (29)
Lucidone (30)
Ganodermanontriol (31)
Ganodermanondiol (32)
Caffeic acid phenethyl ester (CAPE, 33)Phenolic compoundsPropolisOral human cancer and human cancer cell lines of the promyelocytic leukemia (HL-60)Inhibition of T cell receptor-mediated T cell proliferation[150,151,154,158,159]
Artepilin C (34)
Abbreviations: AP-1, activator protein 1; NF-kB, nuclear factor kappa of activated B cells; CD, cluster of differentiation; MHC, Major histocompatibility complex; IFN-γ, interferon-gamma; IL-2, interleukin 2; IL-4, interleukin 4; IL-6, interleukin 6; IL-10, interleukin 10; IL-12, interleukin 12; IL-18, interleukin 18; NK, natural killer; TNFα, tumor necrosis factor α; PD-L1, Programmed death-ligand 1; Th2, T helper 2; DC, dendridic cell; CREB, cAMP response element-binding protein; MAPK, mitogen-activated protein kinase; ED50, Median Effective Dose.

3. Compounds from Marine Environment

The sea, covering a large part of Earth surface, is a basin of huge biological diversity, with more than three hundred thousand species of living organisms described until now [162,163]. The hostile and competitive marine habitat makes these organisms a great depot of molecules with biological properties undiscosed in terrestrial environments. In Table 2, were summarized MNPs with immunomodulatory and antitumor properties herein cited.
Marine macroalgae (especially brown algae) contain a relevant amount of soluble biologically active polysaccharides, such as alginates and fucoidans, and small molecules (tripeptides, phlorotannins, glycolipids, and carotenoids) that have a potential function as dietary fiber and are also used as immune activators [164]. Mekabu is an edible alga exerting anticancer and immunogenic effects [164]. In vivo studies of activity of fucoidans found in Mekabu highlighted the reduction of the tumor growth through Th1 and NK cell responses [165]. Carrageenans, a family of sulfated galactans isolated from marine red algae, exhibit significant antitumor and immunomodulatory activities at different extent [164]. Several studies showed that the molecular weight (MW) of these polysaccharides affect their activity [166,167]. The highest inhibition rate on sarcoma S180 and hepatoma H22 were recorded with MW between 9.3 and 15 kDa, with 66.15 and 68.97% inhibition at the dose of 200 mg kg−1 per day, respectively. The antitumor activities of these compounds are not due to cytotoxic effects, thus the authors put forward the activation of body immunocompetence to explain the results.
From the marine dinoflagellate Alexandrium minutum has been identified a glycopeptide (Table 2) that induces mitophagic cell death in the cancer cell line without affecting normal cell line viability [162]. This mechanism causes the lysosomal secretion of ATP, which stimulates myeloid cells and can induce immunogenic cell death [21].
Similarly, the polyunsaturated short-chain aldehydes produced by Thalassiosira rotula, Skeletonema costatum, and Pseudonitzschia delicatissima [167,168,169,170,171] (Table 2), cause specific programmed cell death in lung and colon adenocarcinoma, inducing the release of ATP and other immune signals, which are known as ICD inducers [21].
The depsipeptide coibamide A (35, Figure 7), isolated from the marine cyanobacterium Leptolyngbya sp. induces caspase-independent cell death in breast cancer cells with EC50 at nanomolar level [168]. As a consequence of severe lysosome defects, the autophagosome-lysosome fusion was blocked upon treatment with coibamide A. Autophagy is responsible for the release of DAMPS [21].
Dioxinodehydroeckol (36, Figure 7), phlorotannin isolated from seaweed Ecklonia cava, exhibited a remarkable antiproliferative effect on human breast cancer cells (MCF-7). This activity has been associated with the induction of apoptosis through activation of the transcription factors of the NF-kB family [169].
Astaxanthin (37, Figure 7) is a carotenoids with chemopreventive activity found abundantly in seaweeds [170]. The molecule improves antitumor immune responses by inhibiting lipid peroxidation induced by stress [170]. Astaxanthin protected mice from carcinogenesis of the urinary bladder by reducing the incidence of chemically induced bladder carcinoma [171]. Dietary astaxanthin also exerted antitumoral activity in the post-initiation phase of carcinogen-induced colon [172] and oral cancer models [173].
Sponges represent a primary source of bioactive marine substances. Galactosylceramide (α-GalCer) (38) was the first active marine glycolipid derived from sponges (not occurring in mammalian cells). This lipid stimulates NKT cells to produce both Th1 and Th2 cytokines and shows antitumor effects in mice [174,175]. The antitumor properties are mediated by CD1d-restricted iNKT cells that activate NK and antigen presenting cells (APCs) to stimulate antitumor immune responses [174,175]. Clinical trials to evaluate the therapeutical application have been already completed (Galactosylceramide Clinal Trials. Available online: https://clinicaltrials.gov/ct2/results?cond=&term=Galactosylceramide+&cntry=&state=&city=&dist= (accessed on 13 April 2022)).
Screening of marine extracts on monocyte-derived human dendritic cells (hMo-DCs) provided positive results for α-sulfoquinovosildiacylglycerols (SQDGs 39, Figure 7), plastidial sulfolipids occurring in algae and other photosynthetic organisms [176]. Natural SQDGs showed moderate activity but a rational change of their chemical structure led to the synthetic analogues β-sulfoquinovosildiacylglycerols (SULF A 40, Figure 7) wich induces DCs maturation and triggers in vivo immune resposte. The product was tested as adjuvant in an experimental model of tumor vaccine against a murine B16F10 melanoma cell line with very encouraging results [176]. Recent studies on the mechanism of action indicated that SULF A is a ligand of the Triggering Receptor Expressed on Myeloid cells-2 (TREM2) [177]. The TREM2-iduced response is mediated by SYK-NFAT axis and is compromised by blockade and gene silencing of the receptor. Activation by this lipid preserved the DC functions to excite the allogeneic T cell response and promoted interleukin-10 (IL-10) release after lipopolysaccharide (LPS) stimulation [177].
A few successful antitumor compounds derived from tunicates and ascidians have made significant progress through clinical trials in the USA and Europe. Didemnins are bioactive depsipeptides, first isolated from the Caribbean tunicate Trididemnum solidum, exhibited various biological properties [178]. Didemnin B (41, Figure 7) shows cytotoxic activity against L1210 murine leukemia cells at very low concentrations, as well as has an immunosuppressant activity through inhibition of the lymphocyte activation at concentrations of 10 pg/mL (IC50) [179,180]. In 2011, Tsukimoto et al. reported the identification of the molecule from the marine α-proteobacteria Tistrella mobilis and Tistrella bauzanensis [181] that uses a unique post-assembly line maturation process [182]. Didemnin B was the first marine-derived product to advance into clinical trials (phase I and phase II) in the early 1980s. Due to several side effects for an unpredicted toxicity and short half-life [179], these trials were officially terminated in the middle 1990s. Recently anticancer marine peptides have been reviewed by Zhang et al. [183]. Cyclic peptides have significant structural advantages, displaying a large surface area, which provides a high affinity and selectivity for protein targets. Thus, they are simple to modify, handle, and characterize, all essential properties for therapeutics [183]. Aplidin® (dehydrodidemnin B), a second generation didemnin isolated from the Mediterranean tunicate Aplidium albicans, was more effective and less toxic than Didemnin B in clinical trials [184]. The molecule was granted “orphan drug status” in the European Union for acute myeloblastic leukemia. Recent studies on the treatment of adults with SARS-COV-2 requiring hospitalization suggest that, in addition to its antiviral effect [185], dehydrodidemnin B could stimulate the immune response against the virus.
Lissoclibadin 2 (42, Figure 7), a trimeric compound isolated from the tunicate Lissoclinum cf. badium ahowed inhibitory activity, efficient against human colon tumor cells DLD-1 and HCT116, breast tumor cells MDA-MB-231, renal tumor cells ACHN, and non-small-cell lung tumor cells NCI-H460. Moreover, the molecule increases IL-8 production, indicating a link between cancer-killing and the immune system [186,187]. The benzoquinone derivative (43, Figure 7), found in Aplidium glabrum and its synthetic analogues (44, Figure 7), induced apoptosis in JB6 CI41 tumor cell lines and inhibited p53 while increasing AP-1 and NF-kB transcription. Inhibition of cellular modification was shown to strictly depend on terpenoid side chain length [186,188].
Lately, we have identified by bioassay-guided fractionation the alkaloid lepadin A (45) from the tunicate Clavelina lepadiformis sp. B, already known to exhibit significant in vitro cytotoxicity against several human cancer cell lines, as a potent activator of innate immune cells [189]. In this study, lepadin A showed both cytotoxic effect against chronic forms of lung carcinoma, melanoma, and multiple myeloma cells together with maturation of mouse dendritic cells at micromolar concentrations. The combination of the two effects is expected to increase the anticancer properties by a synergistic mechanism deriving from both the role of mature DCs in the generation of antitumor activity and the release of immunogenic molecules by dying cells. In particular, the marine alkaloid triggers a significant over-expression of MHC-II and co-stimulatory molecules that are key signals for naïve T cell differentiation by DCs and for mounting an effective immune response. Moreover, the immune response of 45 occurs at subtoxic concentrations (EC50 = 1.64 µg/mL; IC50 = 4.20 µg/mL), indicating that the molecule can induce a cell stress. On this basis, even though further studies are needed to investigate the mechanism of action, it was suggested that lepadin A could act as ICD inducer [189].
Although their constituents and peculiar activities are still to be studied, the extracts from marine organisms are often tested for immunostimulatory activity together withtumor prophylaxis and therapy. For example, in vitro and in vivo antitumor and immunomodulatory activities of the ethanolic extract of the simple ascidian Microcosmus exasperates [190] and Phallusia nigra [191] have been reported. Sea cucumbers metabolites have been also proposed as potent anticancer agents and raw extracts of these marine organisms can suppress inflammation and increased innate immune responses [192]. Frondanols (sulfated triterpenoid glycoside), frondosides (sulfated glycoside), eicosapentaenoic acid, 12-methyltetradecanoic acid, and fucosylated chondroitin sulfate, as well as canthaxanthin/astaxanthin (detected in little amount in the extracts), are the main bioactive metabolites of sea cucumbers with immunomodulatory properties, ameliorating immune reaction by modulation of innate immune cells [192].
Bryostatin-1 (46, Figure 7), a macrocyclic lactone of the marine Bryozoan, Bugula neritina [193], show antitumor and immunomodulatory activity and it is currently in clinical trial [194]. The molecule is a potent activator of PKC class, lacks cancer-inducing efficacy, and shows antagonistic activity on cancer-inducing phorbol esters. This activity is presumably linked to PKC down-regulation or to peculiar isoform activation. The molecule also induced cytokines secretion, bone marrow progenitor cells, and neutrophils [195,196,197]. Increase in IL-2-promoted proliferative response in PBLs was detected, but not upregulation of IL-6 or tumor necrosis factor (TNF). In vitro, 46 displayed cytotoxic effects toward several leukemia and solid tumor lines [196]. It has also in vivo antitumor activity in different mouse models, as leukemia, lymphoma, ovarian cancer, and melanoma. The potential of Bryostatin-1 on several tumors has been reported by many clinical studies (Bryostatin-1 Clinical Trials. Available online: https://clinicaltrials.gov/ct2/results?cond=cancer&term=Bryostatin-1+&cntry=&state=&city=&dist= (accessed on 13 April 2022)).
Trabectidin, a tetrahydroisoquinoline alkaloid that was initially isolated from the Caribbean tunicate Ecteinascida turbinata with the name of ET-743 [198,199], has been the first marine-derived anti-neoplastic drug approved for the treatment of advanced soft tissue sarcoma and, in combination with pegylated liposomal doxorubicin, for the treatment of patients with relapsed platinum-sensitive ovarian cancer. The alkaloid besides the cytotoxic effect has immunomodulatory activity on several cell types of the microenvironment [200]. Several studies have also underlined that Trabectedin is an immunomodulatory drug with potential use in enhancing the therapeutic response to checkpoint inhibitor-based immunotherapy and in overcoming chemoimmune resistance [200,201]
Table 2. Marine occurring immunomodulators with antitumor activity.
Table 2. Marine occurring immunomodulators with antitumor activity.
MoleculeSourceTumorImmuno System’s RoleReferences
GlycopeptideAlexandrium minutumA549 Lung adenocarcinoma cell lineMitophagy and ICD inducer[202]
Polyunsaturated aldehydesdiatomsProgrammed cell death in lung and colon adenocarcinomaInduce the release of ATP and others immune signals which are known as ICD inducers[21,163,202,203,204,205]
Coibamida A (35)Leptolyngbya sp.Breast camcerCaspase-independent cell death and ICD inducer[168]
Dioxinodehydroeckol (36)Ecklonia cavaHuman cancer cell line of the breast (MCF-7)Induction of apoptosis through NF-kB family and NF-kB-dependent pathway[169]
Astaxanthin (37)SeaweedsAntitumoral activity in the post-initiation phase of carcinogen-induced colon and oral cancer modelsImproves antitumor immune responses by inhibiting lipid peroxidation induced by stress[170,171,172,173]
α Galactosylceramide (38)SpongeAntitumor effects in miceStimulation of NKT cells to produce both Th1 and Th2 cytokines[174]
α-Sulfoquinovosides (39)Marine microalgaeSynthetic β-sulfoquinovosides derivative as adjuvant in vaccine against a murine B16F10 melanoma cell lineMaturation of human DCs.[176,177]
Didemin B (41)Trididemnum solidum Inhibition of lymphocyte activation[181,182,183]
Lissoclibadin 2 (42)Lissoclinum cf. badiumHuman colon cancer lines (DLD-1) and (HCT116),
breast cancer lines MDA-MB-231,
renal cancer line ACHN;
non-small-cell lung cancer line NCI-H460
Increase of IL-8 production[187]
2,3-Dimethoxy-5-(3′,7′-dimethyl-octa-20(E),6′-dienyl)-[1,4] benzoquinone (43)Aplidium glabrumJB6 CI41 cancer cellInhibition of p53;
Increase transcription of AP-1 and NF-kB
[186,188]
Lepadin A (45)Clavelina lepadiformis sp. BHuman lung carcinoma, melanoma, and multiple myelomaMouse DCs[189]
Bryostatin 1 (46)Bugula neritinaAntitumor activity against leukemia, lymphoma ovarian cancer, and melanomaActivation of PKC family;
Stimulation of cytokine production
[193,194]
Abbreviations: AP-1, activator protein 1; NF-kB, nuclear factor kappa of activated B cells; CD, cluster of differentiation; IL-8, interleukin 8; NK, natural killer; Th1, T helper 1; Th2, T helper 2; DC, dendridic cell; PCK, Protein kinase C.

4. Conclusions: Future Prospects of Natural Compounds as Potential Anti-Cancer Agents

Cancer is one of the most challenging medical conditions and requires continuous updating of therapeutic approach. Many natural products have been tested for their anticancer potential and some of these compounds have successfully advanced in clinical trials. Even today, natural molecules continue to offer novel chemicals with a wide range of mechanisms of action to help the world fight cancer.
However, due to the serious collateral effects of the traditional cancer chemotherapy, immunotherapeutic treatments are becoming increasingly widespread. The goal of immunotherapy is to fight cancer by stimulating the immune system, our body’s natural defense system. Immune system cells usually activate themselves against anything they recognize as non self, for example pathogens or mutated cells in order to eliminate them. Regrettably, tumor cells use a number of mechanisms to evade this control, increase in number and spread out in the body. Cancer immunotherapy has demonstrated to prevent these self-masking mechanisms of tumor cells, so that the immune system is no longer cheated. The key breakthrough in the development of cancer immunotherapy has been the discovery of immune checkpoints as molecules involved in the tumor immune evasion. Very important therapeutic results have already been achieved with checkpoint inhibitors in melanoma, non-small-cell lung carcinoma, and urothelial cancer [20].
Immunomodulatory compounds and antitumor molecules of natural origin have attracted great interest for the potential to promote the activation and recruitment of immune cells as macrophage, neutrophils, dendritic cells, NKs, and T lymphocytes lead to infiltration in the cancer microenvironment with effector cells facilitating the cancer eradication. In addition, natural products with antioxidant and anti-inflammatory activity, are used to potentiate the immune response and to prevent induction and formation of tumors.
Most of these studies are related to natural compounds from terrestrial sources. However, micro and macro marine organisms are receiving increasing attention in this context and represent a promising platform for the development of new candidates for anticancer immunotherapy.

Author Contributions

Conceptualization, G.N., E.M., G.d. and A.F.; data curation, writing—original draft preparation, G.N., G.S., C.G., F.A. and L.R.; writing—review and editing, E.M. and A.F. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the project “Antitumor Drugs and Vaccines from the Sea (ADViSE)” project (CUP B43D18000240007–SURF 17061BP000000011) funded by POR Campania FESR 2014–2020 “Technology Platform for Therapeutic Strategies against Cancer”—Action 1.1.2 and 1.2.2.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

AP-1: activator protein 1; APCs, antigen presenting-cell; CASP3, caspase-3; CTX, Cyclophosphamide; COX, cyclooxygenase; COX-2, cyclooxygenase-2; CREB, cAMP response element-binding protein; CTL, cytotoxic T lymphocyte; DC, dendridic cell; Fips, fungal immunomodulatory proteins; hMO-DCs, monocyte-derived human dendritic cells; IAPs, inhibitor of apoptosis proteins; IFN-γ, interferon γ; IL-1, interleukin 1; IL-2, interleukin 2; IL-4, interleukin 4; IL-6, interleukin 6; IL-8, interleukin 8; IL-10, interleukin 10; IL-12, interleukin 12; IL-18, interleukin 18; iNOS, inducible nitric oxide synthase; IDO, indoleamine-2,3 dioxygenase; JAK2, janus kinase 2; LOX, lipooxigenase; LPS, lipopolysaccharides; MAPK, mitogen-activated protein kinase; MAPK/PKB, mitogen-activated protein kinase/protein kinase B; MDSC, myeloid-derived suppressor cells; CNP, natural product; ND, natural derivate; NK, natural killer; NF-kB, nuclear factor kappa of activated B cells; NO, nitric oxide; NFAT, nuclear factor of activated T-cells; NSCLC, non-small-cell lung cancer; PGE2, prostaglandin E2; PLF, flavonoids extract from persimmon leaves; PK, protein kinase; PKB, protein kinase B; STAT3, signal transducer and activator of transcription 3; TME, tumor immune microenvironment; TPA, tissue polypeptide antigen; TNF, tumor necrosis factor; TNFα, tumor necrosis factor α; TNFc, tumor necrosis factor c; TRAIL, TNF-related apoptosis-inducing ligand; TH, T helper; TH1, T helper 1; TH2, T helper 2; WHO, world health organization.

References

  1. Thomford, N.E.; Senthebane, D.A.; Rowe, A.; Munro, D.; Seele, P.; Maroyi, A.; Dzobo, K. Natural Products for Drug Discovery in the 21st Century: Innovations for Novel Drug Discovery. Int. J. Mol. Sci. 2018, 19, 1578. [Google Scholar] [CrossRef] [Green Version]
  2. Reid, R.G.; Sarker, S.D. Isolation of Natural Products by Low-Pressure Column Chromatography. Methods Mol. Biol. 2012, 864, 155–187. [Google Scholar] [CrossRef]
  3. Sticher, O. Natural Product Isolation. Nat. Prod. Rep. 2008, 25, 517–554. [Google Scholar] [CrossRef]
  4. Newman, D.J.; Cragg, G.M. Natural Products as Sources of New Drugs over the Nearly Four Decades from 01/1981 to 09/2019. J. Nat. Prod. 2020, 83, 770–803. [Google Scholar] [CrossRef]
  5. Rayan, A.; Raiyn, J.; Falah, M. Nature Is the Best Source of Anticancer Drugs: Indexing Natural Products for Their Anticancer Bioactivity. PLoS ONE 2017, 12, e0187925. [Google Scholar] [CrossRef]
  6. Heinig, U.; Jennewein, S. Taxol: A complex diterpenoid natural product with an evolutionarily obscure origin. Afr. J. Biotechnol. 2009, 8, 1370–1385. [Google Scholar]
  7. Cuevas, C.; Francesch, A. Development of Yondelis® (Trabectedin, ET-743). A Semisynthetic Process Solves the Supply Problem. Nat. Prod. Rep. 2009, 26, 322–337. [Google Scholar] [CrossRef]
  8. Pommier, Y.; Kohlhagen, G.; Bailly, C.; Waring, M.; Mazumder, A.; Kohn, K.W. DNA Sequence-and Structure-Selective Alkylation of Guanine N2 in the DNA Minor Groove by Ecteinascidin 743, a Potent Antitumor Compound from the Caribbean Tunicate Ecteinascidia Turbinata. Biochemistry 1996, 35, 13303–13309. [Google Scholar] [CrossRef]
  9. Bracci, L.; Schiavoni, G.; Sistigu, A.; Belardelli, F. Immune-Based Mechanisms of Cytotoxic Chemotherapy: Implications for the Design of Novel and Rationale-Based Combined Treatments against Cancer. Cell Death Differ. 2014, 21, 15–25. [Google Scholar] [CrossRef] [Green Version]
  10. Norling, L.V.; Serhan, C.N. Profiling in Resolving Inflammatory Exudates Identifies Novel Anti-Inflammatory and pro-Resolving Mediators and Signals for Termination. J. Intern. Med. 2010, 268, 15–24. [Google Scholar] [CrossRef]
  11. Zhu, H.F.; Li, Y. Small-Molecule Targets in Tumor Immunotherapy. Nat. Prod. Bioprospec. 2018, 8, 297–301. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Ngamkham, S.; Holden, J.E.; Smith, E.L. A Systematic Review: Mindfulness Intervention for Cancer-Related Pain. Asia-Pac. J. Oncol. Nurs. 2019, 6, 161–169. [Google Scholar] [CrossRef] [PubMed]
  13. Yoest, J. Clinical Features, Predictive Correlates, and Pathophysiology of Immune-Related Adverse Events in Immune Checkpoint Inhibitor Treatments in Cancer: A Short Review. Immuno. Targets Ther. 2017, 6, 73–82. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Zhong, Z.; Vong, C.T.; Chen, F.; Tan, H.; Zhang, C.; Wang, N.; Cui, L.; Wang, Y.; Feng, Y. Immunomodulatory Potential of Natural Products from Herbal Medicines as Immune Checkpoints Inhibitors: Helping to Fight against Cancer via Multiple Targets. Med. Res. Rev. 2022, 42, 1246–1279. [Google Scholar] [CrossRef] [PubMed]
  15. Zitvogel, L.; Apetoh, L.; Ghiringhelli, F.; Kroemer, G. Immunological Aspects of Cancer Chemotherapy. Nat. Rev. Immunol. 2008, 8, 59–73. [Google Scholar] [CrossRef]
  16. Galluzzi, L.; Senovilla, L.; Zitvogel, L.; Kroemer, G. The Secret Ally: Immunostimulation by Anticancer Drugs. Nat. Rev. Drug Discov. 2012, 11, 215–233. [Google Scholar] [CrossRef]
  17. Kawano, M.; Tanaka, K.; Itonaga, I.; Iwasaki, T.; Miyazaki, M.; Ikeda, S.; Tsumura, H. Dendritic Cells Combined with Doxorubicin Induces Immunogenic Cell Death and Exhibits Antitumor Effects for Osteosarcoma. Oncol. Lett. 2016, 11, 2169–2175. [Google Scholar] [CrossRef] [Green Version]
  18. Pan, P.; Huang, Y.W.; Oshima, K.; Yearsley, M.; Zhang, J.; Arnold, M.; Yu, J.; Wang, L.S. The Immunomodulatory Potential of Natural Compounds in Tumor-Bearing Mice and Humans. Crit. Rev. Food Sci. Nutr. 2019, 59, 992–1007. [Google Scholar] [CrossRef]
  19. Mohamed, S.I.A.; Jantan, I.; Haque, M.A. Naturally Occurring Immunomodulators with Antitumor Activity: An Insight on Their Mechanisms of Action. Int. Immunopharmacol. 2017, 50, 291–304. [Google Scholar] [CrossRef]
  20. Moody, R.; Wilson, K.; Jaworowski, A.; Plebanski, M. Natural Compounds with Potential to Modulate Cancer Therapies and Self-Reactive Immune Cells. Cancers 2020, 12, 673. [Google Scholar] [CrossRef] [Green Version]
  21. Sansone, C.; Bruno, A.; Piscitelli, C.; Baci, D.; Fontana, A.; Brunet, C.; Noonan, D.M.; Albini, A. Natural Compounds of Marine Origin as Inducers of Immunogenic Cell Death (ICD): Potential Role for Cancer Interception and Therapy. Cells 2021, 10, 1–20. [Google Scholar] [CrossRef]
  22. Katanaev, V.L.; di Falco, S.; Khotimchenko, Y. The Anticancer Drug Discovery Potential of Marine Invertebrates from Russian Pacific. Marine Drugs 2019, 17, 474. [Google Scholar] [CrossRef] [Green Version]
  23. Riccio, G.; Lauritano, C. Microalgae with Immunomodulatory Activities. Marine Drugs 2020, 18, 2. [Google Scholar] [CrossRef] [Green Version]
  24. Aggarwal, B.B.; Kunnumakkara, A.B.; Harlkumar, K.B.; Tharakan, S.T.; Sung, B.; Anand, P. Potential of Spice-Derived Phytochemicals for Cancer Prevention. Planta Med. 2008, 74, 1560–1569. [Google Scholar] [CrossRef] [Green Version]
  25. Shen, Y.-C.; Chen, C.-F.; Chiou, W.-F. Andrographolide Prevents Oxygen Radical Production by Human Neutrophils: Possible Mechanism(s) Involved in Its Anti-Intinflammatory Effect. Br. J. Pharmacol. 2002, 135, 399–406. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Shihman Chang, R.; Ding, L.; Gai-qing, C.; Ze-lin, Z.; SMITHt, K.M. Dehydroandrographolide Succinic Acid Monoester as an Inhibitor against the Human Immunodeficiency Virus (43225). Proc. Soc. Exp. Biol. Med. 1991, 197, 59–66. [Google Scholar] [CrossRef]
  27. Hossain, S.; Urbi, Z.; Karuniawati, H.; Mohiuddin, R.B.; Moh Qrimida, A.; Allzrag, A.M.M.; Ming, L.C.; Pagano, E.; Capasso, R. Andrographis paniculata (Burm. f.) Wall. ex Nees: An Updated Review of Phytochemistry, Antimicrobial Pharmacology, and Clinical Safety and Efficacy. Life 2021, 11, 348. [Google Scholar] [CrossRef]
  28. Puri, A.; Saxena, R.; Saxena, R.P.; Saxena, K.C.; Srivastava, V.; Tandón, J.S. Immunostimulant Agents from Andrographis Paniculata1. J. Nat. Prod. 1993, 56, 995–999. [Google Scholar] [CrossRef]
  29. Ajaya Kumar, R.; Sridevi, K.; Vijaya Kumar, N.; Nanduri, S.; Rajagopal, S. Anticancer and Immunostimulatory Compounds from Andrographis Paniculata. J. Ethnopharmacol. 2004, 92, 291–295. [Google Scholar] [CrossRef]
  30. Ghazalee, N.S.; Jantan, I.; Arshad, L.; Haque, M.A. Immunosuppressive Effects of the Standardized Extract of Zingiber Zerumbet on Innate Immune Responses in Wistar Rats. Phytother. Res. 2019, 33, 929–938. [Google Scholar] [CrossRef]
  31. Girisa, S.; Shabnam, B.; Monisha, J.; Fan, L.; Halim, C.E.; Arfuso, F.; Ahn, K.S.; Sethi, G.; Kunnumakkara, A.B. Potential of Zerumbone as an Anti-Cancer Agent. Molecules 2019, 24, 734. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Arshad, L.; Jantan, I.; Bukhari, S.N.A.; Haque, M.A. Immunosuppressive Effects of Natural α,β-Unsaturated Carbonyl-Based Compounds, and Their Analogs and Derivatives, on Immune Cells: A Review. Front. Pharmacol. 2017, 8, 22. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Prasannan, R.; Kalesh, K.A.; Shanmugam, M.K.; Nachiyappan, A.; Ramachandran, L.; Nguyen, A.H.; Kumar, A.P.; Lakshmanan, M.; Ahn, K.S.; Sethi, G. Key Cell Signaling Pathways Modulated by Zerumbone: Role in the Prevention and Treatment of Cancer. Biochem. Pharmacol. 2012, 84, 1268–1276. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Haque, M.A.; Jantan, I.; Arshad, L.; Bukhari, S.N.A. Exploring the Immunomodulatory and Anticancer Properties of Zerumbone. Food Funct. 2017, 8, 3410–3431. [Google Scholar] [CrossRef] [PubMed]
  35. Murakami, A.; Shigemori, T.; Ohigashi, H. International Conference on Diet, Nutrition, and Cancer Zingiberaceous and Citrus Constituents, 1-Acetoxychavicol Acetate, Zerumbone, Auraptene, and Nobiletin, Suppress Lipopolysaccharide-Induced Cyclooxygenase-2 Expression in RAW264.7 Murine Macrophages through Different Modes of Action. J. Nutr. 2005, 135, 2987S–2992S. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Eguchi, A.; Kaneko, Y.; Murakami, A.; Ohigashi, H. Zerumbone Suppresses Phorbol Ester-Induced Expression of Multiple Scavenger Receptor Genes in THP-1 Human Monocytic Cells. Biosci. Biotechnol. Biochem. 2007, 71, 935–945. [Google Scholar] [CrossRef] [Green Version]
  37. Murakami, A.; Takahashi, D.; Kinoshita, T.; Koshimizu, K.; Kim, H.W.; Yoshihiro, A.; Nakamura, Y.; Jiwajinda, S.; Terao, J.; Ohigashi, H. Zerumbone, a Southeast Asian Ginger Sesquiterpene, Markedly Suppresses Free Radical Generation, Proinflammatory Protein Production, and Cancer Cell Proliferation Accompanied by Apoptosis: The α,β-Unsaturated Carbonyl Group Is a Prerequisite. Carcinogenesis 2002, 23, 795–802. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Abdel Wahab, S.I.; Abdul, A.B.; Alzubairi, A.S.; Mohamed Elhassan, M.; Mohan, S. In Vitro Ultramorphological Assessment of Apoptosis Induced by Zerumbone on (HeLa). J. Biomed. Biotechnol. 2009, 2009. [Google Scholar] [CrossRef]
  39. Yodkeeree, S.; Sung, B.; Limtrakul, P.; Aggarwal, B.B. Zerumbone Enhances TRAIL-Induced Apoptosis through the Induction of Death Receptors in Human Colon Cancer Cells: Evidence for an Essential Role of Reactive Oxygen Species. Cancer Res. 2009, 69, 6581–6589. [Google Scholar] [CrossRef] [Green Version]
  40. Kupchan, S.M.; Court, W.A.; Dailey, R.G., Jr.; Gilmore, C.J.; Bryan, R.F. Triptolide and Tripdiolide, Novel Antileukemic Diterpenoid Triepoxides from Tripterygium Wilfordii. J. Am. Chem. Soc. 1972, 94, 7194–7195. [Google Scholar] [CrossRef]
  41. Tao, X.; Cai, J.J.; Lipsky, P.E. The Identity of Immunosuppressive Components of the Ethyl Acetate Extract and Chloroform Methanol Extract. J. Pharmacol. Exp. Ther. 1995, 272, 1305–1312. [Google Scholar] [PubMed]
  42. Tao, X.; Davis, L.S.; Hashhoto, K.; Lipsky, R.E. The Chinese Herbal Remedy, T2. Inhibits Mitogen-Induced Cytokine Gene Transcription by T Cells, but Not Initial Signal Transduction. J. Pharmacol. Exp.Ther. 1996, 276, 316–325. [Google Scholar]
  43. Qiu, D.; Zhao, G.; Aoki, Y.; Shi, L.; Uyei, A.; Nazarian, S.; Ng, J.C.H.; Kao, P.N. Immunosuppressant PG490 (Triptolide) Inhibits T-Cell Interleukin-2 Expression at the Level of Purine-Box/Nuclear Factor of Activated T-Cells and NF-ΚB Transcriptional Activation. J. Biol. Chem. 1999, 274, 13443–13450. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Lee, K.Y.; Chang, W.T.; Qiu, D.; Kao, P.N.; Rosen, G.D. PG490 (Triptolide) Cooperates with Tumor Necrosis Factor-α to Induce Apoptosis in Tumor Cells. J. Biol. Chem. 1999, 274, 13451–13455. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Chang, W.T.; Kang, J.J.; Lee, K.Y.; Wei, K.; Anderson, E.; Gotmare, S.; Ross, J.A.; Rosen, G.D. Triptolide and Chemotherapy Cooperate in Tumor Cell Apoptosis. A Role for the P53 Pathway. J. Biol. Chem. 2001, 276, 2221–2227. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Rongzhen, X.; Hongjian, J. C14-Hydroxyl Esterified Amino Acid Derivatives of Triptolide, and Preparation Method and Use Thereof. CN Patent. EP3248981A1. Available online: https://patents.google.com/patent/EP3248981A1/en (accessed on 13 April 2022).
  47. Namin, M.H.; Ebrahimzadeh, H.; Ghareyazie, B.; Radjabian, T.; Gharavi, S.; Tafreshi, N.; Lee, H. In Vitro Expression of Apocarotenoid Genes in Crocus Sativus L. Afr. J. Biotechnol. 2009, 8, 5378–5382. [Google Scholar]
  48. Wyss, A. Carotene Oxygenases: A New Family of Double Bond Cleavage Enzymes. J. Nutr. 2004, 134, 246S–250S. [Google Scholar] [CrossRef] [Green Version]
  49. Bolhassani, A.; Khavari, A.; Bathaie, S.Z. Saffron and Natural Carotenoids: Biochemical Activities and Anti-Tumor Effects. Biochim. Et Biophys. Acta-Rev. Cancer 2014, 1845, 20–30. [Google Scholar] [CrossRef]
  50. Abar, L.; Vieira, A.R.; Aune, D.; Stevens, C.; Vingeliene, S.; Navarro Rosenblatt, D.A.; Chan, D.; Greenwood, D.C.; Norat, T. Blood Concentrations of Carotenoids and Retinol and Lung Cancer Risk: An Update of the WCRF–AICR Systematic Review of Published Prospective Studies. Cancer Med. 2016, 5, 2069–2083. [Google Scholar] [CrossRef] [Green Version]
  51. Rakic, J.M.; Liu, C.; Veeramachaneni, S.; Wu, D.; Paul, L.; Chen, C.Y.O.; Ausman, L.M.; Wang, X.D. Lycopene Inhibits Smoke-Induced Chronic Obstructive Pulmonary Disease and Lung Carcinogenesis by Modulating Reverse Cholesterol Transport in Ferrets. Cancer Prev. Res. 2019, 12, 421–432. [Google Scholar] [CrossRef]
  52. Eid, S.Y.; El-Readi, M.Z.; Wink, M. Carotenoids Reverse Multidrug Resistance in Cancer Cells by Interfering with ABC-Transporters. Phytomedicine 2012, 19, 977–987. [Google Scholar] [CrossRef] [PubMed]
  53. Awad, A.B.; Chen, Y.C.; Fink, C.S.; Hennessey, T. Beta-Sitosterol Inhibits HT-29 Human Colon Cancer Cell Growth and Alters Membrane Lipids. Anticancer Res. 1996, 16, 2797–2804. [Google Scholar] [PubMed]
  54. Bouic, P.J.D. The Role of Phytosterols and Phytosterolins in Immune Modulation: A Review of the Past 10 Years. Curr. Opin. Clin. Nutr. Metab. Care 2001, 4, 471–475. [Google Scholar] [CrossRef] [PubMed]
  55. Moreno, J.J. Effect of Olive Oil Minor Components on Oxidative Stress and Arachidonic Acid Mobilization and Metabolism by Macrophages RAW 264.7. Free. Radic. Biol. Med. 2003, 35, 1073–1081. [Google Scholar] [CrossRef]
  56. Awad, A.B.; Toczek, J.; Fink, C.S. Phytosterols Decrease Prostaglandin Release in Cultured P388D 1/MAB Macrophages. Prostaglandins Leukot. Essent. Fat. Acids 2004, 70, 511–520. [Google Scholar] [CrossRef]
  57. Ikemoto, S.; Sugimura, K.; Yoshida, N.; Yasumoto, R.; Wada, S.; Yamamoto, K.; Kishimoto, T. Antitumor Effects of Scutellariae Radix and Its Components Baicalein, Baicalin, and Wogonin on Bladder Cancer Cell Lines. Urology 2000, 55, 951–955. [Google Scholar] [CrossRef]
  58. Kaplya, O.A.; Sherstoboev, E.Y.; Zueva, E.P.; Razina, T.G.; Amosova, E.N.; Krylova, S.G. Effect of Baikal Skullcap Extract Administered Alone or in Combination with Cyclophosphamide on Natural Cytotoxicity System in Mice with Lewis Lung Carcinoma. Bull. Exp. Biol. Med. 2004, 137, 471–474. [Google Scholar] [CrossRef]
  59. Li-Weber, M. New Therapeutic Aspects of Flavones: The Anticancer Properties of Scutellaria and Its Main Active Constituents Wogonin, Baicalein and Baicalin. Cancer Treat. Rev. 2009, 35, 57–68. [Google Scholar] [CrossRef]
  60. Lamer-Zarawska, E.; Wiśniewska, A.; Błach-Olszewska, Z. Anticancer Properties of Scutellaria Baicalensis Root in Aspect of Innate Immunity Regulation Przeciwnowotworowa Aktywność Tarczycy Bajkalskiej w Świetle Regulacji Wrodzonej Odporności. Adv. Clin. Exp. Med. 2010, 19, 419–428. [Google Scholar]
  61. Ma, Z.; Otsuyama, K.-I.; Liu, S.; Abroun, S.; Ishikawa, H.; Tsuyama, N.; Obata, M.; Li, F.-J.; Zheng, X.; Maki, Y.; et al. Baicalein, a Component of Scutellaria Radix from Huang-Lian-Jie-Du-Tang (HLJDT), Leads to Suppression of Proliferation and Induction of Apoptosis in Human Myeloma Cells. Blood 2005, 105, 3312–3318. [Google Scholar] [CrossRef] [Green Version]
  62. Bonham, M.; Posakony, J.; Coleman, I.; Montgomery, B.; Simon, J.; Nelson, P.S. Characterization of Chemical Constituents in Scutellaria Baicalensis with Antiandrogenic and Growth-Inhibitory Activities toward Prostate Carcinoma. Clin. Cancer Res. 2005, 11, 3905–3914. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Kumagai, T.; Müller, C.I.; Desmond, J.C.; Imai, Y.; Heber, D.; Koeffler, H.P. Scutellaria Baicalensis, a Herbal Medicine: Anti-Proliferative and Apoptotic Activity against Acute Lymphocytic Leukemia, Lymphoma and Myeloma Cell Lines. Leuk. Res. 2007, 31, 523–530. [Google Scholar] [CrossRef] [PubMed]
  64. Chi, Y.S.; Lim, H.; Park, H.; Kim, H.P. Effects of Wogonin, a Plant Flavone from Scutellaria Radix, on Skin Inflammation: In Vivo Regulation of Inflammation-Associated Gene Expression. Biochem. Pharmacol. 2003, 66, 1271–1278. [Google Scholar] [CrossRef]
  65. Huang, W.H.; Lee, A.R.; Yang, C.H. Antioxidative and Anti-Inflammatory Activities of Polyhydroxyflavonoids of Scutellaria Baicalensis GEORGI. Biosci. Biotechnol. Biochem. 2006, 70, 2371–2380. [Google Scholar] [CrossRef] [Green Version]
  66. Lim, H.; Park, H.; Kim, H.P. Inhibition of Contact Dermatitis in Animal Models and Suppression of Proinflammatory Gene Expression by Topically Applied Flavonoid, Wogonin. Arch. Pharm. Res. 2004, 27, 442. [Google Scholar] [CrossRef]
  67. Błach-Olszewska, Z.; Jatczak, B.; Rak, A.; Lorenc, M.; Gulanowski, B.; Drobna, A.; Lamer-Zarawska, E. Production of Cytokines and Stimulation of Resistance to Viral Infection in Human Leukocytes by Scutellaria Baicalensis Flavones. J. Interferon Cytokine Res. 2008, 28, 571–581. [Google Scholar] [CrossRef]
  68. Chen, Y.C.; Shen, S.C.; Chen, L.G.; Lee, T.J.; Yang, L.L. Wogonin, Baicalin, and Baicalein Inhibition of Inducible Nitric Oxide Synthase and Cyclooxygenase-2 Gene Expressions Induced by Nitric Oxide Synthase Inhibitors and Lipopolysaccharide. Biochem. Pharmacol. 2001, 61, 1417–1427. [Google Scholar] [CrossRef]
  69. Chen, L.G.; Hung, L.Y.; Tsai, K.W.; Pan, Y.S.; da Tsai, Y.; Li, Y.Z.; Liu, Y.W. Wogonin, a Bioactive Flavonoid in Herbal Tea, Inhibits Inflammatory Cyclooxygenase-2 Gene Expression in Human Lung Epithelial Cancer Cells. Mol. Nutr. Food Res. 2008, 52, 1349–1357. [Google Scholar] [CrossRef]
  70. Kim, H.; Kim, Y.S.; Kim, S.Y.; Suk, K. The Plant flavonoid Wogonin Suppresses Death of Activated C6 Rat Glial Cells by Inhibiting Nitric Oxide Production. Neurosci. Lett. 2001, 309, 67–71. [Google Scholar] [CrossRef]
  71. Fukumura, D.; Kashiwagi, S.; Jain, R.K. The Role of Nitric Oxide in Tumour Progression. Nat. Rev. Cancer 2006, 6, 521–534. [Google Scholar] [CrossRef]
  72. Ke, M.; Zhang, Z.; Xu, B.; Zhao, S.; Ding, Y.; Wu, X.; Wu, R.; Lv, Y.; Dong, J. Baicalein and Baicalin Promote Antitumor Immunity by Suppressing PD-L1 Expression in Hepatocellular Carcinoma Cells. Int. Immunopharmacol. 2019, 75, 105824. [Google Scholar] [CrossRef]
  73. Anila, L.; Vijayalakshmi, N.R. Antioxidant Action of Flavonoids from Mangifera Indica and Emblica Officinalis in Hypercholesterolemic Rats. Food Chem. 2003, 83, 569–574. [Google Scholar] [CrossRef]
  74. Abesundara, K.J.M.; Matsui, T.; Matsumoto, K. α-Glucosidase Inhibitory Activity of Some Sri Lanka Plant Extracts, One of Which, Cassia Auriculata, Exerts a Strong Antihyperglycemic Effect in Rats Comparable to the Therapeutic Drug Acarbose. J. Agric. Food Chem. 2004, 52, 2541–2545. [Google Scholar] [CrossRef] [PubMed]
  75. Rani, P.; Khullar, N. Antimicrobial Evaluation of Some Medicinal Plants for Their Anti-Enteric Potential against Multi-Drug Resistant Salmonella Typhi. Phytother. Res. 2004, 18, 670–673. [Google Scholar] [CrossRef] [PubMed]
  76. Ram, M.S.; Neetu, D.; Yogesh, B.; Anju, B.; Dipti, P.; Pauline, T.; Sharma, S.K.; Sarada, S.K.S.; Ilavazhagan, G.; Kumar, D.; et al. Cyto-Protective and Immunomodulating Properties of Amla (Emblica Officinalis) on Lymphocytes: An in-Vitro Study. J. Ethnopharmacol. 2002, 81, 5–10. [Google Scholar] [CrossRef]
  77. Jose, J.K.; Kuttan, G.; Kuttan, R. Antitumour Activity of Emblica Officinalis. J. Ethnopharmacol. 2001, 75, 65–69. [Google Scholar] [CrossRef]
  78. Rajeshkumar, N.V.; Joy, K.L.; Kuttan, G.; Ramsewak, R.S.; Nair, M.G.; Kuttan, R. Antitumour and Anticarcinogenic Activity of Phyllanthus Amarus Extract. J. Ethnopharmacol. 2002, 81, 17–22. [Google Scholar] [CrossRef]
  79. Liu, X.; Zhao, M.; Wu, K.; Chai, X.; Yu, H.; Tao, Z.; Wang, J. Immunomodulatory and Anticancer Activities of Phenolics from Emblica Fruit (Phyllanthus Emblica L.). Food Chem. 2012, 131, 685–690. [Google Scholar] [CrossRef]
  80. Kawakami, K.; Nishida, H.; Tatewaki, N.; Eguchi-Kasai, K.; Anzai, K.; Eitsuka, T.; Konishi, T.; Hirayama, M. Persimmon Leaf Flavonols Enhance the Anti-Cancer Effect of Heavy Ion Radiotherapy on Murine Xenograft Tumors. J. Cancer Ther. 2013, 4, 1150–1157. [Google Scholar] [CrossRef] [Green Version]
  81. Ling, W.D.; Du, G. Effect of Flavonoid from Diospyros Kaki Leaves on TGF- β 1 and MMP-9 in Blood Glucose Levels and Kidney Tissues of Rats with Diabetic Nephropathy. Chin. J. Exp. Tradit. Med. Form 2016, 22, 139–143. [Google Scholar]
  82. Sun, L.; Zhang, J.; Lu, X.; Zhang, L.; Zhang, Y. Evaluation to the Antioxidant Activity of Total Flavonoids Extract from Persimmon (Diospyros Kaki L.) Leaves. Food Chem. Toxicol. 2011, 49, 2689–2696. [Google Scholar] [CrossRef] [PubMed]
  83. Chen, L.; Wei, Y.; Zhao, S.; Zhang, M.; Yan, X.; Gao, X.; Li, J.; Gao, Y.; Zhang, A.; Gao, Y. Antitumor and Immunomodulatory Activities of Total Flavonoids Extract from Persimmon Leaves in H22 Liver Tumor-Bearing Mice. Sci. Rep. 2018, 8, 1–11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Ganeshpurkar, A.; Saluja, A.K. The Pharmacological Potential of Rutin. Saudi Pharm. J. 2017, 25, 149–164. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Xu, R.; Zhang, Y.; Ye, X.; Xue, S.; Shi, J.; Pan, J.; Chen, Q. Inhibition Effects and Induction of Apoptosis of Flavonoids on the Prostate Cancer Cell Line PC-3 in Vitro. Food Chem. 2013, 138, 48–53. [Google Scholar] [CrossRef] [PubMed]
  86. Hayat, K.; Iqbal, H.; Malik, U.; Bilal, U.; Mushtaq, S. Tea and Its Consumption: Benefits and Risks. Crit. Rev. Food Sci. Nutr. 2015, 55, 939–954. [Google Scholar] [CrossRef]
  87. Crespy, V.; Williamson, G. A Review of the Health Effects of Green Tea Catechins in In Vivo Animal Models. J. Nutr. 2004, 134, 3431S–3440S. [Google Scholar] [CrossRef]
  88. Katiyar, S.; Elmets, C.A.; Katiyar, S.K. Green Tea and Skin Cancer: Photoimmunology, Angiogenesis and DNA Repair. J. Nutr. Biochem. 2007, 18, 287–296. [Google Scholar] [CrossRef]
  89. Shirakami, Y.; Shimizu, M.; Moriwaki, H. Cancer Chemoprevention with Green Tea Catechins: From Bench to Bed. Current Drug Targets 2012, 13, 1842–1857. [Google Scholar] [CrossRef]
  90. Butt, M.S.; Sultan, M.T. Green Tea: Nature’s Defense against Malignancies. Crit. Rev. Food Sci. Nutr. 2009, 49, 463–473. [Google Scholar] [CrossRef]
  91. Fujiki, H.; Watanabe, T.; Sueoka, E.; Rawangkan, A.; Suganuma, M. Cancer Prevention with Green Tea and Its Principal Constituent, EGCG: From Early Investigations to Current Focus on Human Cancer Stem Cells. Mol. Cells 2018, 41, 73–82. [Google Scholar]
  92. Negri, A.; Naponelli, V.; Rizzi, F.; Bettuzzi, S. Molecular Targets of Epigallocatechin—Gallate (EGCG): A Special Focus on Signal Transduction and Cancer. Nutrients 2018, 10, 1936. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Santilli, G.; Piotrowska, I.; Cantilena, S.; Chayka, O.; D’Alicarnasso, M.; Morgenstern, D.A.; Himoudi, N.; Pearson, K.; Anderson, J.; Thrasher, A.J.; et al. Polyphenol e Enhances the Antitumor Immune Response in Neuroblastoma by Inactivating Myeloid Suppressor Cells. Clin. Cancer Res. 2013, 19, 1116–1125. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Tae, H.K.; Jin, H.L.; Chung, K.S.; Hee, D.H.; Byung, C.S.; Pai, S.I.; Hung, C.F.; Trimble, C.; Lim, J.S.; Tae, W.K.; et al. Epigallocatechin-3-Gallate Enhances CD8+ T Cell-Mediated Antitumor Immunity Induced by DNA Vaccination. Cancer Res. 2007, 67, 802–811. [Google Scholar] [CrossRef] [Green Version]
  95. Mantena, S.K.; Roy, A.M.; Katiyar, S.K. Epigallocatechin-3-Gallate Inhibits Photocarcinogenesis Through Inhibition of Angiogenic Factors and Activation of CD8+ T Cells in Tumors. Photochem. Photobiol. 2005, 81, 1174. [Google Scholar] [CrossRef]
  96. Hsieh, D.S.; Wang, H.; Tan, S.W.; Huang, Y.H.; Tsai, C.Y.; Yeh, M.K.; Wu, C.J. The Treatment of Bladder Cancer in a Mouse Model by Epigallocatechin-3-Gallate-Gold Nanoparticles. Biomaterials 2011, 32, 7633–7640. [Google Scholar] [CrossRef]
  97. Rawangkan, A.; Wongsirisin, P.; Namiki, K.; Iida, K.; Kobayashi, Y.; Shimizu, Y.; Fujiki, H.; Suganuma, M. Green Tea Catechin Is an Alternative Immune Checkpoint Inhibitor That Inhibits PD-L1 Expression and Lung Tumor Growth. Molecules 2018, 23, 2071. [Google Scholar] [CrossRef] [Green Version]
  98. Belguendouz, L.; Frémont, L.; Gozzelino, M.-T. Interaction of Transresveratrol with Plasma Lipoproteins. Biochem. Pharmacol. 1998, 55, 811–816. [Google Scholar] [CrossRef]
  99. Kisková, T.; Kassayová, M. Resveratrol Action on Lipid Metabolism in Cancer. Int. J. Mol. Sci. 2019, 20, 2704. [Google Scholar] [CrossRef] [Green Version]
  100. Pace-Asciak, C.R.; Hahn, S.; Diamandis, E.P.; Soleas, G.; Goldberg, D.M. The Red Wine Phenolics Trans-Resveratrol and Quercetin Block Human Platelet Aggregation and Eicosanoid Synthesis: Implications for Protection against Coronary Heart Disease. Clin. Chim. Acta 1995, 235, 207–219. [Google Scholar] [CrossRef]
  101. Rotondo, S.; Rajtar, G.; Manarini, S.; Celardo, A.; Rotilio, D.; de Gaetano, G.; Evangelista, V.; Cerletti, C. Effect of Trans-Resveratrol, a Natural Polyphenolic Compound, on Human Polymorphonuclear Leukocyte Function. Br. J. Pharmacol. 1998, 123, 1691–1699. [Google Scholar] [CrossRef] [Green Version]
  102. Kimura, Y.; Okuda, H.; Kubo, M. Effects of Stilbenes Isolated from Medicinal Plants on Arachidonate Metabolism and Degranulation in Human Polymorphonuclear Leukocytes. J. Ethnopharmacol. 1995, 45, 131–139. [Google Scholar] [CrossRef]
  103. Belguendouz, L.; Fremont, L.; Hard, A. Resveratrol Inhibits Metal Ion-Dependent and Independent Peroxidation of Porcine Low-Density Lipoproteins. Biochem Pharmacol. 1997, 53, 1347–1355. [Google Scholar] [CrossRef]
  104. Man-Ying Chan, M.; Mattiacci, J.A.; Hwang, H.S.; Shah, A.; Fong, D. Synergy between Ethanol and Grape Polyphenols, Quercetin, and Resveratrol, in the Inhibition of the Inducible Nitric Oxide Synthase Pathway. Biochem. Pharmacol. 2000, 60, 1539–1548. [Google Scholar] [CrossRef]
  105. Jang, M.; Cai, L.; Udeani, G.O.; Slowing, K.V.; Thomas, C.F.; Beecher, C.W.W.; Fong, H.H.S.; Farnsworth, N.R.; Kinghorn, A.D.; Mehta, R.G.; et al. Cancer Chemopreventive Activity of Resveratrol, a Natural Product Derived from Grapes. Science 1997, 275, 218–220. [Google Scholar] [CrossRef] [Green Version]
  106. Mgbonyebi, O.P.; Russo, J.; Russo, I.H. Antiproliferative Effect of Synthetic Resveratrol on Human Breast Epithelial Cells. Int. J. Oncol. 1998, 12, 865–874. [Google Scholar] [CrossRef]
  107. Lu, R.; Serrero, G. Resveratrol, a Natural Product Derived from Grape, Exhibits Antiestrogenic Activity and Inhibits the Growth of Human Breast Cancer Cells. J. Cell Physiol. 1999, 179, 297–304. [Google Scholar] [CrossRef]
  108. Hsieh, T.-C.; Wu, J.M. Differential Effects on Growth, Cell Cycle Arrest, and Induction of Apoptosis by Resveratrol in Human Prostate Cancer Cell Lines. Exp Cell Res. 1999, 249, 109–115. [Google Scholar] [CrossRef]
  109. Schneider, Y.; Vincent, F.; Duranton, B.Ã.; Badolo, L.; Gosse, F.; Bergmann, C.; Seiler, N.; Raul, F. Anti-Proliferative Effect of Resveratrol, a Natural Component of Grapes and Wine, on Human Colonic Cancer Cells. Cancer Lett. 2000, 158, 85–91. [Google Scholar] [CrossRef]
  110. Gao, X.; Xu, Y.X.; Janakiraman, N.; Chapman, R.A.; Gautam, S.C. Immunomodulatory Activity of Resveratrol: Suppression of Lymphocyte Proliferation, Development of Cell-Mediated Cytotoxicity, and Cytokine Production. Biochem. Pharmacol. 2001, 62, 1299–1308. [Google Scholar] [CrossRef]
  111. Bergman, M.; Levin, G.S.; Bessler, H.; Djaldetti, M.; Salman, H. Resveratrol Affects the Cross Talk between Immune and Colon Cancer Cells. Biomed. Pharmacother. 2013, 67, 43–47. [Google Scholar] [CrossRef]
  112. Majdalawieh, A.F.; Carr, R.I. In Vitro Investigation of the Potential Immunomodulatory and Anti-Cancer Activities of Black Pepper (Piper Nigrum) and Cardamom (Elettaria Cardamomum). J. Med. Food 2010, 13, 371–381. [Google Scholar] [CrossRef] [PubMed]
  113. Surh, Y.-J. Anti-Tumor Promoting Potential of Selected Spice Ingredients with Antioxidative and Anti-Inflammatory Activities: A Short Review. Food Chem. Toxicol. 2002, 40, 1091–1097. [Google Scholar] [CrossRef]
  114. Kaefer, C.M.; Milner, J.A. The Role of Herbs and Spices in Cancer Prevention. J. Nutr. Biochem. 2008, 19, 347–361. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Pathak, N.; Khandelwal, S. Cytoprotective and Immunomodulating Properties of Piperine on Murine Splenocytes: An in Vitro Study. Eur. J. Pharmacol. 2007, 576, 160–170. [Google Scholar] [CrossRef] [PubMed]
  116. Duessel, S.; Heuertz, R.M.; Ezekiel, U.R. Growth Inhibition of Human Colon Cancer Cells by Plant Compounds. Am. Soc. Clin. Lab. Sci. 2008, 21, 151–157. [Google Scholar]
  117. Menon, L.G.; Kuttan, R.; Kuttan, G. Effect of Rasayanas in the Inhibition of Lung Metastasis Induced by B16F-10 Melanoma Cells. J. Exp. Clin. Cancer Res. 1997, 16, 365–368. [Google Scholar]
  118. Selvendiran, K.; Mumtaz Banu, S.; Sakthisekaran, D. Oral Supplementation of Piperine Leads to Altered Phase II Enzymes and Reduced DNA Damage and DNA-Protein Cross Links in Benzo(a)Pyrene Induced Experimental Lung Carcinogenesis. Mol Cell Biochem. 2005, 268, 141–147. [Google Scholar] [CrossRef]
  119. Selvendiran, K.; Thirunavukkarasu, C.; Prince, J.; Singh, V.; Padmavathi, R.; Sakthisekaran, D. Chemopreventive Effect of Piperine on Mitochondrial TCA Cycle and Phase-I and Glutathione-Metabolizing Enzymes in Benzo(a)Pyrene Induced Lung Carcinogenesis in Swiss Albino Mice. Mol. Cell Biochem. 2005, 271, 101–106. [Google Scholar] [CrossRef]
  120. Krishnakumar, N.; Manoharan, S.; Palaniappan, P.R.; Venkatachalam, P.; Manohar, M.G.A. Chemopreventive Efficacy of Piperine in 7,12-Dimethyl Benz [a] Anthracene (DMBA)-Induced Hamster Buccal Pouch Carcinogenesis: An FT-IR Study. Food Chem. Toxicol. 2009, 47, 2813–2820. [Google Scholar] [CrossRef]
  121. Manoharan, S.; Balakrishnan, S.; Menon, V.P.; Alias, L.M.; Reena, A.R. Chemopreventive Efficacy of Curcumin and Piperine during 7,12-Dimethylbenz [a]Anthracene-Induced Hamster Buccal Pouch Carcinogenesis. Singapore Med. J. 2009, 50, 139. [Google Scholar]
  122. Block, R.M.; Lewis, R.D.; Shea@, J.B.; Fawley, J.; Richmond, V. Cell-Mediated Immune Response to Dog Pulp Tissue Altered by Eugenol within the Root Canal. Oral Surg. Oral Med. Oral Pathol. 1978, 45, 452–463. [Google Scholar] [CrossRef]
  123. Van Duuren, B.L.; Sivak, A.; Segal, A.; Orris, L.; Langseth, L. The Tumor-Promoting Agents of Tobacco Leaf and Tobacco Smoke Condensate. J Natl Cancer Inst. 1966, 37, 519–526. [Google Scholar] [PubMed]
  124. Van Duuren, B.L.; Goldschmidt, B.M. Cocarcinogenic and Tumor-Promoting Agents in Tobacco Carcinogenesis. J. Natl. Cancer Inst. 1976, 56, 1237–1242. [Google Scholar] [CrossRef] [PubMed]
  125. Sukumaran, K.; Unnikrishnan, M.C.; Kuttan, R. Inhjbition of Tumour Promotion in Mice by Eugenol. Indian J. Physiol. Pharmacol. 1994, 38, 306–308. [Google Scholar]
  126. Ghosh, R.; Nadiminty, N.; Fitzpatrick, J.E.; Alworth, W.L.; Slaga, T.J.; Kumar, A.P. Eugenol Causes Melanoma Growth Suppression through Inhibition of E2F1 Transcriptional Activity. J. Biol. Chem. 2005, 280, 5812–5819. [Google Scholar] [CrossRef] [Green Version]
  127. Edwards, R.L.; Luis, P.B.; Varuzza, P.V.; Joseph, A.I.; Presley, S.H.; Chaturvedi, R.; Schneider, C. The Anti-Inflammatory Activity of Curcumin Is Mediated by Its Oxidative Metabolites. J. Biol. Chem. 2017, 292, 21243–21252. [Google Scholar] [CrossRef] [Green Version]
  128. Galet, C.; Gollapudi, K.; Stepanian, S.; Byrd, J.B.; Henning, S.M.; Grogan, T.; Elashoff, D.; Heber, D.; Said, J.; Cohen, P.; et al. Effect of a Low-Fat Fish Oil Diet on Proinflammatory Eicosanoids and Cell-Cycle Progression Score in Men Undergoing Radical Prostatectomy. Cancer Prev. Res. 2014, 7, 97–104. [Google Scholar] [CrossRef] [Green Version]
  129. Panahi, Y.; Darvishi, B.; Ghanei, M.; Jowzi, N.; Beiraghdar, F.; Varnamkhasti, B.S. Molecular Mechanisms of Curcumins Suppressing Effects on Tumorigenesis, Angiogenesis and Metastasis, Focusing on NF-ΚB Pathway. Cytokine Growth Factor Rev. 2016, 28, 21–29. [Google Scholar] [CrossRef]
  130. Wilken, R.; Veena, M.S.; Wang, M.B.; Srivatsan, E.S. Curcumin: A Review of Anti-Cancer Properties and Therapeutic Activity in Head and Neck Squamous Cell Carcinoma. Mol. Cancer 2011, 10, 1–19. [Google Scholar] [CrossRef] [Green Version]
  131. Chandra, D.; Jahangir, A.; Cornelis, F.; Rombauts, K.; Meheus, L.; Jorcyk, C.L.; Gravekamp, C. Cryoablation and Meriva Have Strong Therapeutic Effect on Triple-Negative Breast Cancer. OncoImmunology 2016, 5, e1049802. [Google Scholar] [CrossRef] [Green Version]
  132. Nahar, P.P.; Slitt, A.L.; Seeram, N.P. Anti-Inflammatory Effects of Novel Standardized Solid Lipid Curcumin Formulations. J. Med. Food 2015, 18, 786–792. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Jeong, Y.I.; Kim, S.W.; Jung, I.D.; Lee, J.S.; Chang, J.H.; Lee, C.M.; Chun, S.H.; Yoon, M.S.; Kim, G.T.; Ryu, S.W.; et al. Curcumin Suppresses the Induction of Indoleamine 2,3-Dioxygenase by Blocking the Janus-Activated Kinase-Protein Kinase Cδ-STAT1 Signaling Pathway in Interferon-γ-Stimulated Murine Dendritic Cells. J. Biol. Chem. 2009, 284, 3700–3708. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Lim, S.O.; Li, C.W.; Xia, W.; Cha, J.H.; Chan, L.C.; Wu, Y.; Chang, S.S.; Lin, W.C.; Hsu, J.M.; Hsu, Y.H.; et al. Deubiquitination and Stabilization of PD-L1 by CSN5. Cancer Cell 2016, 30, 925–939. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Ireson, C.R.; Jones, D.J.L.; Orr, S.; Coughtrie, M.W.H.; Boocock, D.J.; Williams, M.L.; Farmer, P.B.; Steward, W.P.; Gescher, A.J. Metabolism of the Cancer Chemopreventive Agent Curcumin in Human and Rat Intestine. Cancer Epidemiol. Prev. Biomark. 2002, 11, 105–111. [Google Scholar]
  136. Moradali, M.F.; Mostafavi, H.; Ghods, S.; Hedjaroude, G.A. Immunomodulating and Anticancer Agents in the Realm of Macromycetes Fungi (Macrofungi). Int. Immunopharmacol. 2007, 7, 701–724. [Google Scholar] [CrossRef]
  137. Wasser, S. Medicinal Mushrooms as a Source of Antitumor and Immunomodulating Polysaccharides. Appl. Microbiol. Biotechnol. 2002, 60, 258–274. [Google Scholar]
  138. Wachtel-Galor, S.; Yuen, J.; Buswell, J.A.; Benzie, I.F.F. Ganoderma lucidum (Lingzhi or Reishi): A Medicinal Mushroom. In Herbal Medicine: Biomolecular and Clinical Aspects, 2nd ed.; Benzie, I.F.F., Wachtel-Galor, S., Eds.; CRC Press/Taylor & Francis: Boca Raton, FL, USA, 2011. Available online: https://www.ncbi.nlm.nih.gov/books/NBK92757/ (accessed on 13 April 2022)Chapter 9.
  139. Gao, Y.; Zhou, S.H.; Chen, G.; Dai, X.; Ye, J. A Phase III Study of a Ganoderma Lucidum (Curt. Fr.) P. Karst. Extract (Ganopoly) in Patients with Advanced Cancer. Int. J. Med. Mushroom 2002, 4, 207–214. [Google Scholar]
  140. Burdock, G.A. Review of the Biological Properties and Toxicity of Bee Propolis (Propolis). Food Chem. Toxicol. 1998, 36, 347–363. [Google Scholar] [CrossRef]
  141. Castaldo, S.; Capasso, F. Propolis, an Old Remedy Used in Modern Medicine. Fitoterapia 2002, 73 (Suppl. 1), S1–S6. [Google Scholar] [CrossRef]
  142. Borrelli, F.; Maffia, P.; Pinto, L.; Ianaro, A.; Russo, A.; Capasso, F.; Ialenti, A. Phytochemical Compounds Involved in the Anti-Inflammatory Effect of Propolis Extract. Fitoterapia 2002, 73 (Suppl. 1), S53–S63. [Google Scholar] [CrossRef]
  143. Oršolić, N.; Bašić, I. Immunomodulation by Water-Soluble Derivative of Propolis: A Factor of Antitumor Reactivity. J. Ethnopharmacol. 2003, 84, 265–273. [Google Scholar] [CrossRef]
  144. Marcucci, M.C.; Ferreres, F.; Custódio, A.R.; Ferreira, M.M.C.; Bankovad, V.S.; García-Viguera, C.; Bretz, W.A. Evalution of Phenolic Compounds in Brazilian Propolis from Different Geographic Regions. Z Naturforsch C J. Biosci. 2000, 55, 76–81. [Google Scholar] [CrossRef] [PubMed]
  145. Cole, N.; Sou, P.W.; Ngo, A.; Tsang, K.H.; Severino, J.A.J.; Arun, S.J.; Duke, C.C.; Reeve, V.E. Topical “sydney” Propolis Protects against UV-Radiation-Induced Inflammation, Lipid Peroxidation and Immune Suppression in Mouse Skin. Int. Arch. Allergy Immunol. 2010, 152, 87–97. [Google Scholar] [CrossRef]
  146. Bankova, V. Recent Trends and Important Developments in Propolis Research. Evid. -Based Complementary Altern. Med. 2005, 2, 29–32. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Lin, W.L.; Liang, W.H.; Lee, Y.J.; Chuang, S.K.; Tseng, T.H. Antitumor Progression Potential of Caffeic Acid Phenethyl Ester Involving P75NTR in C6 Glioma Cells. Chem.-Biol. Interact. 2010, 188, 607–615. [Google Scholar] [CrossRef]
  148. Ang, E.S.M.; Pavlos, N.J.; Chai, L.Y.; Qi, M.; Cheng, T.S.; Steer, J.H.; Joyce, D.A.; Zheng, M.H.; Xu, J. Caffeic Acid Phenethyl Ester, an Active Component of Honeybee Propolis Attenuates Osteoclastogenesis and Bone Resorption via the Suppression of RANKL-Induced NF-ΚB and NFAT Activity. J. Cell. Physiol. 2009, 221, 642–649. [Google Scholar] [CrossRef]
  149. Park, J.H.; Lee, J.K.; Kim, H.S.; Chung, S.T.; Eom, J.H.; Kim, K.A.; Chung, S.J.; Paik, S.Y.; Oh, H.Y. Immunomodulatory Effect of Caffeic Acid Phenethyl Ester in Balb/c Mice. Int. Immunopharmacol. 2004, 4, 429–436. [Google Scholar] [CrossRef]
  150. Huang, M.-T.; Ma, W.; Yen, P.; Xie, J.-G.; Han, J.; Frenkel, K.; Grunberger, D.; Conney, H.-A. Inhibitory Effects of Caffeic Acid Phenethyl Ester (CAPE) on 12-0-Tetradecanoylphorbol-13-Acetate-Induced Tumor Promotion in Mouse Skin and the Synthesis of DNA, RNA and Protein in HeLa Cells. Carcinogenesis 1996, 17, 761–765. [Google Scholar] [CrossRef] [Green Version]
  151. Natarajan, K.; Singh, S.; Burke, T.R.; GRUNBERGERt, D.; Aggarwal, B.B. Caffeic Acid Phenethyl Ester Is a Potent and Specific Inhibitor of Activation of Nuclear Transcription Factor NF-KB (Tumor Necrosis Factor/Okadaic Acid/Ceramide/Phorbol Ester/Hydrogen Peroxide). Proc. Natl. Acad. Sci. USA 1996, 93, 9090–9095. [Google Scholar] [CrossRef] [Green Version]
  152. Roos, T.U.; Heiss, E.H.; Schwaiberger, A.V.; Schachner, D.; Sroka, I.M.; Oberan, T.; Vollmar, A.M.; Dirsch, V.M. Caffeic Acid Phenethyl Ester Inhibits PDGF-Induced Proliferation of Vascular Smooth Muscle Cells via Activation of P38 MAPK, HIF-1α, and Heme Oxygenase-1. J. Nat. Prod. 2011, 74, 352–356. [Google Scholar] [CrossRef]
  153. Basini, G.; Baioni, L.; Bussolati, S.; Grasselli, F.; Daquino, C.; Spatafora, C.; Tringali, C. Antiangiogenic Properties of an Unusual Benzo[k,l]Xanthene Lignan Derived from CAPE (Caffeic Acid Phenethyl Ester). Investig. New Drugs 2012, 30, 186–190. [Google Scholar] [CrossRef] [PubMed]
  154. Márquez, N.; Sancho, R.; Macho, A.; Calzado, M.A.; Fiebich, B.L.; Muñoz, E. Caffeic Acid Phenethyl Ester Inhibits T-Cell Activation by Targeting Both Nuclear Factor of Activated T-Cells and NF-ΚB Transcription Factors. J. Pharmacol. Exp. Ther. 2004, 308, 993–1001. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Wang, L.C.; Lin, Y.L.; Liang, Y.C.; Yang, Y.H.; Lee, J.H.; Yu, H.H.; Wu, W.M.; Chiang, B.L. The Effect of Caffeic Acid Phenethyl Ester on the Functions of Human Monocyte-Derived Dendritic Cells. BMC Immunol. 2009, 10, 1–13. [Google Scholar] [CrossRef] [Green Version]
  156. Wang, L.C.; Chu, K.H.; Liang, Y.C.; Lin, Y.L.; Chiang, B.L. Caffeic Acid Phenethyl Ester Inhibits Nuclear Factor-ΚB and Protein Kinase B Signalling Pathways and Induces Caspase-3 Expression in Primary Human CD4+ T Cells. Clin. Exp. Immunol. 2010, 160, 223–232. [Google Scholar] [CrossRef] [PubMed]
  157. Su, Z.-Z.; Lin, J.; Grunberger, D.; Fisher2, P.B. Growth Suppression and Toxicity Induced by Caffeic Acid Phenethyl Ester (CAPE) in Type 5 Adenovirus-Transformed Rat Embryo Cells Correlate Directly with Transformation Progression. Cancer Res 1994, 54, 1865–1870. [Google Scholar] [PubMed]
  158. Chena, J.-H.; Shaoa, Y.; Huang, M.; Chinb, C.-K.; Hoa, C.-T. Inhibitory Effect of Caffeic Acid Phenethyl Ester on Human Leukemia HL-60 Cells. Cancer Lett. 1996, 108, 211–214. [Google Scholar] [CrossRef]
  159. Lee, Y.-J.; Liao, P.-H.; Chen, W.-K.; Yang, C.-C. Preferential Cytotoxicity of Caffeic Acid Phenethyl Ester Analogues on Oral Cancer Cells. Cancer Lett. 2000, 153, 51–56. [Google Scholar] [CrossRef] [PubMed]
  160. Paulino, N.; Abreu, S.R.L.; Uto, Y.; Koyama, D.; Nagasawa, H.; Hori, H.; Dirsch, V.M.; Vollmar, A.M.; Scremin, A.; Bretz, W.A. Anti-Inflammatory Effects of a Bioavailable Compound, Artepillin C, in Brazilian Propolis. Eur. J. Pharmacol. 2008, 587, 296–301. [Google Scholar] [CrossRef]
  161. von Holtz, R.L.; Fink, C.S.; Awad, A.B. β-Sitosterol Activates the Sphingomyelin Cycle and Induces Apoptosis in LNCaP Human Prostate Cancer Cells. Nutr. Cancer 1998, 32, 8–12. [Google Scholar] [CrossRef]
  162. Galasso, C.; Nuzzo, G.; Brunet, C.; Ianora, A.; Sardo, A.; Fontana, A.; Sansone, C. The Marine Dinoflagellate Alexandrium Minutum Activates a Mitophagic Pathway in Human Lung Cancer Cells. Mar. Drugs 2018, 16, 502. [Google Scholar] [CrossRef] [Green Version]
  163. D’Ippolito, G.; Cutignano, A.; Tucci, S.; Romano, G.; Cimino, G.; Fontana, A. Biosynthetic intermediates and stereochemical aspects of aldehyde biosynthesis in the marine diatom Thalassiosira rotula. Phytochemistry 2006, 67, 314–322. [Google Scholar] [CrossRef]
  164. Kim, S.K.; Thomas, N.V.; Li, X. Anticancer Compounds from Marine Macroalgae and Their Application as Medicinal Foods. In Advances in Food and Nutrition Research; Academic Press Inc.: Cambridge, MA, USA, 2011; Volume 64, pp. 213–224. [Google Scholar]
  165. Maruyama, H.; Tamauchi, H.; Iizuka, M.; Nakano, T.; Maruyamá, H. The Role of NK Cells in Antitumor Activity of Dietary Fucoidan from Undaria Pinnatifida Sporophylls (Mekabu). Planta Med. 2006, 72, 1415–1417. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Zhou, G.; Sun, Y.P.; Xin, H.; Zhang, Y.; Li, Z.; Xu, Z. In Vivo Antitumor and Immunomodulation Activities of Different Molecular Weight Lambda-Carrageenans from Chondrus Ocellatus. Pharmacol. Res. 2004, 50, 47–53. [Google Scholar] [CrossRef] [PubMed]
  167. Sun, L.; Wang, L.; Zhou, Y. Immunomodulation and Antitumor Activities of Different-Molecular-Weight Polysaccharides from Porphyridium Cruentum. Carbohydr. Polym. 2012, 87, 1206–1210. [Google Scholar] [CrossRef]
  168. Shi, W.; Lu, D.; Wu, C.; Li, M.; Ding, Z.; Li, Y.; Chen, B.; Lin, X.; Su, W.; Shao, X.; et al. Coibamide A Kills Cancer Cells through Inhibiting Autophagy. Biochem. Biophys. Res. Commun. 2021, 547, 52–58. [Google Scholar] [CrossRef]
  169. Kong, C.S.; Kim, J.A.; Yoon, N.Y.; Kim, S.K. Induction of Apoptosis by Phloroglucinol Derivative from Ecklonia Cava in MCF-7 Human Breast Cancer Cells. Food Chem. Toxicol. 2009, 47, 1653–1658. [Google Scholar] [CrossRef]
  170. Kurihara, H.; Koda, H.; Asami, S.; Kiso, Y.; Tanaka, T. Contribution of the Antioxidative Property of Astaxanthin to Its Protective Effect on the Promotion of Cancer Metastasis in Mice Treated with Restraint Stress. Life Sci. 2002, 70, 2509–2520. [Google Scholar] [CrossRef]
  171. Tanaka, T.; Morishita, Y.; Suzui, M.; Kojima, T.; Okumura, A.; Mori, H. Chemoprevention of Mouse Urinary Bladder Carcinogenesis by the Naturally Occurring Carotenoid Astaxanthin. Carcinogenesis 1994, 15, 15–19. [Google Scholar] [CrossRef] [PubMed]
  172. Tanaka, T.; Shnimizu, M.; Moriwaki, H. Cancer Chemoprevention by Carotenoids. Molecules 2012, 17, 3202–3242. [Google Scholar] [CrossRef]
  173. Tanaka, T.; Makita, H.; Ohnishi, M.; Mori, H.; Satoh, K.; Hara, A. Chemoprevention of rat oral carcinogenesis by naturally occurring xanthophylls, astaxanthin and canthaxanthin. Cancer Res. 1995, 55, 4059–4064. [Google Scholar]
  174. Chang, Y.-J.; Huang, J.-R.; Tsai, Y.-C.; Hung, J.-T.; Wu, D.; Fujio, M.; Wong, C.-H.; Yu, A.L. Potent Immune-Modulating and Anticancer Effects of NKT Cell Stimulatory Glycolipids. Proc. Natl. Acad. Sci. USA 2007, 104, 10299–10304. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Bae, E.A.; Seo, H.; Kim, I.K.; Jeon, I.; Kang, C.Y. Roles of NKT Cells in Cancer Immunotherapy. Arch. Pharmacol. Res. 2019, 42, 543–548. [Google Scholar] [CrossRef] [PubMed]
  176. Manzo, E.; Cutignano, A.; Pagano, D.; Gallo, C.; Barra, G.; Nuzzo, G.; Sansone, C.; Ianora, A.; Urbanek, K.; Fenoglio, D.; et al. A New Marine-Derived Sulfoglycolipid Triggers Dendritic Cell Activation and Immune Adjuvant Response. Sci. Rep. 2017, 7, 1–10. [Google Scholar] [CrossRef] [PubMed]
  177. Gallo, C.; Ferrera, F.; Castiglia, D.; Fontana, A. Identication of Sulfavant A as the First Synthetic TREM2 Ligand Discloses a Homeostatic Response of Dendritic Cells After Receptor Engagement. Res. Sq. 2022. [Google Scholar] [CrossRef]
  178. Rinehart, K.L.; Gloer, J.B.; Cook, J.C.; Mizsak, S.A.; Scahill, T.A. Structures of the Didemnins, Antiviral and Cytotoxic Depsipeptides from a Caribbean Tunicate. J. Am. Chem. Soc. 1981, 103, 1857–1859. [Google Scholar] [CrossRef]
  179. Thell, K.; Hellinger, R.; Schabbauer, G.; Gruber, C.W. Immunosuppressive Peptides and Their Therapeutic Applications. Drug Discov. Today 2014, 19, 645–653. [Google Scholar] [CrossRef] [Green Version]
  180. Lee, J.; Currano, J.N.; Carroll, P.J.; Joullié, M.M. Didemnins, Tamandarins and Related Natural Products. Nat. Prod. Rep. 2012, 29, 404–424. [Google Scholar] [CrossRef]
  181. Tsukimoto, M.; Nagaoka, M.; Shishido, Y.; Fujimoto, J.; Nishisaka, F.; Matsumoto, S.; Harunari, E.; Imada, C.; Matsuzaki, T. Bacterial Production of the Tunicate-Derived Antitumor Cyclic Depsipeptide Didemnin B. J. Nat. Prod. 2011, 74, 11–2329. [Google Scholar] [CrossRef]
  182. Xu, Y.; Kersten, R.D.; Nam, S.J.; Lu, L.; Al-Suwailem, A.M.; Zheng, H.; Fenical, W.; Dorrestein, P.C.; Moore, B.S.; Qian, P.Y. Bacterial biosynthesis and maturation of the didemnin anti-cancer agents. J. Am. Chem. Soc. 2012, 134, 8625–8632. [Google Scholar] [CrossRef] [Green Version]
  183. Zhang, Q.T.; Liu, Z.D.; Wang, Z.; Wang, T.; Wang, N.; Wang, N.; Zhang, B.; Zhao, Y.F. Recent Advances in Small Peptides of Marine Origin in Cancer Therapy. Mar. Drugs 2021, 19, 115. [Google Scholar] [CrossRef]
  184. Delgado-Calle, J.; Kurihara, N.; Atkinson, E.G.; Nelson, J.; Miyagawa, K.; Galmarini, C.M.; Roodman, G.D.; Bellido, T. Aplidin (plitidepsin) is a novel anti-myeloma agent with potent anti-resorptive activity mediated by direct effects on osteoclasts. Oncotarget 2019, 10, 2709–2721. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. White, K.M.; Rosales, R.; Yildiz, S.; Kehrer, T.; Miorin, L.; Moreno, E.; Jangra, S.; Uccellini, M.B.; Rathnasinghe, R.; Coughlan, L.; et al. Plitidepsin has potent preclinical efficacy against SARS-CoV-2 by targeting the host protein eEF1A. Science 2021, 371, 926–931. [Google Scholar] [CrossRef] [PubMed]
  186. Cooper, E.L.; Yao, D. Diving for Drugs: Tunicate Anticancer Compounds. Drug Discov. Today 2012, 17, 636–648. [Google Scholar] [CrossRef] [PubMed]
  187. Oda, T.; Kamoshita, K.; Maruyama, S.; Masuda, K.; Nishimoto, M.; Xu, J.; Ukai, K.; Mangindaan, R.E.P.; Namikoshi, M. Cytotoxicity of Lissoclibadins and Lissoclinotoxins, Isolated from a Tropical Ascidian Lissoclinum Cf. Badium, against Human Solid-Tumor-Derived Cell Lines. Biol. Pharm. Bull. 2007, 30, 385–387. [Google Scholar] [CrossRef] [Green Version]
  188. Fedorov, S.N.; Radchenko, O.S.; Shubina, L.K.; Balaneva, N.N.; Bode, A.M.; Stonik, V.A.; Dong, Z. Evaluation of cancer-preventive activity and structure-activity relationships of 3-demethylubiquinone Q2, isolated from the ascidian Aplidium glabrum, and its synthetic analogs. Pharm. Res. 2006, 23, 70–81. [Google Scholar] [CrossRef] [Green Version]
  189. Nuzzo, G.; Gallo, C.; Crocetta, F.; Romano, L.; Barra, G.; Senese, G.; dell’Isola, M.; Carbone, D.; Tanduo, V.; Albiani, F.; et al. Identification of the Marine Alkaloid Lepadin A as Potential Inducer of Immunogenic Cell Death. Biomolecules 2022, 12, 246. [Google Scholar] [CrossRef]
  190. Journal, E.; Meenakshi, V.K.; Senthamarai, S.; Paripooranaselvi, M.; Gomathy, S.; Sankaravadivu, S.; Chamundeswari, K.P. Scholars Research Library In Vitro and in Vivo Antitumor and Immunomodulatory Studies of Microcosmus Exasperatus against DLA Bearing Mice. Eur. J. Appl. Eng. Sci. Res. 2013, 23, 18–25. Available online: http://scholarsresearchlibrary.com/archive.html (accessed on 13 April 2022).
  191. Pps APC, P. Antitumor and Immunomodulatory Activity of Phallusia Nigra Savigny, 1816 Against Ehrlich Ascites Carcinoma. Res. J. Pharm. Sci. 2012, 1, 7–12. [Google Scholar] [CrossRef]
  192. Janakiram, N.B.; Mohammed, A.; Rao, C.V. Sea Cucumbers Metabolites as Potent Anti-Cancer Agents. Mar. Drugs 2015, 13, 2909–2923. [Google Scholar] [CrossRef] [Green Version]
  193. Pettit, G.R.; Herald, C.L.; Doubek, D.L.; Herald, D.L.; Arnold, E.; Clardy, J. Isolation and structure of bryostatin-1. Am. Chem. Soc. 1982, 104, 24–6846. [Google Scholar] [CrossRef]
  194. Raghuvanshi, R.; Bharate, S.B. Preclinical and Clinical Studies on Bryostatins, A Class of Marine-Derived Protein Kinase C Modulators: A Mini-Review. Curr Top Med Chem. 2020, 20, 1124–1135. [Google Scholar] [CrossRef] [PubMed]
  195. Trenn, G.; Pettit, G.R.; Takayama, H.; Hu-Li, J.; Sitkovsky, M.V. Immunomodulating Properties of a Novel Series of Protein Kinase C Activators. The Bryostatins. J. Immunol. 1988, 140, 433–439. [Google Scholar] [PubMed]
  196. Hornung, R.L.; Pearson, J.W.; Beckwith, M.; Longo, D.L. Preclinical Evaluation of Bryostatin as an Anticancer Agent against Several Murine Tumor Cell Lines in Vitro versus in Vivo Activity. Cancer Res. 1992, 52, 101–107. [Google Scholar] [PubMed]
  197. Mary Varterasian, B.L.; Mohammad, R.M.; Eilender, D.S.; Hulburd, K.; Rodriguez, D.H.; Pemberton, P.A.; Pluda, J.M.; Dan, M.D.; Chen, B.D.; Al-Katib, A.M. Phase I Study of Bryostatin 1 in Patients with Relapsed. Non-Hodgkin’s Lymphoma and Chronic Lymphocytic Leukemia. J. Clin. Oncol. 1998, 16, 56–62. [Google Scholar] [CrossRef]
  198. Rinehart, K.L.; Holt, T.G.; Fregeau, N.L.; Stroh, J.G.; Keifer, P.A.; Sun, F.; Li, L.H.; Martin, D.G. Ecteinascidins 729, 743, 745, 759A, 759B, and 770: Potent antitumor agents from the Caribbean tunicate Ecteinascidia turbinata. J. Org. Chem. 1990, 55, 4512–4515. [Google Scholar] [CrossRef]
  199. Larsen, A.K.; Galmarini, C.M.; D’Incalci, M. Unique features of Trabectedin mechanism of action. Cancer Chemother. Pharmacol. 2016, 77, 663–671. [Google Scholar] [CrossRef] [PubMed]
  200. Belgiovine, C.; Frapolli, R.; Liguori, M.; Digifico, E.; Colombo, F.S.; Meroni, M.; Allavena, P.; D’Incalci, M. Inhibition of tumor-associated macrophages by trabectedin improves the antitumor adaptive immunity in response to anti-PD-1 therapy. Eur. J. Immunol. 2021, 51, 2677–2686. [Google Scholar] [CrossRef] [PubMed]
  201. Banerjee, P.; Zhang, R.; Ivan, C.; Galletti, G.; Clise-Dwyer, K.; Barbaglio, F.; Scarfò, L.; Aracil, M.; Klein, C.; Wierda, W.; et al. Trabectedin Reveals a Strategy of Immunomodulation in Chronic Lymphocytic Leukemia. Cancer Immunol. Res. 2019, 7, 2036–2051. [Google Scholar] [CrossRef]
  202. D’Ippolito, G.; Iadicicco, O.; Romano, G. Fontana A Detection of short-chain aldehydes in marine organisms: The diatom Thalassiosira rotula. Tetrahedron. Lett. 2002, 43, 6137–6140. [Google Scholar] [CrossRef]
  203. Cutignano, A.; D’Ippolito, G.; Romano, G.; Lamari, N.; Cimino, G.; Febbraio, F.; Nucci, R.; Fontana, A. Chloroplastic Glycolipids Fuel Aldehyde Biosynthesis in the Marine Diatom Thalassiosira rotula. ChemBioChem 2006, 7, 450–456. [Google Scholar] [CrossRef]
  204. D’Ippolito, G.; Romano, G.; Caruso, T.; Spinella, A.; Cimino, G.; Fontana, A. Production of octadienal in the marine diatom Skeletonema costatum. Org. Lett. 2003, 5, 885–887. [Google Scholar] [CrossRef] [PubMed]
  205. D’Ippolito, G.; Tucci, S.; Cutignano, A.; Giovanna, R.; Cimino, G.; Miralto, A.; Fontana, A. The role of complex lipids in the synthesis of bioactive aldehydes of the marine diatom Skeletonema costatum. Biochim. Biophys Acta–Mol. Cell Biol. Lipids 2004, 1686, 100–107. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The “immune synapse”: APC: antigen presenting-cell; PD-L1: programmed death ligand-1; PD-1: programmed death-1; CTLA-4: cytotoxic T-lymphocyte antigen-4; TCR: T cell receptor; MHC: major histocompatibily complex; CD28, CD40, CD40L, CD80/86: cluster of differentiation.
Figure 1. The “immune synapse”: APC: antigen presenting-cell; PD-L1: programmed death ligand-1; PD-1: programmed death-1; CTLA-4: cytotoxic T-lymphocyte antigen-4; TCR: T cell receptor; MHC: major histocompatibily complex; CD28, CD40, CD40L, CD80/86: cluster of differentiation.
Marinedrugs 20 00386 g001
Figure 2. Terpenes from plants with anticancer and immunomodulatory properties.
Figure 2. Terpenes from plants with anticancer and immunomodulatory properties.
Marinedrugs 20 00386 g002
Figure 3. Phenolic compounds from plant with anticancer and immunomodulatory properties.
Figure 3. Phenolic compounds from plant with anticancer and immunomodulatory properties.
Marinedrugs 20 00386 g003
Figure 4. Potential active compounds from spices (black pepper, cardamom, and curcuma).
Figure 4. Potential active compounds from spices (black pepper, cardamom, and curcuma).
Marinedrugs 20 00386 g004
Figure 5. Triterpenoid compounds from macromycetes with anticancer and immunomodulatory properties.
Figure 5. Triterpenoid compounds from macromycetes with anticancer and immunomodulatory properties.
Marinedrugs 20 00386 g005
Figure 6. Potential anticancer and immunomodulatory compounds from propolis.
Figure 6. Potential anticancer and immunomodulatory compounds from propolis.
Marinedrugs 20 00386 g006
Figure 7. Anticancer and immunomodulatory compounds from marine environment.
Figure 7. Anticancer and immunomodulatory compounds from marine environment.
Marinedrugs 20 00386 g007
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Nuzzo, G.; Senese, G.; Gallo, C.; Albiani, F.; Romano, L.; d’Ippolito, G.; Manzo, E.; Fontana, A. Antitumor Potential of Immunomodulatory Natural Products. Mar. Drugs 2022, 20, 386. https://doi.org/10.3390/md20060386

AMA Style

Nuzzo G, Senese G, Gallo C, Albiani F, Romano L, d’Ippolito G, Manzo E, Fontana A. Antitumor Potential of Immunomodulatory Natural Products. Marine Drugs. 2022; 20(6):386. https://doi.org/10.3390/md20060386

Chicago/Turabian Style

Nuzzo, Genoveffa, Giuseppina Senese, Carmela Gallo, Federica Albiani, Lucia Romano, Giuliana d’Ippolito, Emiliano Manzo, and Angelo Fontana. 2022. "Antitumor Potential of Immunomodulatory Natural Products" Marine Drugs 20, no. 6: 386. https://doi.org/10.3390/md20060386

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

Article Metrics

Back to TopTop