Skip to main content

Anticancer potential of algae-derived metabolites: recent updates and breakthroughs

Abstract

Background

Cancer is an increasing medical condition that poses a threat to worldwide populations, despite improvements in scientific research. For normal cancer treatment, a variety of chemotherapeutics, radiation, and medications are available; however, recurrent side effects and multi-drug resistance have limited treatment options and harmed our immune system. Marine algae are a promising source of novel components for the development of new complementary and alternative medications with anti-carcinogenic properties.

Results

In this review, we discussed several breakthrough studies on the anti-carcinogenic effects of several macro- and micro-algal components, demonstrating the inhibition of cancer cell development via multiple mechanisms. These components, often referred to as algal biopolymers, have been demonstrated to exhibit a wide range of chemical compositions and physical properties; as a result, they are used in pharmacological, pharmaceutical, nutraceutical, and microbiological applications in different sectors. Moreover, treatment of antimicrobial-resistant Helicobacter pylori infection-derived gastric cancer prevention may benefit from the use of algae in addition to standard antibiotics. Additionally, in recent years, it has been shown that algae have incredibly promising low-cost biomedical potentials as therapeutic applications for the treatment of cancer.

Conclusion

In recent years, several preclinical studies with the algal bioactive components in the field of novel drug discovery substituting synthetic drugs have been conducted. To demonstrate their potential anticancer actions on various cancerous signaling pathways and consequently reduce cancer, the enormous plasticity of these algae biopolymers has been intensively explored.

Background

Cancer is the world's second most prevalent debilitating disease, accounting for a significant share of all deaths. The multifactorial etiology of cancer encompasses a wide range of illnesses connected to the body's uncontrolled cell development [1]. The three cancer kinds that account for the bulk of instances worldwide are breast, lung, and colorectal cancers [2]. According to the International Agency for Cancer Research, 19.3 million cases of cancer were reported in 2020, and by 2040, that figure is expected to rise by 47% to 28.4 million cases [3]. Furthermore, facilitating replicative immortality, boosting angiogenesis, evading growth promoters, prolonging proliferative signals, resisting cell death, and initiating metastasis and invasion are all trademarks of cancer malignancy [4]. Although the development of new chemotherapeutic drugs for cancer treatment is crucial for halting the disease's progression, improving cancer therapies remains a challenging undertaking [5]. Chemotherapeutic resistance is a key barrier in the treatment of various cancers, as a large proportion of tumors relapse and develop resistance, inevitably leading to multi-drug resistance following exposure to multiple anticancer medications with similar structures and modes of action [6]. Over 195,000 plant-derived bioactive components have been identified as preventing cancer growth, either directly or indirectly through immune system activation [7]. The taxonomically diverse marine flora (microalgae, macroalgae, cyanobacteria, bacteria, actinobacteria, fungi, and other halophytes) constitute over 90% of the oceanic biomass offering a great scope of novel anticancer drugs [8]. Similarly, bioactive components found in algae have recently been identified as having anticancer properties by inducing apoptosis and suppressing cell divisions through interfering with signaling pathways [9]. Although, due to a lack of ethnomedical history, the creation of novel components from marine flora is still in its initial phases, leaving them under-represented in today's pharmacopeia [8]. Algal metabolites, also known as algal biopolymers, have been demonstrated to contain a diverse spectrum of chemical compositions and physical properties, and are thus used in pharmacological, pharmaceutical, nutraceutical, and microbiological applications in different sectors [10]. Later bioactive components found in algal metabolites (polysaccharides, proteins, polyunsaturated fatty acids (PUFAs), phycocolloids, vitamins, soluble dietary fibers, phycobilins, carotenoids, phycocyanins, minerals, tocopherols, and terpenes) have been shown to have biological therapeutic potential [11]. Further, the algae-derived bioactive components were later identified to antagonize cancer malignancy hallmarks [12]. Microalgae and macroalgae are the two types of algae that live in the sea. Microalgae are photosynthetic autotrophic microorganisms that contribute significantly to the marine food chain [13]. Of the top 10 producers, China leads with 54,850 tonnes, followed by Chile, Greece, Tunisia, Burkina Faso, Central African Republic, Chad, Bulgaria, and Spain. The total output of microalgae is expected to reach 56,456 tonnes globally [14]. They have been demonstrated to have substantial nutraceutical and therapeutic potential due to their high bioactive metabolite content [13]. Along with, cyanobacteria (blue-green algae), Spirulina sp., and Nostoc sp. bioactive components have medicinal values [15]. Furthermore, around 5000 years ago, Chinese physicians began using seaweeds, which are macroalgae that predominate the marine flora [13, 16]. The abundance/production of macroalgae worldwide is estimated at 35,762,504 tonnes (wet weight), with Asia contributing the majority of that amount (97.38%), followed by the Americas, Europe, Africa, and Oceania at estimated 1.36%, 0.80%, 0.41%, and 0.05%, respectively [14]. The potential for seaweeds to be used as several therapeutics has piqued the interest of scientists over thirty years. Additionally, seaweeds' medicinal and nutraceutical properties have been applied to the treatment of a number of diseases (stomach ailments, renal disorders, cancer, psoriasis, arteriosclerosis, lung diseases, cancer, gall stones, ulcers, heart disease, and scabies) [13, 17]. Overall algae are known to exhibit anti-tumor, anti-viral, antimicrobial, immune-boosting, and anti-inflammatory activity [15]. Although various research literature works have looked at the potency of anticancer substances, in this article we have focused on the comprehensive anticancer effectiveness of bioactive components derived from algae against a variety of cancer signaling pathways, including gastric cancer caused by the Helicobacter pylori bacteria, as well as various cutting-edge techniques in biomedical applications. Alongside, in this review, many other neoplastic indicators are highlighted in silico, in vivo, and in vitro for the identification of novel pharmaceuticals and biomedical treatments to be used in algae-derived cancer therapy in the near future (Figs. 1, 2).

Fig. 1
figure 1

Mechanism of action of algal bioactive components with anticancer potential via regulating the aberrant expression of cancer signaling pathways. Following a thorough analysis of the literature from articles, the figure is illustrated schematically [18,19,20,21]

Fig. 2
figure 2

Percentage of total algae-derived bioactive components [22,23,24,25,26,27]

Main text

Cancer biology: a molecular immunopathology

Cancer in humans has been prevalent for a long time, even before the advent of innovation and the use of synthetic substances. Percivall Pott discovered the first evidence of cancer in 1775 when he associated scrotal cancer and chimney soot. However, with the creation of improved scientific investigations, the mechanism of carcinogenesis has been widely studied. In 1971, a war on cancer was declared with the goal of generating new treatments [28, 29]. Cell division is the key phenomenon in the development of a living organism. Approximately 1015 cells are present in an adult human which exhibits cell turnover and regeneration due to the presence of stem cells having compartments with approximately 1012 divisions/day. Throughout an individual's life, several overlapping biological pathways regulate cell differentiation, balancing the ratio of cell proliferation and apoptosis. Any disruption in homeostasis causes neoplasia or uncontrolled cell proliferation [30]. Cancer has traditionally been studied through the lens of Darwin's three fundamental contextual evolutionary principles (variation, heredity, and selection), which Peter Nowell postulated to be an evolutionary process after analyzing carcinogenesis in advanced malignancies [31]. Furthermore, the mathematical idea of Darwinian evolution has been widely employed to comprehend somatic selection, diversity, and extinction [28]. A succession of gene mutations disturbs cellular function and creates gene dysfunction, resulting in cancer [18].

The intrinsic and non-intrinsic factors that link them to deoxyribonucleic acid (DNA) damage impacting cellular homeostasis due to discordant signaling pathways substantially influence the underlying etiology of carcinogenesis [32]. Random replication mistakes owing to spontaneous mutation are intrinsic factors, whereas proto-oncogene mutations are non-intrinsic factors. Radiation, chemical carcinogens, xenobiotics, a terrible routine, viruses, and other external and endogenous causes (hormone levels, abnormal immune system, biological metabolism, repair machinery, etc.) [32, 33]. Chemical carcinogens or xenobiotic components directly or indirectly affect the cellular cytoplasm and/or nucleus which induces proto-oncogenes leading to genetic disorders and mutations [18]. Apart from these carcinogenic factors, infectious oncogenic pathogens contribute 15% of global malignant tumor heterogeneity responsible for thousands of neoplasias [34]. Bacteria, helminths, and fungi cause inflammation and disease-mediated cancer, whereas oncoviruses cause carcinogenesis by oncogene integration with the host genome [35]. The oncopathogens in humans mostly cause organ-specific or site-specific carcinoma. Therefore, the virus-induced cancers include Human papillomavirus (HPV) which causes oropharyngeal, cervical, anal, and penile cancer; Human T-cell leukemia virus (HTLV)-1 which causes adult T-cell leukemia-lymphoma; Hepatitis B virus (HBV) which causes non-Hodgkin lymphoma, breast, hepatocellular, and pancreatic cancer; Hepatitis C virus (HCV) which causes non-Hodgkin lymphoma, thyroid, and liver cancer; Human immunodeficiency virus (HIV) which causes Kaposi sarcoma, non-Hodgkin lymphoma, lung, liver, anal, and oropharyngeal cancer; Epstein–Barr virus (EBV) which causes Burkitt’s lymphoma, gastric, smooth muscle, and nasopharyngeal cancer. Helicobacter pylori confer bacterial-induced carcinoma that causes gastric and pancreatic cancer. Among fungi, Aspergillus sp. causes liver cancer, and Candida sp. causes oral and lung cancer. Various helminths such as Schistosoma haematobium, Schistosoma japonica, Schistosoma mansoni, Plasmodium falciparum, Clonorchis sinensis, and Opisthorchis viverrini are also known to cause cancer [34, 35].

Overall oncogene activation is caused by mutations arising due to erroneous genetic alterations such as point mutations (G12V Ras gene), insertional inactivation (C-myc gene), deletion (Erb-B gene), amplification (N-myc), hypomethylation, hypermethylation, deacetylation, and chromosomal translocation (Abl and Bcr gene) [18, 36]. Furthermore, epigenetic silencing, promoter methylation, and the production of oncometabolite all play a part in oncogenesis [36]. In addition, under normal conditions, including the p21 gene, the p53 gene on human chromosome 17 favorably regulates DNA metabolism, cell differentiation, and cell death. When the p53 gene is altered, cancer cells in the G1 and G2 phases of the cell cycle are generated, followed by a relationship between cyclin-dependent kinase (CDK)1-P2 and cell division cycle (CDC)2. The p53 protein binds to DNA after other genes have produced DNA damage, causing the WAF1 gene to be stimulated. This action causes p53 to bind to CDK2, which then blocks the effect of p21 for the following juncture of the cell cycle. Furthermore, the anticancer effect of p53 causes apoptosis in addition to stopping the cell cycle throughout the G1/S phase [18]. The dysregulation, progression, and dissemination of cancer cells are fueled by several signaling pathways such as receptor-tyrosine kinase mitogen-activated protein /rat sarcoma virus /receptor tyrosine kinase (MAP/RAS/RTK)-kinase signaling, Hippo signaling, Notch signaling, Phosphoinositide 3-kinase (PI3K) signaling, oxidative stress response/Nrf, transforming growth factor-beta (TGFβ) signaling, PI3K-Akt signaling, nuclear factor-kappa B (NF-κB) signaling, β-catenin/wnt signaling, Jun N-terminal kinases (JNK)/p38 signaling, and Ras-extracellular signal-regulated kinase (Ras-ERK) signaling [36,37,38]. Carcinomatous signaling pathways were activated by these alterations, causing cancer cells to stop dying and proliferate by supplying them with extra metabolites [39]. These cancer cells spread and migrate by accessing the extracellular matrix (ECM), which leads to circulation by alternate migration such as collective cell, mesenchymal, and amoeboid cell migration, despite the fact that they are rarely investigated [40].

Cancer development eludes immune monitoring due to immune checkpoint dysregulation caused by malignancies. Furthermore, immune factor activity is suppressed by hyperactivation of signal transducer and activator of transcription (STAT)-3, a signal transducer and activator of transcription [41]. Both STAT-3 and NF-κB activate anti-apoptotic proteins (B-cell lymphoma (Bcl-2 and Bcl-XL)) that enhance tumor growth by interfering with p53 [42]. Neutrophils are a controversial topic due to their dual function, i.e., tumor-promoting and attacking plasticity. Angiogenesis, metastasis, and immunosuppression are all facilitated by tumor-associated neutrophils (TAN) [43]. Simultaneously, tumor-associated macrophages (TAM) inhibit T-cell and natural killer (NK) cell proliferation by releasing cytokines and immune-suppressive factors thereby stimulating tumor progression [44]. Cancer initiation, progression, and metastasis are all influenced by inflammation. Several mediator molecules, including tumor necrosis factor (TNF-α), NF-κB, and signaling pathways, link inflammation and cancer [45]. Inflammation promotes cancer cell proliferation by raising mutation rates, which are mostly caused by chemical carcinogens and pathogenic microorganisms [42]. Tumor cells boost neutrophil production by secreting growth factors (interleukins (IL)-3, granulocyte–macrophage colony-stimulating factor (GMCSF), and granulocyte colony-stimulating factor (GCSF) or inflammatory cytokines (IL-1/6/17 and TNF-α), which promote tumor progression by inducing cancer-related inflammation. Anti-tumor responses are mediated by TANs, which destroy tumor cells [43, 46]. Alongside, macrophages, dendritic cells, B cells, and T cells also exhibit dual functions as neutrophils. In contrast, mast cells and TH2 cells only promote tumorigenesis whereas NK cells only exhibit anti-tumor immunity [42]. Moreover, the major histocompatibility complex (MHC) system, cytokines, lymphocytes (B and T cells), and antigen-presenting cells (APCs) are also used by the host's adaptive immune system to recognize and kill tumor cells with abnormal cell surface antigens. As a result, using a functional adaptive immune system to target mediators and inhibit immunological checkpoints is a strong cancer therapeutic technique. In the near future, further development of cancer vaccines and modified T-lymphocytes will be the most effective technique for treating cancer with fewer/no side effects [45] (Fig. 3).

Fig. 3
figure 3

Schematic representation of the molecular mechanism of cancer topography. Following a thorough analysis of the literature from articles, the figure is illustrated schematically and adapted [4, 18, 47, 48]

Anticancer potentials of algae-derived metabolites

Microalgae anticancer potential

Microalgae are photosynthetic microorganisms that are categorized into prokaryotic (Cyanobacteria) and eukaryotic microalgae (diatoms, dinoflagellates, and coccolithophores). These phytoplanktons, which are found in practically all biomes (temperate to extreme) can be widely classified as fresh and marine water microalgae, providing up to 40% of global productivity [19]. These microalgae can be widely classified as autotrophic, heterotrophic, oligotrophic, and mixotrophic depending on their nutrient requirements [49]. Microalgae can be used to produce a wide range of bioactive compounds with various biotechnological purposes. They can be grown easily in photobioreactors and have quick generation times. Several factors influence the bioactive potential of microalgae, including species, growth phase, and culture conditions (temperature, nutrient availability, and light conditions). Although, due to its tremendous prominence in the field of biofuel production, microalgae's medical potential has been overlooked more than that of macroalgae [50]. Microalgae, in addition to marine bacteria and fungi, are ecologically important as producers and decomposers in the aquatic environment. Second, after food and biorefinery, their metabolic plasticity may stimulate therapeutic development to combat a diversity of diseases [51]. These algae have been shown to produce a variety of bioactive components (carotenoids, polysaccharides, and fatty acids) that have gained popularity due to their antimicrobial and antioxidant characteristics [19]. In recent eras, new therapeutic components can be developed and synthesized from natural resources by means of modern technology. Using the Discovery Studio 3.1 platform, the 3D models of the ligand ((9-Ethyliminomethyl-12-(morpholin-4-ylmethoxy)-5,8,13,16-tetraaza-hexacene-2,3-dicarboxylic acid) EMTAHDCA) obtained from cyanobacterium Nostoc sp. MGL001 were found to have a functional resemblance to existing drugs against 11 cancer-related proteins [52]. However, the in silico characterization of anticancer bioactive components from microalgae has only lately been addressed to a limited extent. Further, the microalgal compounds also have anti-inflammatory and immunomodulatory characteristics, making them a potential immunotherapeutic weapon against cancer. Sulfo-polysaccharides, PUFAs, sulfated lipids, and carotenoids (astaxanthin) are all microalgal immune-stimulatory components that drive macrophage and dendritic cell proliferation and maturation. Apart from APCs, the dendritic cells often entitled “nature's adjuvant,” are known to trigger cytotoxic T-lymphocytes, ultimately culminating in neoplastic cell death [53]. However, the current study indicates that microalgae might be a source of cancer treatments that work by promoting natural killer cells production, apoptosis-mediated cell death, cell cytotoxicity, and reducing tumor cells invasion either via a caspase-dependent or caspase-independent mechanism [19].

Microalgal carotenoids

Carotenoids from microalgae have been identified as a potential regime for treating inflammatory disease and cancer. Chlorella sp. (Chlorella sorokiniana, Chlorella zoofingiensis, Chlorella vulgaris, Auxenochlorella prothecoides, Auxenochlorella pyrenoidosa, Chlorodium saccharophilum, Jaagichlorella luteoviridis) is the major source of carotenoids followed by Arthospira platensis (cyanobacteria), Dunaliella salina, Chlamydomonas reinhardtii, Tetraselmis suecica, Tetraselmis striata, Scenedesmus quadricauda, Dactylococcus dissociates, Asterarcys quadricellulare, Odontella aurita, Chlorobotrys regularis, Isochyrsis galbana, Chlorobotrys gloeothece, Nitzschia laevis, Chaetoceros neogracili, Munda aquilonaris, Phaeodactylum tricornutum, Porphyridium purpureum, Cantharellus cinnabarinus, Haematococcus lacustris, etc. Carotenoids have been studied in vitro, in vivo, and in humans for their anti-inflammatory, anti-tumor, and anticancer activities [20, 54].β-carotene derived from D. salina, C. reinhardtii, T. suecica, I. galbana exhibit anticancer potentials against neuroblastoma, non-Hodgkin lymphoma, prostate, breast, liver, pancreatic, colorectal, and gastric cancer [54]. Among other microalgae that have been shown to kill prostate cancer cells by triggering apoptosis, D. salina is the main source of β-carotene [55]. Studies conducted in vitro indicate that β-carotene has been found to inhibit the Ku proteins, M2 macrophage polarization, and NF-κB activation [56]. Additionally, caveolin-1 protein expression, calcium/calmodulin-dependent protein kinase IV activity, the NF-κB/Akt pathway, the PI3K/Akt pathway, and the ERK pathway are all downregulated by β-carotene's antiproliferative activity, thereby arresting cell cycle and inducing apoptosis [54]. The inhibition of these signaling pathways has arrested the in vitro growth of distinct cancer cells/cell lines such as colorectal cancer cells (HT-29, Caco-2), hepatic cancer cells (HepG2, SK-Hep-1), colon cancer cells (HCT116), H. pylori-infected gastric cancer cells, esophageal carcinoma cells, adrenocorticotropic hormone-secreting pituitary adenoma cells (AtT-20) (further inhibiting cervical, breast, and hepatoma cancer cells), and lymphoblast cells (K562) apoptosis [54]. In in vivo murine model studies, β-carotene administration for a specific time period resulted in various anticancer actions. For example, β-carotene administration for 11 weeks suppresses M2 macrophage polarization thereby reducing colitis-associated colon malignancy [56]. Alongside, β-carotene's anticancer potential demonstrates DNA methylation, epigenetic modulation, and miRNA expression, all of which reduce the ability of colon cancer stem cells to proliferate and self-renew [57]. Oral treatment of β-carotene reduced the tumor weight of rat models suffering from liver cancer in hepatic cell lines (H22) [58]. In humans, β-carotene has been linked to the prevention of numerous malignancies due to its powerful antioxidant properties that reduce reactive oxygen species (ROS) formation, although further research is required to fully comprehend their potential [54].β-cryptoxanthin obtained from P. trichornutum, A. pyrenoidosa, P. purpureum, and Cyanophora paradoxa has been identified with antiproliferative, anti-migratory, and anticancer potentials. In vitro analysis has demonstrated suppressed migration, inhibition, and cell viability with increased apoptosis in the lung, colon (HCT116), and gastric cancer cells [54]. By causing caspase and cytochrome C mediated apoptosis as well as halting the cell cycle at the G0/G1 phase in a nude mouse xenograft, in vivo murine trials with β-cryptoxanthin treatment for 20 days have reduced angiogenesis and tumorigenesis of gastric cancer [59]. Furthermore, β-cryptoxanthin supplementation at 10 and 20 mg/kg inhibited tumor growth by downregulating sirtuin-1 (SIRT-1), retinoic acid receptor-β, and p53 [60], while 1 and 10 mg/kg treatment inhibited nicotinic acetylcholine receptor α7, both of which suppressed lung cancer in mice [61]. In human studies, β-cryptoxanthin reduced the risk of non-Hodgkin lymphoma, lung, breast, renal, head, and neck cancer [54]. Additionally, β-cryptoxanthin has been shown to cause apoptosis in human skin, lung, breast, and HeLa cancer cells and demonstrate cytotoxicity [20]. Further combination treatment with a chemotherapeutic drug (oxaliplatin) in colon cancer reduces the drug’s toxicity [62].

Astaxanthin, a carotenoid with a potent antioxidant potential, shields cells from cyto- and genotoxicity brought on by ROS, epigenetic changes, cell cycle arrest in the G0/G1 or G2/M phase, activation of anti-apoptotic proteins, blocking angiogenesis and metastasis, and chromatin remodeling, ultimately enhancing tumor immunity [19, 63]. Haematococcus pluvialis is the predominant microalgal supplier of astaxanthin followed by Tetraselmis sp., G. sulphuraria, Chlorococcum sp., C. sorokiniana, and C. Zofingiensi [20]. This carotenoid induces several tumor suppressors (MAPK4, mapsin, breast cancer metastasis suppressor 1, and kail) [64]. Apoptosis is triggered by stopping the cell cycle in a number of in vitro experiments on various cell lines. Astaxanthin induced cytotoxicity against ovarian cell lines by inhibiting NF-κB and stimulating apoptosis [65]. According to research, astaxanthin has the ability to decrease angiogenesis and metastasis in a variety of cell lines, including glioblastoma, murine hepatoma cells (H22), and adenocarcinoma gastric cell lines (AGS, KATO-III, MKN-45, and SNU-1). It also has the ability to control epigenetic changes [54]. In colon cancer, astaxanthin therapy was shown to downregulate Akt phosphorylation, cyclin D1, and Bcl-2 expression, as well as promote the production of p53, p21, p27, Bax, and caspase-3 [66]. Several dose-dependent in vivo applications of astaxanthin showed an anticancer effect on different malignancies such as gastrointestinal cancer (downregulates ERK-2, NF-κB, and cyclooxygenase (COX)-2; activates apoptosis), colon cancer (downregulates NF-κB and oxidative stress markers), oral cancer (downregulates Wnt/B-catenin and NF-κB signaling), esophageal cancer (downregulates NF-κB and COX-2), hepatic cancer (downregulates oxidative stress; upregulates serum adiponectin protein), skin cancer (downregulates tyrosinase activity), and lung metastatic myeloma (downregulates Bcl-2, ERK, and NF-κB; upregulates apoptosis) in murine models [54]. Moreover, astaxanthin is a good antioxidant agent that has been identified to elevate IL-6 and TNF-α in murine models prior to tumor initiation [67]. In human studies, astaxanthin is majorly evidenced to inhibit immune dysfunction alongside regulating the inflammatory response [68]. Additionally, the antiproliferative effect of carbendazim in MCF-7 cells in the G2/M phase is enhanced by the addition of astaxanthin [69]. Nevertheless, there is still a paucity of information on astaxanthin-related dose-dependent human cancer investigations.

Lutein obtained from C. sorokiniana, C. zoofingiensis, A. protothecoides, D. salina, T. suecica, and C. reinhardtii has demonstrated anticancer and anti-proliferative activity against non-Hodgkin lymphoma, renal cell carcinoma, hepatocellular carcinoma, pharyngeal, esophageal, neck, pancreatic, colon, bladder, and breast cancer [54]. According to in vitro studies in breast cancer lines (MCF-7 and MDA-MB-231), lutein inhibits transcription factor Nrf2 (including genes superoxide dismutase (SOD)-2 and HO-1), glycolysis, cell growth, and progression, as well as down-regulating NF-κB, pAkt, and pERK markers, inducing p53 signaling, transcription factor hairy and enhancer of split (HES)-1, and cellular apoptosis [70]. Other cell lines, including prostate cancer (PC)-3, sarcoma S180, lung cancer A549, colon adenocarcinoma, and leukemia cells, were also investigated to determine lutein's anticancer activity [54]. Lutein's anti-proliferation slows the progression of the cancer cell cycle by downregulating biomarker genes in prostate cancer and culminates breast cancer by upregulating pro-apoptotic genes and p53 signaling pathway inducing apoptosis alongside downregulating Bcl-2 genes further generating ROS [20]. In vivo studies reported lutein administration of 50 mg/Kg for 1 month alongside 4T1 cells injection, inhibiting breast cancer in murine models [71]. Similarly, 0.002% of dietary lutein downregulated cell proliferative proteins (β-catenin, K-ras, and Akt/protein kinase B) thereby reducing tumor formation [72]. Alongside the suppression of cytochrome P450 phase I enzyme in N-nitrosodiethylamine-stimulated hepatocellular carcinoma was also observed via lutein administration in murine models [73]. Further coadministration of lutein with doxorubicin exhibited higher inhibition of sarcoma S180 cells proliferation in mice [74]. The human dietary consumption of lutein has reduced the efficacy of different cancers which are discussed before in this topic.

Zeaxanthin is a xanthophyll mostly obtained from Nannochloropsis oculata, Chloroidium saccharophilum, and Dunaliella sp. with good anticancer potentials. Few in vitro, in vivo, and human investigations have examined the chemopreventive activity of zeaxanthin, despite its limited research [54]. This carotenoid has been identified to activate gastric cancer cell apoptosis by upregulating pro-apoptotic factors and MAPK signaling pathway alongside downregulating anti-apoptotic factors (Bcl-2) [75]. The anti-melanoma potential of zeaxanthin has also activated human uveal melanoma cells apoptosis by downregulating the melanoma cell-induced fibroblast migration and platelet-derived growth factor [54].

Many microalgae also contain the orange-colored marine xanthophyll molecule known as fucoxanthin such as Chaetoceros neogracili, Isochrysis sp., Cylindrotheca closterium, Pleurochrysis carterae, Odontella aurita, Phaeodactylum tricornutum, Nitzschia laevis, Conticribra weissflogii, Tisochrysis lutea, Thalassiosira sp., Navicula sp., Amphora sp., and Pavlova sp., OPMS 30543 [20, 54]. The anticarcinogenic characteristics of fucoxanthin include decreased tumor incidence, cancer cell inhibition, cell cycle arrest, induction of apoptosis, and controlled metastasis. Furthermore, Bcl-2 protein, caspase pathway (caspase-3, caspase-8, caspase-9), signaling pathways (MAPK, JAK/STAT, stress-activated protein kinases (SAPK)/JNK, and PI3K/Akt/mechanistic target of rapamycin (mTOR)), growth arrest and DNA-damage-inducible protein (growth arrest and DNA damage (GADD)45α), NF-B, CYP3A4 enzyme, connexin genes, expression of N-myc oncogene, angiogenesis, and survival are all involved in fucoxanthin-induced apoptosis. Alongside apoptosis, fucoxanthin also confers chromatin condensation, DNA laddering, and degradation [20, 76]. According to in vitro research, various cancer cell lines have shown fucoxanthin to have anticancer potential. Regarding gastric cancer, fucoxanthin suppresses myeloid cell leukemia 1 protein and cyclin B1 via JAK/STAT signaling pathway alongside the reduction in Bcl-2 thereby inducing autophagy and apoptosis by stimulating cleaved caspase-3, beclin-1, and microtubule-associated protein 1 light chain 3 [77, 78]. Similarly, fucoxanthin's anticancer properties reported beta-glucuronidase activity and NF-κB mediated pro-apoptotic activity in DLD-1 and HCT116 colorectal cancer cells, respectively [79, 80]. The further combined therapy of fucoxanthin with 5-fluorouracil exhibited a cytotoxic effect on both HCT116 and HT29 cell lines [81]. Alongside, the antiproliferative potential of fucoxanthin has been observed to downregulate the NF-κB pathway/expression in hepatic carcinoma (HepG2), Burkitt’s and Hodgkin’s lymphoma, and breast cancer (MCF-7 and MDA-MB-231) cells. Fucoxanthin's ability to kill human cervical cancer cell lines (HeLa) has also been linked to the downregulation of the Akt/mTOR pathway, PI3K/Akt, NF-κB, and a member of the histone cluster 1 H3 family [54]. Fucoxanthin has also been found to boost GADD45 expression in HepG2 and HTLV-1-infected T cells, causing G1 cell cycle arrest [82]. Regarding lung cancer, fucoxanthin has been identified to exhibit inhibitory effects by upregulating the pro-apoptotic p53 gene and Fas, alongside suppressing Bcl-2 [83]. Moreover, activation of different cell lines via mortalin (anti-apoptotic)-p53 binding can be suppressed via fucoxanthin application [84]. Regarding the central nervous system, fucoxanthin not only modulates the MAPK pathway but also downregulates PI3K/Akt/mTOR and p38 signaling pathway thereby stimulating ROS-triggered apoptosis by reducing invasion, angiogenesis, and cell proliferation. Based on in vivo studies, fucoxanthin has shown good chemopreventive potentials against colon cancer, lung cancer, hepatocellular carcinoma, cervical cancer, adenocellular carcinoma, and various tumor xenografts in various murine or rat models [54].

Violaxanthin, a compound isolated from Dunaliella tertiolecta, induces apoptosis in MCF-7 breast cancer cells without fragmenting DNA. Alongside, violaxanthin from Chlorella ellipsoidea exhibits apoptosis in colon cancer cells [20]. This carotenoid also results in the reversion of multi-drug resistance (MDR) in human MDR1 gene-transfected mice lymphoma cells (L1210) and human breast cancer cells (MDA-MB-231 and MCF-7) [51]. Furthermore, it has been demonstrated that violaxanthin from Eustigmatos cf. polyphem has radical scavenging activity [20].

Neoxanthin, being a xanthophyll carotenoid has been evidenced to upregulate cytotoxic effect upon treatment on HeLa and A549 cancer cells [51].

Siphonaxanthin, a keto-carotenoid obtained from Codium fragile, Caulerpa lentillifera, and Umbraulva japonica, has been evidenced with anticancer potential on various cancer [51]. Regarding the human leukemia cell line (HL-60), siphonaxanthin induces apoptosis by downregulating Bcl-2 expression. Simultaneously the condensation of chromatin, GADD45α, and apoptosis-inducing death receptor-5 (DR5) are upregulated [82]. Moreover, the anti-angiogenic effect of siphonaxanthin exhibits downregulated expression of mRNA, fibroblast growth factor receptor (FGFR)-1, early growth response (EGR)-1, and fibroblast growth factor (FGF)-1 [51].

Canthaxanthin is a keto-carotenoid primarily obtained from the mushroom Cantharellus cinnabarinus. Later, this carotenoid was also found in microalgae such as Dactylococcys dissociates, H. pluvialis, Chlorella emersonii, C. zofingiensis, Coelastrella sp., Chlorococcum sp., and cyanobacteria (Aphanizomenon flos-aqua, Trichormus variabilis, Nodularia spumigena, and Anabaena sp.). This carotenoid is known to exhibit anti-tumorigenic, chemopreventive, and antioxidant activity against human colon adenocarcinoma, melanoma cells, prostate cancer cells, and in vitro oral cancer [20].

Microalgal polysaccharides

Polysaccharides derived from microalgae are broadly classified as intracellular and extracellular (structural/cell-bound/cell wall) polysaccharides. The important parameters used in microalgae cultivation boost biomass productivity. Microalgae, on the other hand, produce fewer exopolysaccharides (EPS) than bacteria under normal growth conditions [85]. Furthermore, stress and limited nutrient availability have been shown to increase EPS content in microalgae. The primary EPS composition of microalgae includes polysaccharides, lipids, DNA, and proteins [86]. As a result, two-stage cultivation is required for efficient polysaccharide production. Although microalgae polysaccharides are mostly used for industrial purposes, their biostimulant characteristics have been related to anticancer properties [87]. Sulfate concentration and molecular weight affect polysaccharide potentiality. Therefore, the polysaccharides obtained from several microalgae/cyanobacteria are Chlorella vulgaris, Chlorella pyrenoidosa, Arthospora platensis, Dixoniella grisea, Neochloris oleoabundans, Nostoc carneum, Porphyridium aerugineum, Dunaliella salina, Phaeodactylum tricornutum, Haematococcus pluvalis, Botryococcus braunii UC 58, Nostoc flageliforme, Rhodella violacea, Chlamydomonas reinhardtii, Anabaena sp. 33,047, Gloeocapsa sp., Graesiella sp., Spirulina sp. LEB18 [85]. The sugar composition of these microalgae includes glucose, fructose, xylose, fucose, arabinose, rhamnose, mannose, galactose, maltose, and lactose. The inclusion of uronic acids, pyruvates, and carbohydrate acyl groups thus gives EPS its anionic properties [88]. In contrast to macroalgae, microalgae have a lower proportion of sulfated and methylated polysaccharides. The partially purified EPS obtained from C. pyrenoidosa FACHB-9, Scenedesmus sp., and Chlorococcum sp. has been explored with radical scavenging generating abilities and anti-tumor activities (inhibiting cell viabilities and reducing colony count) upon treatment on HCT116 and HCT8 cell lines [89]. In vivo and in vitro studies using Graffi myeloid tumors, sarcoma S180 tumor cells, and breast cancer cells revealed additional anti-proliferative, anti-tumor, immunostimulatory, and cytotoxic characteristics of EPS produced from Porphyridum cruentum [90]. Simultaneously, nostoglycan, derived from the microalgae Nostoc sphaeroids, has been shown to enhance caspase-3-dependent apoptosis, limiting lung cancer cell proliferation while also protecting against ROS generation [91]. Moreover, dinoflagellate Gymnodinium sp. A3 EPS (GA3P (D-galactan sulfate, associated with L-( +)-lactic acid)) has been identified with both anticancer and enzyme inhibition (DNA topoisomerase I and II) activity [92]. Microalgae's polysaccharides have anticancer, antibacterial, and anti-adhesion capabilities that have been demonstrated to be crucial in the management of gastric carcinoma brought on by H. pylori [93]. Additionally, the anticancer and anti-proliferative properties of chrysolaminarin polysaccharides derived from the diatom Synedra acus have been demonstrated in HCT116 and DLD-1 cell lines [94]. Contrarily, despite a paucity of data and information, Navicula sp., Tribonema sp., and P. cruentum microalgal sulfated polysaccharides (SPs) have been investigated for anticancer potential in vivo and in vitro [20]. Nonetheless, the anticancer potentials of microalgal polysaccharides have received far less attention, with far fewer data available than for macroalgal polysaccharides.

Microalgal peptides

Therapeutic peptides, which are known to offer greater advantages than antibodies or proteins, have lately been explored in microalgae [45]. Enzymatically degraded microalgal byproducts produced from protein hydrolysates are the most common source of these bioactive peptides with unique amino acid residues. As a result, antiproliferative, antioxidant, and anti-microtubule action has been demonstrated on numerous cancer cell lines using these isolated therapeutic bioactive peptides [95]. Additionally, peptide-driven immune responses in cancer patients have produced previously unheard-of reactions. Microalgae hold great promise for the extraction of bioactive peptides for cancer treatment due to their accessibility and inexpensive cost [45]. However, only a few microalgae have been recognized as containing bioactive peptides exhibiting anticancer potential. Among all other microalgae, Chlorella sp. (C. vulgaris, C. sorokiniana, and C. pyrenoidosa) is mostly used for the production of bioactive peptides followed by Dunaliella sp. and Pavlova lutheri [45, 96]. Biologically active peptides extracted from C. vulgaris pepsin hydrolysate induced anti-proliferation and death of AGS cells after 24 h of exposure, arresting cell growth after the G1 phase. Additionally, antioxidant characterization showed that peptide-induced ROS generation is accountable for a number of harmful events in biological systems, including the attack on crucial biological components (DNA, protein, and lipid), and has been suggested as a prospective chemopreventive therapeutic for gastric carcinoma [97]. Human liver cancer cells (HepG2) were shown to be inhibited by enzymatic hydrolyzed derived polypeptides from C. pyrenoidosa by triggering apoptosis and necrotic death. The altered modifications, such as cell membrane shrinkage, nuclear condensation and disintegration, and the generation of black apoptotic bodies, were corroborated using phase-contrast microscopy [98]. Furthermore, malignant tumors gain the ability to spread by generating numerous metalloproteinases (MMP) that promote tumor migration and invasion, considering them potential targets for cancer treatment. Human fibrosarcoma (HT1080) cells are inhibited by bioactive peptides derived from P. lutheri via suppressing mRNA and MMP-9 expression [45, 99]. Tyrosinase activity can be decreased to lessen the risk of melanoma, which is brought on by UV radiation exposure that damages DNA. In mouse melanoma (B16F10) cells, bioactive peptides from P. lutheri have been shown to limit tyrosinase and melanogenesis activity, as well as reduced ROS generation, by boosting ERK phosphorylation [100]. Dolastatins derived from Lyngbya sp. and Symploca sp. has been shown to inhibit ovarian and cancer cell lines in humans. In addition, a dolastatin 10 derivative (TZT-1027) suppresses solid tumors (B-16 melanoma, colon 26 adenocarcinoma, M5076 sarcoma, and human cancer xenograft) in mouse models. Furthermore, although being less potent than dolastatin, auristatin PYE had better outcomes against colon cancer cells (DLD-1, HT 29, and COLO 205) [19]. Apart from dolstatins, grassypeptolide and curacin A are other bioactive peptides obtained from cyanobacterium L. confervoides and L. majuscule, respectively. Moreover, a wide variety of cyanobacterium-derived peptides (apratoxin (A-D, F), aurilides, coibamade A, lyngbyabellin (A, B, E, F, G, H, I, N), hoiamide (A-B), homodolstatin 16, largazole, obyanamide, majusculamide C, desmethoxymajusculamide, Palau amide, palmyramide, pitipeptolide (A and B), ulongapeptin, tasipeptin (A-B), veraguamide (A-G), wewak peptin (A-D), nostocyclopeptide (A1-A2), symplocamide A, belamide A, etc.) have been investigated with anticancer potentials against different cancer cell lines [51]. However, only a few studies on the anticancer activities of microalgal peptides have been conducted, with positive results on six different cancer types [45]. Phycocyanin from cyanobacteria (Arthospora platensis) and red algae are phycobiliproteins that have been studied for their ability to stop cell cycle (G0/G1 or G2/M phase), reduce Bcl-2/Bax, COX-2, p-ERK, PEG2, CDK4, cyclin D1, NF-B, Fas, p53, ICAM-1, CD44, chromatin condensation, Cyt c release. By suppressing the Akt/mTOR/p70S6K pathways, phycocyanin also inhibits angiogenesis and metastasis while also inducing autophagy [101]. Furthermore, amino acid supplementation has been shown to reduce muscle protein breakdown while also suppressing inflammation. It has been discovered that microalgae contain glutamic acid in addition to 18 other amino acids. Along with glycine, C. vulgaris and C. sorokiniana have higher levels of alanine, valine, and leucine. Furthermore, antioxidant-active Mycosporine-like amino acids (MAA) are abundant in Glenodinium foliaceum, Scenedesmus sp., and C. sorokiniana [20].

Microalgal lipids

Microalgae lipids are classified into two types: polar (glycerophospholipids) and non-polar (triacylglycerols). Long-chain fatty acids combine with polar lipids to generate PUFAs, which are divided into three classes: Docosahexaenoic acid (DHA), Docosapentaenoic acid (DPA), and Eicosapentaenoic acid (EPA). Non-polar lipids, on the other hand, are primarily involved in energy conservation. Polar lipids are involved in the functioning of cellular signaling pathways in addition to maintaining structural integrity and membrane fluidity [102]. EPA and DHA are the omega (ω)-3 PUFAs obtained from Porphyridium sp., Phaeodactylum sp., Nannochloropsis sp., Skeletonema sp., Thalassiosira sp., Cryptomonas sp., Tetraselmis sp., Heterocapsa niei, Isochrysis sp., and Chaetoceros sp. [20]. DHA is the largest ω-3 (n-3) fatty acid among all PUFAs, and it has been demonstrated to have anti-tumor effects by triggering apoptosis via regulating the nucleus and mitochondria, culminating in lipid peroxidation (generating ROS) and cell cytotoxicity [19]. Alongside, PUFAs' anti-angiogenic characteristics aid in the generation of anti-metastatic activity in many malignancies. Moreover, PUFAs with a double bond location n-3 (EPA and DHA) have been investigated to confer better anticancer activity compared to PUFAs with n-6 (ω-6). Unlike unsaturated lipids, saturated lipids with shorter chain lengths (≤ C10) are only known to demonstrate anti-tumor activity [103]. Multiple cancer cell lines, including breast cancer (MDA-MB-231, MCF-7, and KPL-1), prostate cancer, pancreatic cancer, and colon cancer (ACL-15 and HT-29), have been linked to dietary supplementation with n-3 PUFAs. In contrast, the anti-tumorigenic property of n-6 PUFAs has been disputed, as it has been shown in numerous human studies to promote carcinogenesis, which is inhibited when n-3 PUFAs are consumed [103, 104]. However, to date, inadequate data are available suggesting n-3 PUFA’s anticancer potentials against skin carcinoma [105]. Alongside, atherosclerosis, increased pro-inflammatory eicosanoids/cytokines, cardiovascular and autoimmune diseases can all result from an excess of n-6 PUFAs consumption [104]. Additionally, DHA-mediated apoptosis is promoted in gastric cancer by activating JNK, ERK, and actuator protein (AP)-1, halting cell growth by increasing the levels of p53, Bax, and intracellular cytochrome c [106]. Among the n-3 PUFAs, DHA and EPA have been examined for their capacity to elicit cell cycle arrest in regard to ROS production, which downregulates death-regulating factors (Bcl-2) and releases mitochondrial cytochrome c to the cytoplasm, activating intrinsic pathway-induced caspase-dependent cytotoxicity [107, 108]. When cytochrome c is released as a result of stress-induced mitochondrial permeabilization, it activates caspase-3 by attaching to the N-terminal caspase-recruitment domain (CARD), which then activates caspase-9 by recruiting to the apoptosome, resulting in biochemical and cellular apoptosis [109]. Simultaneously, the interaction of n-3 and n-6 PUFAs alongside their molecular pathways in cancer therapy is still contentious and complicated, and there is a need for more research [104]. Anticancer medications are further modified by conjugating them with fatty acid molecules (such as doxorubicin conjugates, paclitaxel conjugates, cytarabine conjugates, gemcitabine conjugates, and ciprofloxacin conjugates), which boosts the efficacy of therapeutic selectivity against various cancer cells with lower doses [103]. In advanced breast cancer, a combination of ω-3 PUFAs, doxorubicin, cyclophosphamide, and fluorouracil chemotherapy, as well as mastectomy, inhibits proliferation and angiogenesis by downregulating Ki-67 and vascular endothelial growth factors (VEGF) expression. In addition, vitamin D supplementation decreases inflammatory markers (IL-1b, IL-6, IL-8, TNF-α) and tumor markers in colorectal malignancies. In cancer patients receiving chemotherapy, supplementing with ω-3 fatty acids reduces cancer-related fatigue [20]. Simultaneously, fluorouracil conjugated with DHA has been shown to be more efficient in treating gastric cancer [110]. Additional research and clinical studies (phases I–III) are needed, however, to ensure and define the biochemical processes and pharmacokinetics of these novel conjugates.

Polyunsaturated aldehydes (PUAs) are oxylipins produced by a variety of marine and freshwater diatoms when subjected to various environmental stresses. The abundance of various microbial (bacterial, virus, and plankton) communities have been hypothesized to be influenced by PUAs [111]. After cell disruption, PUAs are produced by oxidative degradation of PUFAs [112]. The diatoms that produce PUAs are mainly Skeletonema costatum, Thalassiosira rotula, Skeletonema marinoi, Attheya longicornis, Chaetoceros socialis, Porosira glacialis, Chaetoceros furcellatus, and Pseudo-nitzschia delicatissima. When grown in Conway's medium, Daigo IMK medium, Guillard's F/2 medium, or versions of both media, these diatoms/microalgae exhibit anticancer properties [5]. PUAs have been shown to have anti-proliferative activity, reducing the sustainability of the human colon adenocarcinoma cell line (Caco-2) to 0% after 48 h of incubation at a concentration of 11-17 µg/mL. To validate the presence of apoptosis, the TUNEL assay was employed [113]. The cytotoxic potential of PUAs has also been established on cancer cells (lung (A549), colon (COLO 205), and adenocarcinoma cells, but not on healthy cells when incubated for 24 and 72 h [114].

On the other hand, few microalgae such as Chlorella sp., Chlamydomonas sp., Scenedesmus sp., Ankistrodesmus sp., Nannochloropsis limnetica, Stephanodiscus hantzschii, Gomphonema parvulum, Cyclotella meneghiniana, Cryptomonas sp., and Monoraphidium sp. have been evidenced for alternatively producing commercial sterols (β-sitosterol, stigmasterol, ergosterol, campesterol, and brassicasterol) [115, 116]. According to research, sterols have cytotoxic and anticancer properties. Furthermore, sterols suppress tumor growth, metastasis, and angiogenesis by inducing caspase-3-dependent apoptosis, Bax/Bcl2 increase, or blood cholesterol reduction, reducing the risk of cancer [20].

Other miscellaneous microalgal components

Vitamins, minerals, polyphenols, and Coenzyme Q, besides carotenoids, were demonstrated to possess strong anticancer properties [20]. Vitamin A obtained from various microalgae (Tetraselmis suecica, Dunaliella tertiolecta, Chlorella stigrnatophora, Skeletonema costatum, Isochrysis galbana, Aphanizomenon flos-aquae, Tetradesmus Obliquus, and Spirulina sp.) is composed of retinol, once in the body, it is metabolized into retinoic acid and retinoids [20, 117]. However, retinoic acid's activity is contradictory because it can activate the ERK pathway, which promotes angiogenesis and metastasis. In combination with other chemotherapeutic medicines and antioxidants, retinoic acid, on the other hand, prevents various cancer prognoses, enhancing the patient's survival rate [118]. Vitamin C is derived from various microalgae (Nannochloropsis oculata, Nannochloris atomus, Chaetoceros muelleri, Pavlova lutheri, Rhodomonas salina, Skeletonema costatum, etc.) and has been shown to have higher anticancer potential when administered intravenously rather than orally [119]. Cancer cells are sensitized and killed by vitamin C via a number of methods, including oxidative stress, immune cell stimulation, inflammation modulation, and signaling pathway interference [20]. Furthermore, vitamin C has been shown to cause protein modification and mitochondrial malfunction in cancer cells when it enters through sodium-dependent vit C transporter2 (SVCT2) and glucose transporters (GLUTs), respectively, boosting cancer cell mortality [120]. Furthermore, exposure to sunlight is the principal source of vitamin D, sometimes described as the “sunshine” vitamin. Microalgae, compared to terrestrial and aquatic plants and animals, have been found to synthesize more vitamin D when exposed to UVB. Several microalgae such as Nannochloropsis oceanica, Skeletonema costatum, Pavlova lutheri, Isochrysis galbana, and Tetraselmis suecica are excellent producers of vitamin D [20]. Although there is a lack of research and evidence on vitamin D from microalgae as an anticancer agent. However, fewer studies suggest that it has anticancer potential by interfering with gene expression and improving cancer patients' relapse-free survival [121, 122]. Among all other vitamins, marine microalgae (Skeletonema costatum, Pavlova lutheri, Isochrysis galbana, Chlorella stigmatophora, Spirulina sp., Tetraselmis suecica, and Dunaliella tertiolecta) is a good source of vitamin E. Supplementing with vitamin E (300–1000 mg/day) has been shown to reduce patient mortality [20, 123]. Vitamin E comes in eight different major isoforms (α, β, δ, γ-tocopherols and -tocotrienol). Vitamin E is frequently used to treat nephrotoxicity and ototoxicity brought on by the drug cisplatin [20]. Vitamin E (especially tocotrienol) has been found to have anticancer properties in addition to its neuroprotective ones, inhibiting cell proliferation, angiogenesis, and cell cycle arrest while simultaneously inducing autophagy, paratopsis, and apoptosis through various mechanisms involving the Bax/Bcl ratio, death receptor activation, and caspase-9 activations [20, 124]. There are two forms of vitamin K, sometimes known as “Koagulation vitamin”: vitamin K1 (phylloquinone) and vitamin K2 (menaquinone). Vitamin K and its derivatives have been shown to have anticancer properties against a variety of malignancies. Several microalgae, including Chlorella ellipsoidea, Tetraselmis suecica, Skeletonema costatum, Isochrysis galbana, and Pavlova lutheri are good sources of vitamin K [20]. Furthermore, it has been demonstrated that vitamin K activates p21 and CDK1 inhibitors through a number of methods, killing cancer cells, including upregulation (Fas/FasL, NF-kB, and p53) and downregulation (Bcl-2/Bcl-xl and Bax/Bak) of numerous factors, as well as caspase-3 activation pathways [125].

Marine microalgae (Tetraselmis chuii, Botryococcus braunii, Phaeodactylum tricornutum, Chlorella sp., and Spirulina sp.) are high in macrominerals and microminerals, both of which have been shown to have antioxidant properties, lowering cancer risk [20]. Further antioxidant multivitamin and mineral (AMM) supplementation reduces oxidative damage caused by chemotherapy and radiotherapy in cancer patients, restoring endogenous and exogenous antioxidants and trace elements [126].

Microalgae (Diacronema lutheri, Phaeodactylum tricornutum, Haematococcus pluvialis, Chlorella vulgaris, and Tetraselmis suecica) produce polyphenols and their derivatives (phenols, flavonoids, dihydrochalcones, and proanthocyanidins). Researchers have discovered that polyphenols have anticancer and antioxidant capabilities [20, 127]. Certain cancers are inhibited from proliferating by the antioxidant characteristics of polyphenols (phenols and flavonoids), which elevate radical scavenging capability [128]. By activating pro-apoptotic, anti-proliferative, and anti-metastatic pathways, polyphenols (genistein, quercetin, and ellagic acid) have been demonstrated to alter molecular targets, suggesting their anticancer potential [20].

Ubiquinone, often known as coenzyme Q (CoQ10), is a well-known inducer of mitochondrial oxidative phosphorylation and adenosine triphosphate (ATP) production [129]. Few microalgae (Porphyridium purpureum, Chlorella pyrenoidosa, and Isochrysis galbana) have been shown to produce more CoQ10, either naturally or when freeze-dried. CoQ10 and Alpha-Lipoic acid (ALA) combination therapy has been shown to reduce inflammation and cancer risk by considerably enhancing antioxidant activity. Nonetheless, a higher risk of cancer has also been connected to low ubiquinone levels [20] (Table 1).

Table 1 List of compounds with anticancer properties derived from various microalgae

Macroalgae anticancer potential

Marine macroalgae are a substantial source of bioactive substances including polysaccharides, lipids, and proteins (primary metabolites) as well as phenolic compounds, halogenated compounds, sterols, terpenes, and short peptides (secondary metabolites) [144]. Based on their morphology and pigmentation, macroalgae are divided into three groups: green (Chlorophyta), red (Rhodophyta), and brown (Phaeophyta) [21]. Several biological properties of macroalgae have been considered notably anti-diabetic, anti-inflammatory, anticancer, antimicrobial, antihypertensive, anti-viral, neuroprotective, and fat-lowering activities [144]. The biological properties of the macroalgae-derived bioactive compounds depend on their extraction process which is available in detail in [21].

Macroalgal polysaccharides

SPs are anionic polymers biosynthesized by macroalgae as an important component of their cell walls and are regarded to be vital for physiological adaptation to the high ionic strength of the marine environment. The SPs that are widely used as potential bioactive compounds include ulvans, galactans (agarans and carrageenans), and fucoidans from green, red, and brown macroalgae, respectively [145]. In terms of anticancer activity, SP with low molecular weight and high sulfate content is considered advantageous [1]. The sulfate groups are covalently bonded in varying quantities (0 to 41%) to the carbohydrate atoms via ether bonds [146].

Ulvan is a polysaccharide of green macroalgae derived from various genera of Ulva, Caulerpa, Monostroma, Codium, and others. The structure of ulvan consists of xylose, rhamnose, uronic acids (glucuronic and iduronic acid), sulfate groups, and trace amounts of mannose, and galactose. Both D-glucuronosyl-(1,4)-L-rhamnose 3-sulfate and L-iduronic acid-(1,4)-L-rhamnose 3-sulfate are repeating disaccharide units that constitute the compound; sulfate content (18.9%) with molecular weights ranging between 1.8 × 105–2 × 106 [147]. Ulvan's interaction with the Toll-like receptor (TLR)4 receptor leads to the P13K/Akt and NF-κB signaling pathways to be activated, which causes the expression of IL-8, TNF-, and CCL20 to be induced to prevent tumor growth [148]. In a study using the human hepatoma (HepG2) cell line, ulvan administration triggered apoptosis by activating the caspases-mediated mitochondrial signaling system that produced cytochrome c (Cyt c), activated caspase-3, − 9, and Bax-Bcl-2 ratio [149].

Carrageenan is mainly composed of a linear chain of alternating α-1,3- and β-1,4-glycosidic linkages connecting 3-linked β-D-galactopyranose units and 4-linked 3, 6-anhydro-α-galactopyranose [10, 150] and are extracted from Kappaphycus alvarezii, Eucheuma denticulatum, Chondrus crispus, Chondrus pinnulatus, Chondrus armatus, Chondrus yendoi [10, 146]. The molecular mass of carrageenan range from 500 to 1000 kDa. According to their sulfate content and position, carrageenans can be divided into six categories: kappa (κ-), mu (µ-), iota (ɩ-), beta (β-), lambda (λ-), theta (θ), and nu (ν-) carrageenan. Among them, the most important types are κ-, ι-, and λ- carrageenans with 20%, 33%, and 41% of sulfate content, respectively [146]. In the signaling pathway for Wnt/β-catenin, Wnt interacts with the Frizzleds (Fr) receptors and coreceptors like low-density lipoprotein receptors (LPR5/6) activating the dishevelled (Dvl) protein. It destabilizes the destructing complex (Wnt/Fr/LRP/Dvl/Axin), thus accumulating β-catenin. This pathway is considered crucial for the formation of cancer stem cells [146, 151]. The application of carrageenan on different cell lines has both the pro-tumor and anti-tumor activity of Wnts in line with the type of tumor and the Wnt ligand involved. Contradictory results on Wnt-cascade signaling have noted the tumor-suppressing efficacy against leukemias, neuroblastoma, thyroid cancer, melanoma, and ductal breast cancer as well as tumor-promoting activity against gastric, prostate, pancreatic, melanoma skin, and non-small cell pulmonary cancer [151].

Fucoidan obtained from brown macroalgae (Ascophyllum nodosum, Ecklonia cava, Undaria pinnatifida, Fucus vesiculosus, Sargasssum hemiphyllum) [152] have molecular mass categorized into three groups: high molecular mass (> 10,000 kDa), intermediate molecular mass (10–10,000 kDa), and low molecular mass (< 10 kDa). It is composed of a backbone of 3-linked α-l-fucopyranose units or alternating 3-linked α-l-fucopyranose and 4-linked α- l-fucopyranose units, along with traces of glucose, mannose, xylose, galactose, rhamnose [152, 153]. Oversulfation of fucoidans promotes their bioactivity and is found to be a strong inhibitor of angiogenesis. The interaction of fucoidan with several cancer-related pathways makes it a multipotent compound. It inhibits the phosphorylation of phosphatidylinositol-4,5-bisphosphate 3-kinase/protein kinase B (PI3K/AKT), mTOR while decreases the level of MMP-2 and MMP-9 on different cancer cell lines [153]. P13K/AKT signaling regulates the pro-apoptotic Bcl-2 subunits, which cause Cyt c to be released from the mitochondria and activate the caspase pathway [154]. In an in vitro investigation by Cho et al. [155], fucoidan has been shown to reduce NF-κB activity on bladder carcinoma cell lines. The same study revealed that fucoidan treatment induced the level of p21WAF1, a cell cycle inhibitor, through upregulation of Akt signaling pathway. Fucoidan has shown promising results against various carcinoma cell lines including acute myeloid leukemia (NB4, HL-60), colon, breast, lung, uterine, ovarian, endometrial, and colorectal cancer [153]. In the MAPK/ERK pathway, the protein level of the phosphorylated ERK1/2 is reduced by fucoidan for apoptotic induction. However, both the inhibitory and stimulatory expression of p38 MAPK has an anti-proliferative effect on colon, leukemia, and gastric cancer cells [153]. Fucoidan's anticancer effects are also linked to its capacity to obstruct a plethora of growth-related receptors, including the estrogen receptor, TGF-β, bone morphogenetic proteins (BMPs), and VEGFs (ER) [156,157,158,159].

Macroalgal peptides

Peptides with anticancer properties are low molecular weight cationic peptides. In comparison to conventional chemotherapy, anticancer peptides are known to efficiently inhibit tumor growth, migration, and angiogenesis. Several mechanisms of anticancer peptides are involved in inhibiting tumorogenic activities including cell membrane destruction, apoptosis, inhibition of tumor angiogenesis, and immune regulation [160]. A study of papain- and pepsin-digested hydrolysates obtained from Pyropia haitanensis showed anti-proliferative activity against breast (MCF-7), liver (HepG2), and lung (A549) carcinoma cell lines. They exhibited an IC50 value ranging between 59.09 to 272.67 μg/ml. In addition, a novel peptide (QTDDNHSNVLWAGFSR) was isolated with an inhibitory effect of 61.36% (at 500 μg/ml) on the HepG2 cancer cell line [161]. Another study by Fan X et al. [162] showed that polypeptides isolated from Porphyra haitanensis exhibited anti-proliferative activity against A549, HepG2, HT-29, MCF-7, SGC-7901 cancer cell lines with an IC50 value within 191.61 and 316.95 μg/ml. Furthermore, two novel peptides (VPGTPKNLDSPR and MPAPSCALPRSVVPPR) showed anti-proliferative activity against MCF-7 (IC50 = 200.97 μg/ml) and HepG2 (IC50 = 276.85 μg/ml). By halting the cell cycle at the G0/G1 phase and causing apoptotic cell death, polypeptides exerted an anticancer action on cancer cells. Undaria pinnatifida is a green macroalgae that is rich in proteins and has remarkable bioactive qualities; however, there is little evidence of its anticancer potential [163]. In an investigation by Rafiquzzaman [164], glycoprotein isolated from Undaria pinnatifida functions as a natural, bioavailable antioxidant with DNA-protective properties. Because molecular weight and structural properties govern the migration and penetration of peptides inside the body, the low molecular weight hydrolysates of protein and peptides of U. pinnatifida are likely to exhibit high radical scavenging action [163].

Macroalgal lipids

PUFAs, which contain ω-6 and ω-3 fatty acids, are crucial for maintaining various metabolic processes that lower the risk of heart disease, cancer, and inflammatory diseases. ω-6 fatty acids are the linoleic and arachidonic (AA) acids and ω-3 fatty acids are EPA and DHA [165]. The short-chain PUFAs are the ω-3 alpha-linolenic acid (ALA), and ω-6 linoleic acid (LA), while the long-chain PUFAs are the ω-3 EPA and DHA, and ω-6 AA [166]. Mammals lack the enzymes essential for the synthesis of PUFAs and hence have to be obtained through diet. The dietary ratio of ω-6:ω-3 has to be 2:1 in healthy individuals; however, excessive intake of ω-6 can lead to diseases like cancer [165, 167]. Epidemiological evidence on the association between PUFA and cancer indicates that ω-3 PUFA prevents cancer whereas ω-6 PUFA induces it [167]. Macroalgae possess an abundant amount of long-chain PUFAs acting as a good source with nutritional value. In comparison to green and red macroalgae, brown macroalgae contain the highest quantity of PUFAs [168, 169]. The potential sources of PUFAs from macroalgae are Gracilaria corticata, Gelidium sp., Pyropia sp., Undaria pinnatifida, Ulva, Gelidiella sp., Polysiphonia sp., Monostroma, Caulerpa, Rhodymenia sonderi, Acanthophora sp., Acrosiphonia, Bryopsis, Cryptonemia undulata, Halymenia sp., and Udotea [170]. PUFA from Adenocystis utricularis displayed growth inhibition of human breast tumor cells (MCF-7 and MDA-MB-231) between 61.04% and 69.78%. The cell viability for MCF-7 (68.7%) and MDA-MB-231 (89%) reduced on exposure to fatty acids from Adenocystis utricularis for 72 h. Furthermore, the 100 μg/ml concentration of fatty acids had > 50% anti-proliferative effect against breast tumor cell lines [171]. Fucus spiralis fatty acid-containing petroleum-ether fraction was cytotoxic to the HeLa cell line due to its anti-migratory, anti-angiogenic, and cell cycle-arresting effects. Its IC50 value was 43.74 μg/ml [172]. Sterols, belonging to a subset of steroids, are the amphipathic lipids with a hydroxyl group (C3 carbon atom) and a branching chain (C17 carbon atom). Many types of sterols such as cholesterol, clionasterol, isofucosterol, fucosterol, sargasterol, and others are found in macroalgae several biological properties [173]. According to Li et al. [174], saringosterol acetate from Sargassum fusiformis had anti-proliferative action on the MCF-7 cell line via inducing mitochondrial-mediated apoptosis (IC50 = 63.16 μg/ml). Commercially purchased fucosterol induced mitochondrial-mediated apoptosis, endoplasmic reticulum stress, and anti-angiogenic effects on human ovarian tumor cell lines (ES2 and OV90) with an IC50 value of 62.4 μM (ES2) and 51.4 μM (OV90) [175].

Macroalgal vitamins

Vitamins are necessary to keep the body's physiological and biochemical processes functioning properly. It has been reported that several macroalgae possess vitamins beneficial for preventing various diseases. Vitamin B12 is known to exist in the highest proportion in red algae (Porphyra sp.). Other species of macroalgae with vitamin B12 are Palmaria longat, Porphyra tenera, Enteromorpha, etc. All species of green, red, and brown macroalgae possess vitamin C and vitamin E (α-tocopherol) including Undaria pinnatifida and Laminaria digitata which have both the vitamins [176]. Vitamin A as already described above have anticancer property. By suppressing the expression of myosin light chain kinase via the MAPK pathway, Zuo et al. [177] found that All-trans retinoic acid (ATRA) has an antimigratory effect against human colorectal carcinoma cells (RKO). B vitamins (B1, B2, B3, B5, B6, B7, B9, B12) are important for generating cofactors required for important cellular and metabolic functions [178]. Vitamin B1 (thiamine) (2 μg/ml) inhibited the proliferation of the MCF-7 breast carcinoma cell line by 63% [179]. Vitamin B6 comprises pyridoxal, pyridoxine, pyridoxamine, along with their phosphorylated forms: pyridoxal-5’-phosphate, pyridoxine-5’-phosphate, pyridoxamine-5’-phosphate [178]. The strong anti-inhibitory activity was observed at a concentration of 20 μM pyridoxal against B16F10 murine melanoma cells [180]. In an in vivo experiment, supplementation of folate and vitamin B12 to azoxymethane-induced carcinogenic mice combats against the cytotoxicity and oxidative stress of azoxymethane [181]. Based on the administration route mentioned above, Vitamin C, often known as ascorbic acid, exhibits anticancer effects. Hepatocellular tumor cells (Hep3B) were treated with low-dose methotrexate and vitamin C in combination to induce H2O2 production and activate caspase-8/-9, hence promoting cell death [182]. The proliferation of the anaplastic thyroid carcinoma cell lines (8505C and C643) was inhibited successfully on treatment with vitamin C at a concentration of 1 mM through ferroptosis via GPX4/PTGS2 pathway [183]. Our skin on exposure to sunlight (UVB, 290–320 nm) produces vitamin D, a seco-steroidal prohormone. It goes through metabolic processes in the liver and kidney to yield calcitriol (biologically active metabolite). Apart from its role in bone metabolism, it is reported to function in cancer treatment and prevention [184]. In a recent study on breast cancer cell lines (MCF-7 and MDA-MB-231), the reduction in the cell viability was 72% (10 μM vitD) at 24 h for MCF-7. This was due to the imbalance in cellular iron homeostasis inducing oxidative stress contributing to cell death [185]. Also, a study by Casadei-Gardini et al. [186] evaluated patients with cholangiocarcinoma undergoing surgery for disease-free survival (DFS) and found that intake of vitamin D improves DFS. Vitamin E and Vitamin K have already been described above and show anticancer activity. α-tocopherol (Vitamin E) exhibits anti-tumor activity on squamous carcinoma cell (ORL-48) at IC50 value of 2.5 μg/ml through apoptotic cell death and sub-G0 phase cell cycle arrest [187]. A recent study on vitamin K2 depicted AMPK-dependent autophagic cell death in human bladder tumor cells (T24, EJ, and J82) on induction of PI3K/AKT/hypoxia-inducible factor-1α (HIF-1α)-mediated glycolysis [188]. Most of the vitamins have shown controversial results on cancer and further investigation needs to be performed for analyzing their exact roles.

Other miscellaneous macroalgal components

Carotenoids are a macroalgal pigment that includes fucoxanthin, β-carotene, astaxanthin, violaxanthin, capsanthin, siphonaxanthin, lutein, neoxanthin, and others [189]. Among them, the major carotenoids are the fucoxanthin widely distributed in brown algae (Undaria pinnatifida, Laminaria japonica, etc.). The structure of fucoxanthin possesses an allenic bond, a 5,6-monoepoxide, and an acetylated group [190]. With regard to anticancer activity, fucoxanthin, and fucoxanthinol (metabolite) induce apoptotic cell death, cell cycle arrest, antiproliferation, and anti-angiogenic effect [189]. Fucoxanthin exhibits its effect by downregulating MAPK, Bcl-2, MMP-9, and mRNA expression levels of CD44, CXCR4 and stimulation of poly-ADP-ribose polymerase (PARP), and caspase-3,-8,-9 [1]. An analysis by Wang et al. [191] showed that the human bladder cancer T24 cell line was inhibited by fucoxanthin at a concentration of 5 μM and 10 μM via G0/G1 cell cycle arrest through downregulation of CDK-2, CDK-4, cyclin D1, cyclin E, and upregulation of p21, CDK-inhibitory protein. Fucoxanthin was also responsible for the downregulation of the mortalin-p53 complex. On treating with Undaria pinnatifida-derived fucoxanthin, the growth of MDA-MB-31 (human breast cancer) and tumor-induced lymphangiogenesis were suppressed by reducing the concentrations of VEGF-C, phospho-AKT, VEGF receptor-3, phospho-P13K, NF-κB in human lymphatic endothelial cells. However, in in vivo MDA-MB-31 nude mouse model micro-lymphatic vascular density (micro-LVD) was reduced [192].

Polyphenols are produced by seaweeds to boost their antioxidant properties and act as radical scavengers. They produce polyphenolic compounds including phlorotannins, flavonoids, bromophenols, mycosporin-like amino acids, and phenolic terpenoids [1]. Phlorotannins are the major polyphenols unique to brown algae [20] such as kelps, rockweeds, Ecklonia cava, Laminaria japonica, and Sargassacean sp. and comprise a monomeric unit phloroglucinol (1,3,5-trihydroxybenzene) [1]. On the basis of the links between the monomeric units, phlorotannins are divided into four classes: phlorethols and fuhalols (ether links), fucols (phenyl links), fucophlorethols (ether and phenyl links), and eckols and carmalols (dibenzodioxin links) [193]. Zenthoefer et al. [194] produced an acetonic extract of Fucus vesiculosus (thallus) for inhibiting the viability of pancreatic cancer cells (Panc89 and PancTu1). The EC50 value for Panc89 was 71.47 μg/ml and PancTu1 was 76.96 μg/ml. Also, the inhibitory rate of Panc89 and PancTu1 was 80.3% and 82.6%, respectively. The application of phlorethols from Costaria costata showed an IC50 value of 92 μg/ml, 94 μg/ml, 96 μg/ml, and 102 μg/ml for HT-29, HCT-116, MCF-7, and SK-MEL-28, respectively [193]. Eckol prohibited Reg3A-induced SW1990 cells from multiplying (pancreatic human cells) [195], while dieckol had anti-proliferative and anti-migratory impact on non-small-cell pulmonary cancer by regulating PI3K/AKT pathway [196].

Similar to microalgae, many macroalgae-derived components have also been investigated using in silico methods, albeit with very scanty data. Although caulerpin from Caulerpa racemosa was molecularly docked, it showed that it was an efficient ligand but had a reduced total binding energy when considering whether it could be used as a therapeutic molecule [197]. Simultaneously, another study of the anticancer activity of metabolites from Caulerpa sp. has been identified as an effective ligand against glucose 6-phosphate dehydrogenase (G6PDH) and 6-phosphogluconate dehydrogenase (6PGD) for targeting the pentose phosphate pathway in colorectal cancer treatment [198] (Fig. 4, Table 2).

Fig. 4
figure 4

Different algal growth and extraction methodologies for bioactive components, as well as their potential applicability to various cancers and cell lines

Table 2 List of compounds with anticancer properties derived from various macroalgae

Can multi-drug-resistant gastric cancer be effectively treated with algal metabolites?

Antimicrobial resistance has rendered many conventional antibiotics ineffective in a number of people with H. pylori infection [245]. Antimicrobial resistance to metronidazole, tetracycline, quinolones, clarithromycin, and rifabutin has emerged as a result of gene modulations or mutations, according to molecular investigations [246, 247]. H. pylori has been classified as priority 2 by the World Health Organization (WHO) due to antibiotic resistance, despite the fact that it is found colonizing in 50% of the human stomach [248]. As a result, a method to lower cancer's morbidity and mortality has been found in algae with special metabolites that can stop the progression of an H. pylori infection into gastric cancer in the era of multi-drug resistance [247]. However, there is a scarcity of data on the antibacterial activity of microalgal bioactive components (mostly carotenoids) against H. pylori. In macroalgae, fucoidan is widely used for its anti-H. pylori activity. Infection with H. pylori raises the risk of gastric and colon cancer. Gastric cancer (GC) arises from a complicated, multi-step process that starts with normal mucosa and progresses to non-atrophic gastritis. In a cascade, the progression of superficial gastritis to atrophic gastritis leads to the production of metaplasia, dysplasia, and intestinal-type cancer [249]. The expression of several outer membrane proteins/adhesins by the bacteria (BabA, SabA, AlpA/B, HopZ, and OipA) aids in the establishment of an intimate relationship with the gastric mucosa cells of the host. Alongside adhesins, numerous virulence factors (vacuolating cytotoxin A (VacA), cytotoxin-associated gene A (CagA), and urease) produced by H. pylori are integrated into host cells via the type 4 secretory system (T4SS) within cag pathogenicity island (cag PAI) for initiating pathogenesis [249, 250]. Instead of using bactericidal drugs, antiadhesives can prevent the pathogenesis that results from H. pylori adhering to stomach mucus via lectin-like molecules. In microalgae, Chlorella sp. and Spirulina sp. can be exploited cost-effectively for their polysaccharides have been evidenced to favor anti-adhesive action in the gastric environment reducing > 90% H. pylori load observed in BALB/c mice models [251]. These algal polysaccharides have not been shown to impede in vitro bacterial growth, but they can prevent H. pylori infection/reinfection due to their anti-adhesive characteristics, making them a safe and cost-effective option [252]. In silico analysis of algal peptide interaction with H. pylori suggested peptides from green microalgae Tetradesmus sp. have been found active inhibitors against three virulent factors (Cag A, VacA, and Htr A) of H. pylori [253]. Additional probiotic therapy utilizing the carrageenan-encapsulated Lactobacillus fermentum UCO-979C strain has demonstrated anti-H. pylori efficacy under fasting conditions (pH 3.0) [248]. Although about 2% to 3% of people develop gastric cancer, this is primarily owing to the infection's persistence. Biofilm development, in addition to mutation, enhances bacterial antibiotic resistance. Considering algae's powerful predator defense mechanisms, bioactive substances identified in them are among the most promising sources. Algal extracts have also been shown to degrade the biofilm polymer matrix, resulting in an anti-film effect that, when coupled with antibiotics, prevents bacterial colonization from progressing and developing into gastric cancer [254]. In addition to genetic and environmental factors, modifications to the stomach adaptive system result in endoplasmic reticulum (ER) stress, which activates the unfolded protein response and causes precancerous lesions to form at the precancerous stage [255]. Because pro-apoptotic proteins (Bim and Bax) are present while VacA interference is present, ER stress causes CHOP transcription, which speeds up apoptosis. Activation of NF-κB, on the other hand, inhibits apoptosis via A20 deubiquitinylase activity, resulting in infection-mediated GC that persists [249, 256]. Furthermore, the increase in autophagosomes as a result of autophagy activation/inhibition enhances invasion and metastasis by causing ROS-mediated oxidative stress in the early stages of cancer. The autophagy generated by VacA exposure has been shown to be contradictory; nevertheless, autophagy inhibition via the CagA protein, which is involved in the c-Met-PI3K/Akt-mTOR signaling pathway, was identified as a confirmation of GC development [257]. Bromophenol and its derivatives (BOS-93 and BOS-102) from marine algae have been shown to inhibit and downregulate the PI3K/Akt/mTOR and MAPK signaling pathways, as well as Bcl-2, MMP, and Cyt-c expression, while upregulating ERK, Bax, Atg14, beclin-1, and phosphorylated p38 expression thereby stimulating apoptosis and preventing carcinogenesis [258]. Inflammation has long been a key factor in the development of cancer. CagA protein and peptidoglycan, as well as activating TLRs, NF-κB, TNF-α, STAT-3, IL-1β, IL-2, IL-4, IL-8, IL-10, IL-12, TNF, IFN, epidermal growth factor response (EFGR), and COX-2/prostaglandin E2 (PGE2) pathways, cause gastric inflammation [44, 159, 249, 259]. Apart from TLR/MYD88 adapter signaling, COX-2 stimulation, as well as Wnt signaling activation and β-catenin accumulation, promotes carcinogenesis. Furthermore, activating PI3K and MAPK signaling via PGE2 signaling causes the growth of CD33 + CD44 + cancer stem cells [159]. Astaxanthin and fucoxanthin from various microalgae and macroalgae have been found to inhibit tumor/cancer progression in vitro experiments by regulating and preventing cell cycle arrest, p27 expression, ERKs expression, NF-κB expression, MMP-2/9 expression, and apoptosis induction [247]. Moreover, the ability of human carcinoma MKN45 gastric cells to invade was shown to be inhibited by a new polysaccharide derived from brown algae (Sargassum sp.). In cancer cells, this polysaccharide caused JNK phosphorylation, p53, Caspase-3 and 9, and ROS, halting the cell cycle (G2/M phase) and triggering apoptosis through the ROS/JNK signaling pathway [260]. This polysaccharide, on the other hand, had no effect on migration and no effect on p38 MAPK signaling or downstream MMP-9/2 [261]. Moreover, porphyran from Porphyra sp. has anticancer characteristics that inhibit in vitro adenocarcinoma cell lines (AGS) by triggering apoptosis via the mitochondrial pathway [262]. Apart from dislodging H. pylori from human AGS cells, fucoidan fraction (Fucus B) has been shown to cause dose-dependent cytotoxicity in AGS cancer cells, which was verified using a lactate dehydrogenase assay. Additionally, 6 gms of fucoidan taken regularly seems to be non-toxic and has the potential for use in treating H. pylori infection and GC formation [263]. Inflammation-mediated GC is accelerated by other inflammatory cytokines (CXCL1, CXCL2, CXCL5, CCL3, CCL4, and TLR2) under COX2/PGE2 signaling [264]. At 100 mg/kg/day, a carotenoid-rich acetone extract of Chlorococcum sp. decreased inflammation, lowered IFN-γ and IL-4 levels in splenocytes, and lowered bacterial density in infected BALB/c mice [265]. Furthermore, brown algae sulfated polysaccharides (fucoidan) have been shown to decrease the expression of IL-1β, IL-6, iNOS, PGE2, NO, and TNF-α. In addition, the anti-inflammatory fucoidan has suppressed complement-related inflammation in the stomach wall [254]. As indicated before, it has been demonstrated that phycobiliproteins from cyanobacteria reduce the production of pro-inflammatory cytokines including NO and COX-2 [266]. Furthermore, oxidative stress caused by H. pylori in the gastric cells contributes to GC. In addition to CagA's fatal impact, pro-oxidant activities like host spermine oxidase, NADPH oxidase, or mitochondria-mediated ROS production reduce antioxidant or glutathione activity in H. pylori-infected patients. Furthermore, nitric oxide (NO) produced in macrophages, gastric cells, and lymphocytes causes DNA adducts and nitrotyrosine, instigating DNA and protein damage [249]. It has been demonstrated that the antioxidant properties of polysaccharides, carotenoids, lipids, peptides, and pigments of micro- and macroalgae can repair ROS-induced damage in cancer cells, as previously mentioned in this review. Consumption of antioxidant or selenium supplements at the same time not only replenishes SOD, catalase, and glutathione levels but also regulates positive gene expression on intracellular and intercellular signaling, preventing deadly damage in GC [267]. Nonetheless, using algae in combination with conventional antibiotics to treat antimicrobial-resistant H. pylori infection and prevent GC may be beneficial. Because these bacteria have developed resistance and there is no cure or prophylactic available, more progress in developing vaccines using biomedical approaches is required.

Biomedical approaches

The development of biomaterials has become one of the foremost significant fields of research in contemporary science, with tremendous promise for biological applications [268]. Furthermore, due to their non-toxic, biodegradable, and biocompatible properties, the biological exploration of natural materials has risen [152]. Despite the fact that various biomaterials have been employed as biological agents to combat drug resistance. Algae, a ubiquitous photosynthetic organism, has long been regarded as interesting naturally active biomaterials with a range of applications including drug administration, bioengineering, wound repair, bioanalysis, and hypoxia-mediated tumor therapy [268]. Microalgae have demonstrated strong targeted drug delivery capabilities in both in vitro and in vivo investigations, with an emphasis on anticancer effects, by loading drug molecules through their active surfaces. Microalgae (Spirulina platensis)-based oral medication delivery systems containing (SP@Curcumin) have been shown to be easily trapped and adhered to intestinal villi and wall, in contrast to conventional oral drug delivery problems. In colon cancer and colitis, the SP@Curcumin has been studied for its ability to operate as a radioprotector by scavenging ROS generated by healthy tissues after X-ray exposure, as well as lowering pro-inflammatory cytokine production and increasing drug bioavailability [269]. Additionally, it has been demonstrated that C. reinhardtii that has been engineered to contain chitosan-coated iron oxide nanoparticles (CSIONPs) coupled to the chemotherapeutic medication doxorubicin (DOX) enhances the drug uptake in SK-BR3 cancer cells [270]. Simultaneously, the development of Spirulina sp. as a biotemplate-based magnetic microrobot ((Pd@Au)/Fe3O4@Sp.-DOX) exhibited excellent synergistic chemo-photothermal therapeutics for both 769P and EC109 cancer cells [271]. When combined with cell-targeting antibodies and chemotherapeutic chemical molecules (camptothecin and 7-ethyl-10-hydroxy-camptothecin), genetically modified biosilica frustules from the altered diatom Thalassiosira pseudonana specifically targeted and killed in vivo neuroblastoma cells in mice models (SH-SY5Y) [272]. Furthermore, it has been demonstrated that lung-targeted administration of positively charged DOX molecules using negatively charged S. platensis kills 4T1 and CT26 tumor cells [273]. Furthermore, tumor hypoxia is caused by unregulated cell proliferation, altered metabolism, and aberrant tumor blood cells, all of which result in inadequate oxygen and nutrient transfer. Hypoxia causes cell cycle arrest, suppresses apoptosis and cell death, modifies the activity of the p53 gene and the mitochondria, expresses the drug efflux pump (P-gp), and reduces oxygenation in the case of chemotherapy cytotoxicity [274]. Since the ROS generated by photodynamic treatment (PDT) and radiation therapy (RT) are converted from oxygen, these two therapeutic modalities rely largely on oxygen. Living algae are therefore anticipated to boost cellular oxygen levels through photosynthesis, acting as a source for the production of ROS and thereby boosting the impact of PDT/RT [268]. Spirulina sp. and Chlorella sp. are microalgae that help to reduce tumor hypoxia by acting as oxygenators in in vivo tumor synergistic therapies (PDT, RT, or PDT/RT) [269]. A cancer-targeted theranostic approach involving biohybrid microswimmers based on engineered S. platensis has been shown to increase oxygen generation in a 4T1 bearing mouse model, as well as innate chlorophyll and magnetic resonance derived fluorescence and photoacoustic imaging for monitoring effective tumor therapy procedures and modifying tumor microenvironment hypoxia [275]. Additionally, an autotrophic light-triggered green affording oxygen engine (ALGAE) made of calcium alginate and C. pyrenoidosa was implanted into 4T1 tumor-bearing mouse tumors. This engine was triggered three times to induce hypoxia-resistant PDT and successfully limit tumor growth and metastasis [269]. Mice with 4T1 tumors were given an intravenous injection of a modified C. vulgaris-based biohybrid Algae@SiO2 system, which inhibited tumor growth. Moreover, ROS generated from Algae@SiO2-derived chlorophyll induces cytotoxicity to cancer cells throughout the photodynamic therapy [276]. In addition, in breast tumors (4T1) and ovarian tumors (SKOV3) mice models, the red blood cell membrane (RBCM) was included in the surface modification of C. vulgaris dramatically lowering tumor development, hypoxia-dependent radioresistance, angiogenesis, and proliferation, triggering death. Downregulation of HIF1 and VEGF, as well as a decrease in Ki67 and CD31 expression, raises cleaved caspase-3, which aids in apoptosis induction and could lead to the development of algae-mediated hypoxia-related tumor therapy in the future [277].

Macroalgae-derived bioactive compounds have a vast array of biomedical applications. Among them, the priority focus is on polysaccharides due to their maximum content in green, brown, and red algae [10]. They are considered advantageous therapeutically as they are biocompatible, non-toxic, biologically tunable, and biodegradable [152]. The presence of biologically active metabolites in seaweeds has gained attention as food supplements in East Asia for centuries. Polysaccharides from macroalgae act as dietary fibers stimulating the production and thickness of intestinal mucus, thus protecting against carcinogenic compounds. Consumption of recommended intake of seaweeds can prevent colon, rectal, stomach, and breast cancer [2, 278,279,280]. Dietary intake of Laminaria sp., Saccharina sp., Undaria pinnatifida, and Porphyra/Pyropia sp. are all known to reduce the risk of breast cancer [278, 281]. Through the activation of NK cells, macrophages, and T cells, seaweeds have immunomodulating properties that enable them to recognize cancer antigens and harm target cells while also enhancing the immune system to prevent the growth of cancer [2]. Laminaria digitata-derived Laminarin stimulated the maturation of dendritic cells and production of Tc cells, IFN-γ, and TNF-α at a concentration of 25 mg/kg to inhibit B16-ovalbumin melanoma tumor growth and metastasis in a mouse model [282]. According to research by Sun et al. [283], fucoidan from Fucus vesiculosus suppressed MHCC-97H (human hepatoma cell line) motility by downregulating CCL22 in M2 macrophages, hence reducing NF-B-dependent transcription. Specifically, the seaweed polysaccharides being hydrophilic are suitable to act as drug delivery agents for hydrophobic anticancer drugs. Nevertheless, these polysaccharides have the ability to reduce the side effects of drugs, prevent the dispersion of chemotherapeutic agents throughout the body as well as optimize the release of anticancer compounds [284]. Oligocarrageenan obtained from the K. alvarezii κ-carrageenan was modified with polycaprolactone (PCL) chains to form PCL-grafted oligocarrageenan nano-micelles (187 nm) that encapsulated curcumin (hydrophobic drug). This enhanced the anti-inflammatory activity in TNF-triggered inflammatory trials [285]. Hydrophobic anticancer compound docetaxel is encapsulated with fucoidan-poly(lactic-co-glycolic acid) nanocarrier (FPN-DTX) [284]. According to Kahya et al. [286], the controlled release of methotrexate is achieved by crosslinking sodium alginate (NaAlg)/sodium carboxymethyl cellulose (NaCMC) composite hydrogel beads with a barium chloride solution. A superabsorbent hydrogel was prepared using carboxymethylagarose (CMA) and polyacrylamide (PAm) while extracting agarose from Gracilaria dura and forming CMA-Ag-PAm. The hydrogel showed extensive pH-responsive behavior causing an enhanced release of DOX with a decline in pH from 7.4 to 5.0. This DOX-loaded hydrogel attributes to cytotoxicity against A549 and Hep-G2 cell lines [287]. Nanoparticles derived from macroalgae show a wide spectrum of anticancer properties. In order to create biocompatible silver nanoparticles with an IC50 value of 95.35 g/ml against Ehlrich Ascites Carcinoma cell lines, the aqueous extract of Enteromorpha compressa is used [205]. Fabrication of methyl gallate encapsulated zeolitic imidazole framework (MG@ZIF-L) prepared from the extracts of Gracilaria debilis showed good biocompatibility, high loading capacity, and rapid release of drugs in the tumor microenvironment. The nanocomposite is cytotoxic to A549 (lung cancer cell line) because of enhanced ROS generation, which causes mitochondrial damage and encourages apoptosis. The cytotoxicity of MG@ZIF-L was tested in an in vivo zebrafish embryo model system and found to be non-toxic [288]. The harmful effects of radiotherapy/chemotherapy can be overcome by using protective agents that are derived from macroalgae. In this area, the most exploited compound of seaweeds is phlorotannins. Phlorotannins extracted from Ecklonia cava include phloroglucinol, dieckol, eckol, and triphlorethol A that show radioprotection activity against radiation-induced damage and oxidative stress through inhibition of apoptosis [289,290,291]. Similarly, methanolic extracts of Polyopes lancifolia have been found to contain higher amounts of SOD and catalase (cytoprotective enzymes) that demonstrate radioprotective activity via antioxidant processes [292].

Conclusions

The abundance of bioactive components found in algae has aroused the interest of many experts, who have proposed potential applications in the industrial and medical sectors. Over the past few decades, numerous in vitro and in vivo investigations have demonstrated that algae offer a wide range of applications in cancer therapy. By activating either the caspase-dependent or caspase-independent apoptosis pathway, which is followed by the upregulation of various tumor suppressor factors and the downregulation of particular cancer genes, markers, and signaling pathways, it has been shown that these algal-derived components are effective at inducing cytotoxicity and cellular death. Furthermore, these ingredients' anti-adhesive, immunostimulating, and anti-inflammatory properties boost the effectiveness of the anticancer potential that has been shown to be effective in the treatment of H. pylori-infected stomach cancer. Algal metabolites therefore can be used to protect humans from various cancers, in addition to their biomedical applications, which are still being studied. Furthermore, based on their molecular weight and viscosity, algal metabolites have been used in a number of other nutraceuticals, proving their plasticity. In the realm of innovative drug development, which replaces manufactured pharmaceuticals, various preclinical investigations using these bioactive components have also been carried out recently. Microalgae have, however, been used much less frequently as anticancer drugs than macroalgal metabolites. Advanced extraction procedures, as well as ideal growth factors and genetic engineering, must be considered for improved metabolite production. Nonetheless, to solve the conundrum, thorough clinical trials and standard dosage recommendations must be devised, allowing the potentiality of these bioactive components to be assessed.

Availability of data and materials

This is a comprehensive review article, and the information has been provided from the literature mentioned in the references. The data that support the findings of this study are available from the corresponding author upon reasonable request.

Abbreviations

CDK:

Cyclin-dependent kinase

CDC:

Cell division cycle

MAP/RAS/RTK:

Mitogen-activated protein/rat sarcoma virus/receptor tyrosine kinase

PI3K:

Phosphoinositide 3-kinase

TGFβ:

Transforming growth factor-beta

NF-κB:

Nuclear factor-kappa B

JNK:

Jun N-terminal kinases

Ras-ERK:

Ras-extracellular signal-regulated kinase

STAT:

Signal transducer and activator of transcription

TAM:

Tumor-associated macrophages

TAN:

Tumor-associated neutrophils

TNF:

Tumor necrosis factor

IL:

Interleukins

GMCSF:

Granulocyte–macrophage colony-stimulating factor

GCSF:

Granulocyte colony-stimulating factor

ROS:

Reactive oxygen species

SIRT-1:

Sirtuin-1

COX:

Cyclooxygenase

HES:

Hairy and enhancer of split

PC:

Prostate cancer

SAPK:

Stress-activated protein kinases

mTOR:

Mechanistic target of rapamycin

GADD:

Growth arrest and DNA damage

FGFR:

Fibroblast growth factor receptor

EGR:

Early growth response

FGF:

Fibroblast growth factor

EPS:

Exopolysaccharides

MAA:

Mycosporine-like amino acids

ω:

Omega

CARD:

Caspase recruitment domain

VEGF:

Vascular endothelial growth factors

SVCT2:

Sodium-dependent vit C transporter2

GLUTs:

Glucose transporters

AMM:

Antioxidant multivitamin and mineral

ALA:

Alpha-lipoic acid

κ:

Kappa

µ:

Mu

ɩ:

Iota

β:

Beta

λ:

Lambda

θ:

Theta

ν:

Nu

BMPs:

Bone morphogenetic proteins

ATRA:

All-trans retinoic acid

PARP:

Poly-ADP-ribose polymerase

micro-LVD:

Micro-lymphatic vascular density

VacA:

Vacuolating cytotoxin A

CagA:

Cytotoxin-associated gene A

cag PAI:

Cag pathogenicity island

ER:

Endoplasmic reticulum

EFGR:

Epidermal growth factor response

CSIONPs:

Chitosan-coated iron oxide nanoparticles

ALGAE:

Autotrophic light-triggered green affording oxygen engine

RBCM:

Red blood cell membrane

MG@ZIF-L:

Methyl gallate encapsulated zeolitic imidazole framework

References

  1. Lee H, Selvaraj B, Lee JW (2021) Anticancer effects of seaweed-derived bioactive compounds. Appl Sci 11:11261

    Article  CAS  Google Scholar 

  2. Cotas J, Pacheco D, Gonçalves AMM, Silva P, Carvalho LG, Pereira L (2021) Seaweeds’ nutraceutical and biomedical potential in cancer therapy: a concise review. J Cancer Metastasis Treat 7:13

    CAS  Google Scholar 

  3. Sung H, Ferlay J, Siegel RL, Laversanne M, Soerjomataram I, Jemal A, Bray F (2021) Global cancer statistics 2020: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J Clin 71:209–249

    Article  PubMed  Google Scholar 

  4. Hanahan D, Weinberg RA (2011) Hallmarks of cancer: the next generation. Cell 144:646–674

    Article  CAS  PubMed  Google Scholar 

  5. Martínez Andrade KA, Lauritano C, Romano G, Ianora A (2018) Marine microalgae with anti-cancer properties. Mar Drugs 16:165

    Article  PubMed  PubMed Central  Google Scholar 

  6. Moghadamtousi SZ, Karimian H, Khanabdali R, Razavi M, Firoozinia M, Zandi K, Kadir HA (2014) Anticancer and antitumor potential of fucoidan and fucoxanthin, two main metabolites isolated from brown algae. Sci World J 2014:768323

    Google Scholar 

  7. Banerjee P, Erehman J, Gohlke BO, Wilhelm T, Preissner R, Dunkel M (2015) Super Natural II-a database of natural products. Nucleic Acids Res 43:D935–D939

    Article  CAS  PubMed  Google Scholar 

  8. Boopathy NS, Kathiresan K (2010) Anticancer drugs from marine flora: an overview. J Oncol 2010:214186

    Google Scholar 

  9. Blunt JW, Copp BR, Keyzers RA, Munro MHG, Prinsep MR (2015) Marine natural products. Nat Prod Rep 32:116–211

    Article  CAS  PubMed  Google Scholar 

  10. Pacheco-Quito E-M, Ruiz-Caro R, Veiga M-D (2020) Carrageenan: drug delivery systems and other biomedical applications. Mar Drugs 18:583

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Alves C, Silva J, Pinteus S, Gaspar H, Alpoim MC, Botana LM, Pedrosa R (2018) From marine origin to therapeutics: the antitumor potential of marine algae-derived compounds. Front Pharmacol 9:777

    Article  PubMed  PubMed Central  Google Scholar 

  12. Ouyang Y, Qiu Y, Liu Y, Zhu R, Chen Y, El-Seedi HR, Chen X, Zhao C (2021) Cancer-fighting potentials of algal polysaccharides as nutraceuticals. Food Res Int 147:110522

    Article  CAS  PubMed  Google Scholar 

  13. Saadaoui I, Rasheed R, Abdulrahman N, Bounnit T, Cherif M, Al Jabri H, Mraiche F (2020) Algae-derived bioactive compounds with anti-lung cancer potential. Mar Drugs 18:197

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. FAO (2021) Seaweeds and microalgae: an overview for unlocking their potential in global aquaculture development. https://www.fao.org/3/cb5670en/cb5670en.pdf Accessed on 5 April 2023

  15. Ramadan KMA, El-Beltagi HS, Shanab SMM, El-fayoumy EA, Shalaby EA, Bendary ESA (2021) Potential antioxidant and anticancer activities of secondary metabolites of Nostoc linckia cultivated under Zn and Cu stress conditions. Processes 9:1972

    Article  CAS  Google Scholar 

  16. Liu L, Heinrich M, Myers S, Dworjanyn SA (2012) Towards a better understanding of medicinal uses of the brown seaweed Sargassum in Traditional Chinese Medicine: a phytochemical and pharmacological review. J Ethnopharmacol 142:591–619

    Article  PubMed  Google Scholar 

  17. Abdelwahab R (2017) Therapeutic and pharmaceutical application of seaweeds. In: Nabti E (ed) Biotechnological applications of seaweeds. Nova, New York

  18. Hassanpour SH, Dehghani M (2017) Review of cancer from perspective of cancer. J Cancer Res Pract 4:127–129

    Article  Google Scholar 

  19. Abd El-Hack ME, Abdelnour S, Alagawany M, Abdo M, Sakr MA, Khafaga AF, Mahgoub SA, Elnesr SS, Gebriel MG (2019) Microalgae in modern cancer therapy: current knowledge. Biomed Pharmacother 111:42–50

    Article  CAS  PubMed  Google Scholar 

  20. Ferdous UT, Yusof ZNB (2021) Medicinal prospects of antioxidants from algal sources in cancer therapy. Front Pharmacol 12:593116

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Trivedi N, Mondal AS, Sharma R, Mone D (2022) Marine macroalgal biorefinery: Recent developments and future perspectives. In: Bhatia SK, Mehariya S, Karthikeyan OP (eds) Algal biorefineries and the circular bioeconomy. CRC Press, Boca Raton, pp 1–36

    Google Scholar 

  22. Novoveská L, Ross ME, Stanley MS, Pradelles R, Wasiolek V, Sassi J-F (2019) Microalgal carotenoids: a review of production, current markets, regulations, and future direction. Mar Drugs 17(11):640

    Article  PubMed  PubMed Central  Google Scholar 

  23. Sun X-M, Ren L-J, Zhao Q-Y, Ji X-J, Huang H (2018) Microalgae for the production of lipid and carotenoids: a review with focus on stress regulation and adaptation. Biotechnol Biofuels 11:272

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Wang Y, Tibbetts SM, McGinn PJ (2021) Microalgae as sources of high-quality protein for human food and protein supplements. Foods 10:3002

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Cheng D, Li D, Yuan Y, Zhou L, Li X, Wu T, Wang L, Zhao Q, Wei W, Sun Y (2017) Improving carbohydrate and starch accumulation in Chlorella sp. AE10 by a novel two-stage process with cell dilution. Biotechnol Biofuels 10:75

    Article  PubMed  PubMed Central  Google Scholar 

  26. Obrien R, Hayes M, Sheldrake G, Tiwari B, Walsh P (2022) Macroalgal proteins: a review. Foods 11:571

    Article  Google Scholar 

  27. Biris-Dorhoi E-S, Michiu D, Pop CR, Rotar AM, Tofana M, Pop OL, Socaci SA, Farcas AC (2020) Macroalgae—a sustainable source of chemical compounds with biological activities. Nutrients 12:3085

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Pepper JW, Scott Findlay C, Kassen R, Spencer SL, Maley CC (2009) Cancer research meets evolutionary biology. Evol Appl 2:62–70

    Article  PubMed  PubMed Central  Google Scholar 

  29. Peters JM, Gonzalez FJ (2018) The evolution of carcinogenesis. Toxicol Sci 165:272–276

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Bertram JS (2000) The molecular biology of cancer. Mol Aspects Med 21:167–223

    Article  CAS  PubMed  Google Scholar 

  31. Vendramin R, Litchfield K, Swanton C (2021) Cancer evolution: Darwin and beyond. EMBO J 40:e108389

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Patterson AD, Gonzalez FJ, Perdew GH, Peters JM (2018) Molecular regulation of carcinogenesis: Friend and Foe. Toxicol Sci 165:277–283

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Wu S, Zhu W, Thompson P, Hannun YA (2018) Evaluating intrinsic and non-intrinsic cancer risk factors. Nat Commun 9:3490

    Article  PubMed  PubMed Central  Google Scholar 

  34. Yasunaga JI, Matsuoka M (2018) Oncogenic spiral by infectious pathogens: Cooperation of multiple factors in cancer development. Cancer Sci 109:24–32

    Article  CAS  PubMed  Google Scholar 

  35. Azevedo MM, Pina-Vaz C, Baltazar F (2020) Microbes and cancer: friends or faux? Int J Mol Sci 21:3115

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Sever R, Brugge JS (2015) Signal transduction in cancer. Cold Spring Harb Perspect Med 5:a006098

    Article  PubMed  PubMed Central  Google Scholar 

  37. Morris MA, Dawson CW, Young LS (2009) Role of the Epstein-Barr virus-encoded latent membrane protein-1, LMP1, in the pathogenesis of nasopharyngeal carcinoma. Future Oncol 5:811–825

    Article  CAS  PubMed  Google Scholar 

  38. Sanchez-Vega F, Mina M, Armenia J, Chatila WK, Luna A, La KC, Dimitriadoy S et al (2018) Oncogenic signaling pathways in the Cancer Genome Atlas. Cell 173:321–337

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Derakhshani A, Rostami Z, Taefehshokr S, Safarpour H, Astamal RV, Taefehshokr N, Alizadeh N, et al (2020) Oncogenic signaling pathways in cancer: An overview. Preprints. https://www.preprints.org/manuscript/202003.0201/v1

  40. Wu JS, Jiang J, Chen BJ, Wang K, Tang YL, Liang XH (2021) Plasticity of cancer cell invasion: Patterns and mechanisms. Transl Oncol 14:100899

    Article  CAS  PubMed  Google Scholar 

  41. Zou S, Tong Q, Liu B, Huang W, Tian Y, Fu X (2020) Targeting STAT3 in cancer immunotherapy. Mol Cancer 19:145

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Grivennikov SI, Greten FR, Karin M (2010) Immunity, inflammation, and cancer. Cell 140:883–899

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Ozel I, Duerig I, Domnich M, Lang S, Pylaeva E, Jablonska J (2022) The good, the bad, and the ugly: neutrophils, angiogenesis, and cancer. Cancers (Basel) 14:536

    Article  CAS  PubMed  Google Scholar 

  44. Munn LL (2016) Cancer and inflammation. WIREs Syst Biol Med. https://doi.org/10.1002/wsbm.1370

    Article  Google Scholar 

  45. Skjånes K, Aesoy R, Herfindal L, Skomedal H (2021) Bioactive peptides from microalgae: focus on anti-cancer and immunomodulating activity. Physiol Plant 173:612–623

    Article  PubMed  Google Scholar 

  46. Jaillon S, Ponzetta A, Di Mitri D, Santoni A, Bonecchi R, Mantovani A (2020) Neutrophil diversity and plasticity in tumor progression and therapy. Nat Rev Cancer 20:485–503

    Article  CAS  PubMed  Google Scholar 

  47. Somarelli JA (2021) The hallmarks of cancer as ecologically driven phenotypes. Front Ecol Evol 9:661583

    Article  PubMed  PubMed Central  Google Scholar 

  48. Basu AK (2018) DNA damage, mutagenesis and cancer. Int J Mol Sci 19:970

    Article  PubMed  PubMed Central  Google Scholar 

  49. Vu CHT, Lee HG, Chang YK (2018) Axenic cultures for microalgal biotechnology: establishment, assessment, maintenance, and applications. Biotechnol Adv 36:380–396

    Article  CAS  PubMed  Google Scholar 

  50. Khan MI, Shin JH, Kim JD (2018) The promising future of microalgae: current status, challenges, and optimization of a sustainable and renewable industry for biofuels, feed, and other products. Microb Cell Fact 17:36

    Article  PubMed  PubMed Central  Google Scholar 

  51. Mondal A, Bose S, Banerjee S, Patra JK, Malik J, Mandal SK, Kilpatrick KL, Das G, Kerry RG, Fimognari C, Bishayee A (2020) Marine cyanobacteria and microalgae metabolites—a rich source of potential anticancer drugs. Mar Drugs 18:476

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Niveshika, Verma E, Maurya SK, Mishra R, Mishra AK (2017) The combined use of in silico, in vitro, and in vivo analyses to assess anti-cancerous potential of a bioactive compound from Cyanobacterium Nostoc sp. MGL001. Front Pharmacol 8:873

  53. Riccio G, Lauritano C (2020) Microalgae with immunomodulatory activities. Mar Drugs 18:2

    Article  CAS  Google Scholar 

  54. Ávila-Román J, García-Gil S, Rodríguez-Luna A, Motilva V, Talero E (2021) Anti-inflammatory and anticancer effects of microalgal carotenoids. Mar Drugs 19:531

    Article  PubMed  PubMed Central  Google Scholar 

  55. Jayappriyan KR, Rajkumar R, Venkatakrishnan V, Nagaraj S, Rengasamy R (2013) In vitro anticancer activity of natural β-carotene from Dunaliella salina EU5891199 in PC-3 cells. Biomed Prev Nutr 3:99–105

    Article  Google Scholar 

  56. Lee NY, Kim Y, Kim YS, Shin JH, Rubin LP, Kim Y (2020) β-Carotene exerts anti-colon cancer effects by regulating M2 macrophages and activated fibroblasts. J Nutr Biochem 82:108402

    Article  CAS  PubMed  Google Scholar 

  57. Kim D, Kim Y, Kim Y (2019) Effects of β-carotene on expression of selected microRNAs, histone acetylation, and DNA methylation in colon cancer stem cells. J Cancer Prev 24:224–232

    Article  PubMed  PubMed Central  Google Scholar 

  58. Cui B, Liu S, Wang Q, Lin X (2012) Effect of β-carotene on immunity function and tumor growth in hepatocellular carcinoma rats. Molecules 17:8595–8603

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Gao M, Dang F, Deng C (2019) β-Cryptoxanthin induced anti-proliferation and apoptosis by G0/G1 arrest and AMPK signal inactivation in gastric cancer. Eur J Pharmacol 859:172528

    Article  CAS  PubMed  Google Scholar 

  60. Iskandar AR, Liu C, Smith DE, Hu KQ, Choi SW, Ausman LM, Wang XD (2013) β-cryptoxanthin restores nicotine-reduced lung SIRT1 to normal levels and inhibits nicotine-promoted lung tumorigenesis and emphysema in A/J mice. Cancer Prev Res 6:309–320

    Article  CAS  Google Scholar 

  61. Iskandar AR, Miao B, Li X, Hu KQ, Liu C, Wang XD (2016) β-Cryptoxanthin reduced lung tumor multiplicity and inhibited lung cancer cell motility by downregulating nicotinic acetylcholine receptor A7 signaling. Cancer Prev Res 9:875–886

    Article  CAS  Google Scholar 

  62. Millán CS, Soldevilla B, Martín P, Gil-Calderón B, Compte M, Pérez-Sacristán B, Donoso E, Peña C, Romero J, Granado-Lorencio F, Bonilla F, Domínguez G (2015) β-Cryptoxanthin synergistically enhances the antitumoral activity of oxaliplatin through ΔNP73 negative regulation in colon cancer. Clin Cancer Res 21:4398–4409

    Article  Google Scholar 

  63. Faraone I, Sinisgalli C, Ostuni A, Armentano MF, Carmosino M, Milella L, Russo D, Labanca F, Khan H (2020) Astaxanthin anticancer effects are mediated through multiple molecular mechanisms: a systematic review. Pharmacol Res 155:104689

    Article  CAS  PubMed  Google Scholar 

  64. Badak B, Aykanat NEB, Kacar S, Sahinturk V, Arik D, Canaz F (2021) Effects of astaxanthin on metastasis suppressors in ductal carcinoma A preliminary study. Ann Ital Chir 92:565–574

    PubMed  Google Scholar 

  65. Su X-Z, Chen R, Wang C-B, Ouyang X-L, Jiang Y, Zhu M-Y (2019) Astaxanthin combine with human serum albumin to abrogate cell proliferation, migration, and drug-resistant in human ovarian carcinoma SKOV3 cells. Anticancer Agents Med Chem 19:792–801

    Article  CAS  PubMed  Google Scholar 

  66. Palozza P, Torelli C, Boninsegna A, Simone R, Catalano A, Mele MC, Picci N (2009) Growth-inhibitory effects of the astaxanthin-rich alga Haematococcus pluvialis in human colon cancer cells. Cancer Lett 283:108–117

    Article  CAS  PubMed  Google Scholar 

  67. Nakao R, Nelson OL, Park JS, Mathison BD, Thompson PA, Chew BP (2010) Effect of dietary astaxanthin at different stages of mammary tumor initiation in BALB/c mice. Anticancer Res 30:2171–2175

    CAS  PubMed  Google Scholar 

  68. Xu S, Chaudhary O, Rodríguez-Morales P, Sun X, Chen D, Zappasodi R, Xu Z et al (2021) Uptake of oxidized lipids by the scavenger receptor CD36 promotes lipid peroxidation and dysfunction in CD8+ T cells in tumors. Immunity 54:1561-1577.e7

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Atalay PB, Kuku G, Tuna BG (2019) Effects of carbendazim and astaxanthin co-treatment on the proliferation of MCF-7 breast cancer cells. In Vitro Cell Dev Biol Anim 55:113–119

    Article  CAS  PubMed  Google Scholar 

  70. Kavalappa YP, Gopal SS, Ponesakki G (2021) Lutein inhibits breast cancer cell growth by suppressing antioxidant and cell survival signals and induces apoptosis. J Cell Physiol 236:1798–1809

    Article  CAS  PubMed  Google Scholar 

  71. Baraya YSA, Yankuzo HM, Wong KK, Yaacob NS (2021) Strobilanthes crispus bioactive subfraction inhibits tumor progression and improves hematological and morphological parameters in mouse mammary carcinoma model. J Ethnopharmacol 267:113522

    Article  CAS  PubMed  Google Scholar 

  72. Reynoso-Camacho R, González-Jasso E, Ferriz-Martínez R, Villalón-Corona B, Loarca-Pina GF, Salgado LM, RamosGomez M (2011) Dietary supplementation of lutein reduces colon carcinogenesis in DMH-treated rats by modulating K-ras, PKB, and β-catenin proteins. Nutr Cancer 63:39–45

    CAS  PubMed  Google Scholar 

  73. Sindhu ER, Firdous AP, Ramnath V, Kuttan R (2013) Effect of carotenoid lutein on N-nitrosodiethylamine-induced hepatocellular carcinoma and its mechanism of action. Eur J Cancer Prev 22:320–327

    Article  CAS  PubMed  Google Scholar 

  74. Luan RL, Wang PC, Yan MX, Chen J (2018) Effect of lutein and doxorubicin combinatorial therapy on S180 cell proliferation and tumor growth. Eur Rev Med Pharmacol Sci 22:1514–1520

    PubMed  Google Scholar 

  75. Sheng Y-N, Luo Y-H, Liu S-B, Xu W-T, Zhang Y, Zhang T, Xue H, Zuo W-B, Li Y-N, Wang C-Y, Jin C-H (2020) Zeaxanthin induces apoptosis via ROS-regulated MAPK and Akt signaling pathway in human gastric cancer cells. Onco Targets Ther 2020:10995–11006

    Article  Google Scholar 

  76. Kumar SR, Hosokawa M, Miyashita K (2013) Fucoxanthin: a marine carotenoid exerting anti-cancer effects by affecting multiple mechanisms. Mar Drugs 11:5130–5147

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Yu RX, Hu XM, Xu SQ, Jiang ZJ, Yang W (2011) Effects of fucoxanthin on proliferation and apoptosis in human gastric adenocarcinoma MGC-803 cells via JAK/STAT signal pathway. Eur J Pharmacol 657:10–19

    Article  CAS  PubMed  Google Scholar 

  78. Yu RX, Yu RT, Liu Z (2018) Inhibition of two gastric cancer cell lines induced by fucoxanthin involves downregulation of Mcl-1 and STAT3. Hum Cell 31:50–63

    Article  CAS  PubMed  Google Scholar 

  79. Kawee-Ai A, Kim SM (2014) Application of microalgal fucoxanthin for the reduction of colon cancer risk: Inhibitory activity of fucoxanthin against beta-glucuronidase and DLD-1 cancer cells. Nat Prod Commun 9:921–924

    CAS  PubMed  Google Scholar 

  80. Tamura S, Narita T, Fujii G, Miyamoto S, Hamoya T, Kurokawa Y, Takahashi M, Miki K, Matsuzawa Y, Komiya M, Terasaki M, Yano T, Mutoh M (2019) Inhibition of NF-kappaB transcriptional activity enhances fucoxanthinol-induced apoptosis in colorectal cancer cells. Genes Environ 41:1

    Article  PubMed  PubMed Central  Google Scholar 

  81. Lopes-Costa E, Abreu M, Gargiulo D, Rocha E, Ramos AA (2017) Anticancer effects of seaweed compounds fucoxanthin and phloroglucinol, alone and in combination with 5-fluorouracil in colon cells. J Toxicol Environ Health A 80:776–787

    Article  CAS  PubMed  Google Scholar 

  82. Ganesan P, Noda K, Manabe Y, Ohkubo T, Tanaka Y, Maoka T, Sugawara T, Hirata T (2011) Siphonaxanthin, a marine algal carotenoids from green algae, effectively induces apoptosis in human leukemia (HL-60) cells. Biochim Biophys Acta 1810:497–503

    Article  CAS  PubMed  Google Scholar 

  83. Mei CH, Zhou SC, Zhu L, Ming JX, Zeng FD, Xu R (2017) Antitumor effects of Laminaria extract fucoxanthin on lung cancer. Mar Drugs 15:39

    Article  PubMed  PubMed Central  Google Scholar 

  84. Garg S, Afzal S, Elwakeel A, Sharma D, Radhakrishnan N, Dhanjal JK, Sundar D, Kaul SC, Wadhwa R (2019) Marine carotenoid fucoxanthin possesses anti-metastasis activity: Molecular evidence. Mar Drugs 17:338

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Costa JAV, Lucas BF, Alvarenga AGP, Moreira JB, de Morais MG (2021) Microalgae polysaccharides: an overview of production, characterization, and potential applications. Polysaccharides 2:759–772

    Article  CAS  Google Scholar 

  86. Abidizadegan M, Peltomaa E, Blomster J (2021) The potential of Cryptophyte algae in biomedical and pharmaceutical applications. Front Pharmacol 11:618836

    Article  PubMed  PubMed Central  Google Scholar 

  87. Delattre C, Pierre G, Laroche C, Michaud P (2016) Production, extraction and characterization of microalgal and cyanobacterial exopolysaccharides. Biotechnol Adv 34:1159–1179

    Article  CAS  PubMed  Google Scholar 

  88. Kumar D, Kaštánek P, Adhikary SP (2018) Exopolysaccharides from cyanobacteria and microalgae and their commercial application. Curr Sci 115:234–241

    Article  CAS  Google Scholar 

  89. Zhang J, Liu L, Ren Y, Chen F (2019) Characterization of exopolysaccharides produced by microalgae with antitumor activity on human colon cancer cells. Int J Biol Macromol 128:761–767

    Article  CAS  PubMed  Google Scholar 

  90. Gargouch N, Elleuch F, Karkouch I, Tabbene O, Pichon C, Gardarin C, Rihouey C, Picton L, Abdelkafi S, Fendri I, Laroche C (2021) Potential of exopolysaccharide from Porphyridium marinum to contend with bacterial proliferation, biofilm formation, and breast cancer. Mar Drugs 19:66

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Li H, Su L, Chen S, Zhao L, Wang H, Ding F, Chen H, Shi R, Wang Y, Huang Z (2018) Physicochemical characterization and functional analysis of the polysaccharide from the edible microalga Nostoc sphaeroides. Molecules 23:508

    Article  PubMed  PubMed Central  Google Scholar 

  92. Umemura K, Yanase K, Suzuki M, Okutani K, Yamori T, Andoh T (2003) Inhibition of DNA topoisomerases I and II, and growth inhibition of human cancer cell lines by a marine microalgal polysaccharide. Biochem Pharmacol 66:481–487

    Article  CAS  PubMed  Google Scholar 

  93. Amaro HM, Barros R, Guedes AC, Sousa-Pinto I, Malcata FX (2013) Microalgal compounds modulate carcinogenesis in the gastrointestinal tract. Trends Biotechnol 31:92–98

    Article  CAS  PubMed  Google Scholar 

  94. Kusaikin MI, Ermakova SP, Shevchenko NM, Isakov VV, Gorshkov AG, Vereshchagin AL, Grachev MA, Zvyagintseva TN (2010) Structural characteristics and antitumor activity of a new chrysolaminaran from the diatom alga Synedra acus. Chem Nat Compd 46:1–4

    Article  CAS  Google Scholar 

  95. Kang K-H, Kim S-K (2013) Beneficial effect of peptides from microalgae on anticancer. Curr Protein Pept Sci 14:212–217

    Article  CAS  PubMed  Google Scholar 

  96. Nethravathy MU, Mehar JG, Mudliar SN, Shekh AY (2019) Recent advances in microalgal bioactives for food, feed, and healthcare products: Commercial potential, market space, and sustainability. Compr Rev Food Sci Food Saf 18:1882–1897

    Article  Google Scholar 

  97. Sheih IC, Fang TJ, Wu TK, Lin PH (2010) Anticancer and antioxidant activities of the peptide fraction from algae protein waste. J Agric Food Chem 58:1202–1207

    Article  CAS  PubMed  Google Scholar 

  98. Wang X, Zhang X (2013) Separation, antitumor activities, and encapsulation of polypeptide from Chlorella pyrenoidosa. Biotechnol Prog 29:681–687

    Article  PubMed  Google Scholar 

  99. Ko S-C, Heo S-Y, Choi S-W, Qian Z-J, Heo S-J, Kang D-H, Kim N, Jung W-K (2018) A heptameric peptide isolated from the marine microalga Pavlova lutheri suppresses PMA-induced secretion of matrix metalloproteinase-9 through the inactivation of the JNK, p38, and NF-κB pathways in human fibrosarcoma cells. J Appl Phycol 30:2367–2378

    Article  CAS  Google Scholar 

  100. Oh G-W, Ko S-C, Heo S-Y, Nguyen V-T, Kim GH, Jang CH, Park WS, Choi I-W, Qian Z-J, Jung W-K (2015) A novel peptide purified from the fermented microalga Pavlova lutheri attenuates oxidative stress and melanogenesis in B16F10 melanoma cells. Process Biochem 50:1318–1326

    Article  CAS  Google Scholar 

  101. Jiang L, Wang Y, Yin Q, Liu G, Liu H, Huang Y, Li B (2017) Phycocyanin: a potential drug for cancer treatment. J Cancer 8:3416–3429

    Article  PubMed  PubMed Central  Google Scholar 

  102. Alishah Aratboni H, Rafiei N, Garcia-Granados R, Alemzadeh A, Morones-Ramίrez JR (2019) Biomass and lipid induction strategies in microalgae for biofuel production and other applications. Microb Cell Fact 18:178

    Article  PubMed  PubMed Central  Google Scholar 

  103. Jóźwiak M, Filipowska A, Fiorino F, Struga M (2020) Anticancer activities of fatty acids and their heterocyclic derivatives. Eur J Pharmacol 871:172937

    Article  PubMed  Google Scholar 

  104. Marventano S, Kolacz P, Castellano S, Galvano F, Buscemi S, Mistretta A, Grosso G (2015) A review of recent evidence in human studies of n-3 and n-6 PUFA intake on cardiovascular disease, cancer, and depressive disorders: does the ratio really matter? Int J Food Sci Nutr 66:611–622

    Article  CAS  PubMed  Google Scholar 

  105. Noel SE, Stoneham AC, Olsen CM, Rhodes LE, Green AC (2014) Consumption of omega-3 fatty acids and the risk of skin cancers: a systematic review and meta-analysis. Int J Cancer 135:149–156

    Article  CAS  PubMed  Google Scholar 

  106. Abd-Ellatef GF, Ahmed OM, Abdel-Reheim ES, Abdel-Hamid AZ (2017) Ulva lactuca polysaccharides prevent Wistar rat breast carcinogenesis through the augmentation of apoptosis, enhancement of antioxidant defense system, and suppression of inflammation. Breast Cancer (Dove Med Press) 9:67–83

    CAS  PubMed  Google Scholar 

  107. Das UN (2004) From bench to the clinic: gamma-linolenic acid therapy of human gliomas. Prostagl Leukot Essent Fat Acids 70:539–552

    Article  CAS  Google Scholar 

  108. So WW, Liu WN, Leung KN (2015) Omega-3 polyunsaturated fatty acids trigger cell cycle arrest and induce apoptosis in human neuroblastoma LA-N-1 cells. Nutrients 7:6956–6973

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  109. Fulda S, Debatin K-M (2006) Extrinsic versus intrinsic apoptosis pathways in anticancer chemotherapy. Oncogene 25:4798–4811

    Article  CAS  PubMed  Google Scholar 

  110. Albino AP, Juan G, Traganos F, Reinhart L, Connolly J, Rose DP, Darzynkiewicz Z (2000) Cell cycle arrest and apoptosis of melanoma cells by docosahexaenoic acid: association with decreased pRb phosphorylation. Cancer Res 60:4139–4145

    CAS  PubMed  Google Scholar 

  111. Paul C, Reunamo A, Lindehoff E, Bergkvist J, Mausz MA, Larsson H, Richter H, Wängberg S-Å, Leskinen P, Båmstedt U, Pohnert G (2012) Diatom derived polyunsaturated aldehydes do not structure the planktonic microbial community in a mesocosm study. Mar Drugs 10:775–792

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  112. Pezzolesi L, Pichierri S, Samorì C, Totti C, Pistocchi R (2017) PUFAs and PUAs production in three benthic diatoms from the northern Adriatic Sea. Phytochemistry 142:85–91

    Article  CAS  PubMed  Google Scholar 

  113. Miralto A, Barone G, Romano G, Poulet SA, Ianora A, Russo GL, Buttino I, Mazzarella G, Laabir M, Cabrini M, Giacobbe MG (1999) The insidious effect of diatoms on copepod reproduction. Nature 402:173–176

    Article  CAS  Google Scholar 

  114. Sansone C, Braca A, Ercolesi E, Romano G, Palumbo A, Casotti R, Francone M, Ianora A (2014) Diatom-derived polyunsaturated aldehydes activate cell death in human cancer cell lines but not normal cells. PLoS ONE 9:e101220

    Article  PubMed  PubMed Central  Google Scholar 

  115. Martin-Creuzburg D, Merkel P (2016) Sterols of freshwater microalgae: potential implications for zooplankton nutrition. J Plankton Res 38:865–877

    Article  CAS  Google Scholar 

  116. Randhir A, Laird DW, Maker G, Trengove R, Moheimani NR (2020) Microalgae: a potential sustainable commercial source of sterols. Algal Res 46:101772

    Article  Google Scholar 

  117. Fritz H, Kennedy D, Fergusson D, Fernandes R, Doucette S, Cooley K, Seely A, Sagar S, Wong R, Seely D (2011) Vitamin A and retinoid derivatives for lung cancer: a systematic review and meta analysis. PLoS ONE 6:e21107

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  118. Tripathi SK, Pandey K, Panda M, Spinella MJ, Rengasamy KR, Biswal BK (2019) The potential of retinoids for combination therapy of lung cancer: updates and future directions. Pharmacol Res 147:104331

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  119. Blaszczak W, Barczak W, Masternak J, Kopczyński P, Zhitkovich A, Rubiś B (2019) Vitamin C as a modulator of the response to cancer therapy. Molecules 24:453

    Article  PubMed  PubMed Central  Google Scholar 

  120. Satheesh NJ, Samuel SM, Büsselberg D (2020) Combination therapy with vitamin C could eradicate cancer stem cells. Biomolecules 10:79

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  121. Chatterjee R, Erban JK, Fuss P, Dolor R, LeBlanc E, Staten M, Sheehan P, Pittas A, D2d Research Group (2019) Vitamin D supplementation for prevention of cancer: The D2d cancer outcomes (D2dCA) study. Contemp Clin Trials 81:62–70

  122. Urashima M, Okuyama M, Akutsu T, Ohdaira H, Kaji M, Suzuki Y (2020) Effect of vitamin D supplementation on survival of digestive tract cancer patients with low bioavailable 25-hydroxyvitamin D levels: A post hoc analysis of the AMATERASU randomized clinical trial. Cancers 12:347

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  123. Köpcke W (2019). In: Weber P, Birringer M, Blumberg JB, Eggersdorfer M, Frank J (eds) Vitamin E in human health. Humana Press, Switzerland, p 467

    Google Scholar 

  124. Abraham A, Kattoor AJ, Saldeen T, Mehta JL (2019) Vitamin E and its anticancer effects. Crit Rev Food Sci Nutr 59:2831–2838

    Article  CAS  PubMed  Google Scholar 

  125. Dasari S, Ali SM, Zheng G, Chen A, Dontaraju VS, Bosland MC, Kajdacsy-Balla A, Munirathinam G (2017) Vitamin K and its analogs: potential avenues for prostate cancer management. Oncotarget 8:57782–57799

    Article  PubMed  PubMed Central  Google Scholar 

  126. Patil RY, More HN (2020) Antioxidants with multivitamin and mineral supplementation attenuates chemotherapy or radiotherapy-induced oxidative stress in cancer patients. Indian J Pharm Educ Res 54:484–490

    Article  CAS  Google Scholar 

  127. Goiris K, Muylaert K, Voorspoels S, Noten B, De Paepe D, Baart GJE, De Cooman L (2014) Detection of flavonoids in microalgae from different evolutionary lineages. J Phycol 50:483–492

    Article  CAS  PubMed  Google Scholar 

  128. Jayshree A, Jayashree S, Thangaraju N (2016) Chlorella vulgaris and Chlamydomonas reinhardtii: effective antioxidant, antibacterial and anticancer mediators. Indian J Pharm Sci 78:575–581

    Article  CAS  Google Scholar 

  129. Raizner AE (2019) Coenzyme Q10. Methodist Debakey Cardiovasc J 15:185–191

    Article  PubMed  PubMed Central  Google Scholar 

  130. Hafsa MB, Ismail MB, Garrab M, Aly R, Gagnon J, Naghmouchi K (2017) Antimicrobial, antioxidant, cytotoxic and anticholinesterase activities of water-soluble polysaccharides extracted from microalgae Isochrysis galbana and Nannochloropsis oculata. J Serbian Chem Soc 82:509–522

    Article  Google Scholar 

  131. Nigjeh SE, Yusoff F, Banu N, Alitheen M, Rasoli M, Keong YS, Rahman A (2013) Cytotoxic effect of ethanol extract of microalga, Chaetoceros calcitrans, and its mechanisms in inducing apoptosis in human breast cancer cell line. Biomed Res Int 2013:1–9

    Article  Google Scholar 

  132. Shah MR, Samarakoon KW, Ko J-Y, Lakmal HHC, Lee J-H, An S-J, Jeon Y-J, Lee J-B (2014) Potentiality of benthic dinoflagellate cultures and screening of their bioactivities in Jeju Island, Korea. Afr J Biotechnol 13:792–805

    Article  Google Scholar 

  133. Lauritano C, Andersen JH, Hansen E, Albrigtsen M, Escalera L, Esposito F, Helland K, Hanssen KØ, Romano G, Ianora A (2016) Bioactivity screening of microalgae for antioxidant, anti-inflammatory, anticancer, anti-diabetes and antibacterial activities. Front Mar Sci 3:1–12

    Article  Google Scholar 

  134. Lin P-Y, Tsai C-T, Chuang W-L, Chao Y-H, Pan I-H, Chen Y-K, Lin C-C, Wang B-Y (2017) Chlorella sorokiniana induces mitochondrial-mediated apoptosis in human non-small cell lung cancer cells and inhibits xenograft tumor growth in vivo. BMC Complement Altern Med 17:88

    Article  PubMed  PubMed Central  Google Scholar 

  135. Pasquet V, Morisset P, Ihammouine S, Chepied A, Aumailley L, Berard JB, Serive B, Kaas R, Lanneluc I, Thiery V, Lafferriere M, Piot J-M, Patrice T, Cadoret J-P, Picot L (2011) Antiproliferative activity of violaxanthin isolated from bioguided fractionation of Dunaliella tertiolecta extracts. Mar Drugs 9:819–831

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  136. Kim Y-S, Li X-F, Kang K-H, Ryu B, Kim S-K (2014) Stigmasterol isolated from marine microalgae Navicula incerta induces apoptosis in human hepatoma HepG2 cells. BMB Rep 47:433–438

    Article  PubMed  PubMed Central  Google Scholar 

  137. Su J, Guo K, Huang M, Liu Y, Zhang J, Sun L, Li D, Pang K-L, Wang G, Chen L, Liu Z, Chen Y, Chen Q, Huang L (2019) Fucoxanthin, a marine xanthophyll isolated from Conticribra weissflogii ND-8: preventive anti-inflammatory effect in a mouse model of sepsis. Front Pharmacol 10:906–917

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  138. Juin C, de Oliveira Junior RG, Fleury A, Oudinet C, Pytowski L, Bérard JB, Nicolau E, Thiéry V, Lanneluc I, Beaugeard L, Prunier G, Da Silva Almeida JRG, Picot L (2018) Zeaxanthin from Porphyridium purpureum induces apoptosis in human melanoma cells expressing the oncogenic BRAF V600E mutation and sensitizes them to the BRAF inhibitor vemurafenib. Rev Bras Farmacogn 28:457–467

    Article  CAS  Google Scholar 

  139. Sanjeewa KKA, Fernando IPS, Samarakoon KW, Lakmal HHC, Kim EA, Kwon ON, Dilshara MG, Lee J-B, Jeon Y-J (2016) Anti-Inflammatory and anti-cancer activities of sterol rich fraction of cultured marine microalga Nannochloropsis oculata. Algae 31:277–287

    Article  CAS  Google Scholar 

  140. Elkhateeb W, El-Sayed H, Fayad W, Galib Al Kolaibe A, Emam M, Daba G (2020) In vitro anti-breast cancer and antifungal bio-efficiency of some microalgal extracts. Egypt J Aquat Biol 24:263–279

    Article  Google Scholar 

  141. Montone CM, Capriotti AL, Cavaliere C, La Barbera G, Piovesana S, Zenezini Chiozzi R, Laganà A (2018) Peptidomic strategy for purification and identification of potential ACE-inhibitory and antioxidant peptides in Tetradesmus obliquus microalgae. Anal Bioanal Chem 410:3573–3586

    Article  CAS  PubMed  Google Scholar 

  142. Nagaraj S, Rajaram MG, Arulmurugan P, Baskaraboopathy A, Karuppasamy K, Jayappriyan KR, Sundararaj R, Rengasamy R (2012) Antiproliferative potential of astaxanthin rich alga Haematococcus pluvialis flotow on human hepatic cancer (HepG2) cell line. Biomed Prev Nutr 2:149–153

    Article  Google Scholar 

  143. Yang S, Wan H, Wang R, Hao D (2019) Sulfated polysaccharides from Phaeodactylum tricornutum: Isolation, structural characteristics, and inhibiting HepG2 growth activity in vitro. PeerJ 7:e6409

    Article  PubMed  PubMed Central  Google Scholar 

  144. Rosa GP, Tavares WR, Sousa PMC, Pagѐs AK, Seca AML, Pinto DCGA (2020) Seaweed secondary metabolites with beneficial health effects: An overview of successes in in vivo studies and clinical trials. Mar Drugs 18:8

    Article  CAS  Google Scholar 

  145. Ciancia M, Fernández PV, Leliaert F (2020) Diversity of sulfated polysaccharides from cell walls of coenocytic green algae and their structural relationships in view of green algal evolution. Front Plant Sci 11:554585

    Article  PubMed  PubMed Central  Google Scholar 

  146. Khotimchenko M, Tiasto V, Kalitnik A, Begun M, Khotimchenko R, Leonteva E, Bryukhovetskiy I, Khotimchenko Y (2020) Antitumor potential of carrageenans from marine red algae. Carbohydr Polym 246:116568

    Article  CAS  PubMed  Google Scholar 

  147. Thanh TTT, Quach TMT, Nguyen TN, Luong DV, Bui ML, Tran TTV (2016) Structure and cytotoxic activity of ulvan extracted from green seaweed Ulva lactuca. Int J Biol Macromol 93:695–702

    Article  CAS  PubMed  Google Scholar 

  148. Berri M, Olivier M, Holbert S, Dupont J, Demais H, Goff ML, Collen PN (2017) Ulvan from Ulva armoricana (Chlorophyta) activates the PI3K/Akt signalling pathway via TLR4 to induce intestinal cytokine production. Algal Res 28:39–47

    Article  Google Scholar 

  149. Wang X, Chen Y, Wang J, Liu Z, Zhao S (2014) Antitumor activity of a sulfated polysaccharide from Enteromorpha intestinalis targeted against hepatoma through mitochondrial pathway. Tumor Biol 35:1641–1647

    Article  CAS  Google Scholar 

  150. Lajili S, Ammar HH, Mzoughi Z, Amor HBH, Muller CD, Majdoub H, Bouraoui A (2019) Characterization of sulfated polysaccharide from Laurecia obtusa and its apoptotic, gastroprotective and antioxidant activities. Int J Biol Macromol 126:326–336

    Article  CAS  PubMed  Google Scholar 

  151. Tabatabai R, Linhares Y, Bolos D, Mita M, Mita A (2017) Targeting the Wnt pathway in cancer: a review of novel therapeutics. Target Oncol 12:623–641

    Article  PubMed  Google Scholar 

  152. Bilal M, Iqbal HMN (2020) Marine seaweed polysaccharides-based engineered cues for the modern biomedical sector. Mar Drugs 18:7

    Article  CAS  Google Scholar 

  153. van Weelden G, Bobiński M, Okła K, van Weelden WJ, Romano A, Pijnenborg JMA (2019) Fucoidan structure and activity in relation to anti-cancer mechanisms. Mar Drugs 17:32

    Article  PubMed  PubMed Central  Google Scholar 

  154. Atashrazm F, Lowenthal RM, Woods GM, Holloway AF, Dickinson JL (2015) Fucoidan and cancer: a multifunctional molecule with anti-tumor potential. Mar Drugs 13:2327–2346

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  155. Cho TM, Kim WJ, Moon SK (2014) AKT signaling is involved in fucoidan-induced inhibition of growth and migration of human bladder cancer cells. Food Chem Toxicol 64:344–352

    Article  CAS  PubMed  Google Scholar 

  156. Hsu HY, Lin TY, Wu YC, Tsao SM, Hwang PA, Shih YW, Hsu J (2014) Fucoidan inhibition of lung cancer in vivo and in vitro: role of the Smurf2-dependent ubiquitin proteasome pathway in TGF beta receptor degradation. Oncotarget 5:7870–7885

    Article  PubMed  PubMed Central  Google Scholar 

  157. Wang WC, Chen HJ, Zhang L, Qin Y, Cong QF, Wang PP, Ding K (2016) A fucoidan from Nemacystus decipiens disrupts angiogenesis through targeting bone morphogenetic protein 4. Carbohydr Polym 144:305–314

    Article  CAS  PubMed  Google Scholar 

  158. Yang YZ, Gao ZX, Ma YH, Teng HM, Liu ZD, Wei HY, Lu YB, Cheng XF, Hou L, Zou XY (2016) Fucoidan inhibits lymphangiogenesis by downregulating the expression of VEGFR3 and PROX1 in human lymphatic endothelial cells. Oncotarget 7:38025–38035

    Article  PubMed  PubMed Central  Google Scholar 

  159. Zhang JQ, Riby JE, Conde L, Grizzle WE, Cui XQ, Skibola CF (2016) A Fucus vesiculosus extract inhibits estrogen receptor activation and induces cell death in female cancer cell lines. BMC Complement Altern Med 16:151

    Article  PubMed  PubMed Central  Google Scholar 

  160. Xie M, Liu D, Yang Y (2020) Anti-cancer peptides: Classification, mechanism of action, reconstruction and modification. Open Biol 10:200004

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  161. Mao X, Bai L, Fan X, Zhang X (2017) Anti-proliferation peptides from protein hydrolysates of Pyropia haitanensis. J Appl Phycol 29:1623–1633

    Article  CAS  Google Scholar 

  162. Fan X, Bai L, Mao X, Zhang X (2017) Novel peptides with anti-proliferation activity from the Porphyra haitanesis hydrolysate. Process Biochem 60:98–107

    Article  CAS  Google Scholar 

  163. Nadeeshani H, Hassouna A, Lu J (2021) Proteins extracted from seaweed Undaria pinnatifida and their potential uses as foods and nutraceuticals. Crit Rev Food Sci Nutr 62:6187–6203

    Article  PubMed  Google Scholar 

  164. Rafiquzzaman SM, Kim E-Y, Kim Y-R, Nam T-J, Kong I-S (2013) Antioxidant activity of glycoprotein purified from Undaria pinnatifida measured by an in vitro digestion model. Int J Biol Macromol 62:265–272

    Article  CAS  PubMed  Google Scholar 

  165. Zanoaga O, Jurj A, Raduly L, Cojocneanu-Petric R, Fuentes-Mattei E, Wu O, Braicu C, Gherman CD, Berindan-Neagoe I (2018) Implications of dietary ω-3 and ω-6 polyunsaturated fatty acids in breast cancer (review). Exp Ther Med 15:1167–1176

    CAS  PubMed  Google Scholar 

  166. Fabian CJ, Kimier BF, Hursting SD (2015) Omega-3 fatty acids for breast cancer prevention and survivorship. Breast Cancer Res 17:62

    Article  PubMed  PubMed Central  Google Scholar 

  167. Gu Z, Suburu J, Chen H, Chen YQ (2013) Mechanisms of omega-3 polyunsaturated fatty acids in prostate cancer prevention. Biomed Res Int 2013:824563

    Article  PubMed  PubMed Central  Google Scholar 

  168. Ganesan AR, Tiwari U, Rajauria G (2019) Seaweed nutraceuticals and their therapeutic role in disease prevention. Food Sci Hum Wellness 8:252–263

    Article  Google Scholar 

  169. Cotas J, Leandro A, Pacheco D, Gonçalves AMM, Pereira LA (2020) A comprehensive review of the nutraceutical and therapeutic applications of red seaweeds (Rhodophyta). Life (Basel) 10:19

    CAS  PubMed  Google Scholar 

  170. Kumari P, Kumar M, Reddy CRK, Jha B (2013) Algal lipids, fatty acids and sterols. In: Domίnguez H (ed) Functional ingredients from algae for foods and nutraceuticals. Woodhead Publishing, pp 87–134

    Chapter  Google Scholar 

  171. Pacheco BS, dos Santos MAZ, Schultze E, Martins RM, Lund RG, Seixas FK, Colepicolo P, Collares T, Paula FR, De Pereira CMP (2018) Cytotoxic activity of fatty acids from Antarctic macroalgae on the growth of human breast cancer cells. Front Bioeng Biotechnol 6:185

    Article  PubMed  PubMed Central  Google Scholar 

  172. Grozdanic N, Djuricic I, Kosanic M, Zdunic G, Savikin K, Etahiri S, Assobhei O, Benba J, Petovic S, Matic IZ, Stanojkovic TP (2020) Fucus spiralis extract and fractions: Anticancer and pharmacological potentials. J BUON 25:1219–1229

    PubMed  Google Scholar 

  173. Hentati F, Tounsi L, Djomdi D, Pierre G, Delattre C, Ursu AV, Fendri I, Abdelkafi S, Michaud P (2020) Bioactive polysaccharides from seaweeds. Molecules 25:3152

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  174. Li X, Zhou D-Y, Li F-T, Jiang Y-F, Dai Y-L, Jeon Y-J (2022) Saringosterol acetate isolated from Sargassum fusiformis induces mitochondrial-mediated apoptosis in MCF-7 breast cancer cells. Chem Biodivers 19:e202100848

    CAS  PubMed  Google Scholar 

  175. Bae H, Lee J-Y, Song G, Lim W (2020) Fucosterol suppresses the progression of human ovarian cancer by inducing mitochondrial dysfunction and endoplasmic reticulum stress. Mar Drugs 18:261

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  176. Hamed I, Özogul F, Özogul Y, Regenstein JM (2015) Marine bioactive compounds and their health benefits: a review. Compr Rev Food Sci Food Saf 14:446–465

    Article  CAS  Google Scholar 

  177. Zuo L, Yang X, Lu M, Hu R, Zhu H, Zhang S, Zhou Q, Chen F, Gui S, Wang Y (2016) All-trans retinoic acid inhibits human colorectal cancer cells RKO migration via downregulating myosin light chain kinase expression through MAPK signaling pathway. Nutr Cancer 68:1225–1233

    Article  CAS  PubMed  Google Scholar 

  178. Peterson CT, Rodionov DA, Osterman AL, Peterson SN (2020) B vitamins and their role in immune regulation and cancer. Nutrients 12:3380

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  179. Liu X, Montissol S, Uber A, Ganley S, Grossestreuer AV, Berg K, Heydrick S, Donnino MW (2018) The effects of thiamine on breast cancer cells. Molecules 23:1464

    Article  PubMed  PubMed Central  Google Scholar 

  180. Matsuo T, Fujiwara A, Nakamura K, Sadzuka Y (2018) The effects of vitamin B6 compounds on cell proliferation and melanogenesis in B16F10 melanoma cells. Oncol Lett 15:5181–5184

    PubMed  PubMed Central  Google Scholar 

  181. Padmanabhan S, Waly MI, Taranikanti V, Guizani N, Ali A, Rahman MS, Al-Attabi Z, Al-Malky RN, Al-Maskari SNM, Al-Ruqaishi BRS, Dong J, Deth RC (2019) Folate/vitamin B12 supplementation combats oxidative-stress associated carcinogenesis in a rat model of colon cancer. Nutr Cancer 71:100–110

    Article  CAS  PubMed  Google Scholar 

  182. Yiang G-T, Chou P-L, Hung Y-T, Chen J-N, Chang W-J, Yu Y-L, Wei C-W (2014) Vitamin C enhances anticancer activity in methotrexate-treated Hep3B hepatocellular carcinoma cells. Oncol Rep 32:1057–1063

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  183. Wang X, Xu S, Zhang L, Cheng X, Yu H, Bao J, Lu R (2021) Viatmin C induces ferroptosis in anaplastic thyroid cancer cells by ferritinophagy activation. Biochem Biophys Res Commun 551:46–53

    Article  CAS  PubMed  Google Scholar 

  184. Jeon S-M, Shin E-A (2018) Exploring vitamin D metabolism and function in cancer. Exp Mol Med 50:20

    Article  PubMed  PubMed Central  Google Scholar 

  185. Bajbouj K, Sahnoon L, Shafarin J, Al-Ali A, Muhammad JS, Karim A, Guraya SY, Hamad M (2021) Vitamin D-mediated anti-cancer activity involves iron homeostatic balance disruption and oxidative stress induction in breast cancer. Front Cell Dev Biol 9:766978

    Article  PubMed  PubMed Central  Google Scholar 

  186. Casadei-Gardini A, Filippi R, Rimini M, Rapposelli IG, Fornaro L, Silvestris N, Aldrighetti L et al (2021) Effects of metformin and vitamin D on clinical outcome in cholangiocarcinoma patients. Oncology 99:292–299

    Article  CAS  PubMed  Google Scholar 

  187. Zulkapli R, Razak FA, Zain RB (2017) Vitamin E (α-tocopherol) exhibits antitumor activity on oral squamous carcinoma cells ORL-48. Integr Cancer Ther 16:414–425

    Article  CAS  PubMed  Google Scholar 

  188. Duan F, Mei C, Yang L, Zheng J, Lu H, Xia Y, Hsu S, Liang H, Hong L (2020) Vitamin K2 promotes PI3K/AKT/HIF-1α-mediated glycolysis that leads to AMPK-dependent autophagic cell death in bladder cancer cells. Sci Rep 10:7714

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  189. Almeida TP, Ramos AA, Ferreira J, Azqueta A, Rocha E (2020) Bioactive compounds from seaweed with anti-leukemic activity: a mini-review on carotenoids and phlorotannins. Mini Rev Med Chem 20:39–53

    Article  CAS  PubMed  Google Scholar 

  190. Satomi Y (2017) Antitumor and cancer-preventative function of fucoxanthin: a marine carotenoid. Anticancer Res 37:1557–1562

    Article  CAS  PubMed  Google Scholar 

  191. Wang L, Zeng Y, Liu Y, Hu X, Li S, Wang Y, Li L, Lei Z, Zhang Z (2014) Fucoxanthin induces growth arrest and apoptosis in human bladder cancer T24 cells by upregulation of p21and down-regulation of mortalin. Acta Biochim Biophys Sin 46:877–884

    Article  CAS  PubMed  Google Scholar 

  192. Wang J, Ma Y, Yang J, Jin L, Gao Z, Xue L, Hou L, Sui L, Liu J, Zou X (2019) Fucoxanthin inhibits tumor-related lymphangiogenesis and growth of breast cancer. J Cell Mol Med 23:2219–2229

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  193. Malyarenko OS, Imbs TI, Ermakova SP (2020) In-vitro anticancer and radiosensitizing activities of phlorethols from the brown alga Costaria costata. Molecules 25:3208

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  194. Zenthoefer M, Geisen U, Hofmann-Peiker K, Fuhrmann M, Kerber J, Kirchhӧfer R, Hennig S, Peipp M, Geyer R, Piker L, Kalthoff H (2017) Isolation of polyphenols with anticancer activity from Baltic Sea brown seaweed Fucus vesiculosus using bioassay-guided fractionation. J Appl Phycol 29:2021–2037

    Article  CAS  Google Scholar 

  195. Zhang M, Zhou W, Zhao S, Li S, Yan D, Wang J (2019) Eckol inhibits Reg3A-induced proliferation of human SW1990 pancreatic cancer cells. Exp Ther Med 18:2825–2832

    CAS  PubMed  PubMed Central  Google Scholar 

  196. Wang C-H, Li X-F, Jin L-F, Zhao Y, Zhu G-J, Shen W-Z (2019) Dieckol inhibits non-small-cell lung cancer cell proliferation and migration by regulating the PI3K/AKT signaling pathway. J Biochem Mol Toxicol 33:e22346

    Article  PubMed  PubMed Central  Google Scholar 

  197. Dissanayake IH, Bandaranayake U, Keerthirathna LR, Manawadu C, Silva RM, Mohamed B, Ali R, Peiris DC (2022) Integration of in vitro and in-silico analysis of Caulerpa racemosa against antioxidant, antidiabetic, and anticancer activities. Sci Rep 12:20848

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  198. Mert-Ozupek N, Calibasi-Kocal G, Olgun N, Basbinar Y, Cavas L, Ellidokuz H (2022) In-silico molecular interactions among the secondary metabolites of Caulerpa spp. and colorectal cancer targets. Front Chem 10:1046313

  199. Pangestuti R, Haq M, Rahmadi P, Chun B-S (2021) Nutritional value and biofunctionalities of two edible green seaweeds (Ulva lactuca and Caulerpa racemosa) from Indonesia by subcritical water hydrolysis. Mar Drugs 19:578

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  200. Tanna B, Yadav S, Mishra A (2020) Anti-proliferative and ROS-inhibitory activities reveal the anticancer potential of Caulerpa species. Mol Biol Rep 47:7403–7411

    Article  PubMed  Google Scholar 

  201. Zhao C, Lin G, Wu D, Liu D, You L, Hӧgger P, Simal-Gandara J, Wang M, da Costa JGM, Marunaka Y, Daglia M, Khan H, Filosa R, Wang S, Xiao J (2020) The algal polysaccharide ulvan suppresses growth of hepatoma cells. Food Front 1:83–101

    Article  Google Scholar 

  202. Haq SH, Al-Ruwaished G, Al-Mutlaq MA, Naji SA, Al-Mogren M, Al-Rashed S, Ain QT, Al-Amro AA, Al-Mussallam A (2019) Antioxidant, anticancer activity and phytochemical analysis of green algae, Chaetomorpha collected from the Arabian Gulf. Sci Rep 9:18906

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  203. Marques MLM, Presa FB, Viana RLS, Costa MSSP, Amorim MOR, Bellan DL, Alves MGCF, Costa LS, Trindade ES, Rocha HAO (2019) Anti-thrombin, anti-adhesive, anti-migratory, and anti-proliferative activities of sulfated galactans from the tropical green seaweed, Udotea flabellum. Mar Drugs 17:5

    Article  CAS  Google Scholar 

  204. Barbosa JS, Palhares LCGF, Silva CHF, Sabry DA, Chavante SF, Rocha HAO (2021) In vitro antitumor potential of sulfated polysaccharides from seaweed Caulerpa cupressoides var. flabellata. Mar Biotechnol 23:77–89

    Article  Google Scholar 

  205. Ramkumar VS, Pugazhendhi A, Gopalakrishnan K, Sivagurunathan P, Saratale GD, Dung TNB, Kannapiran E (2017) Biofabrication and characterization of silver nanoparticles using aqueous extract of seaweed Enteromorpha compressa and its biomedical properties. Biotechnol Rep 14:1–7

    Article  Google Scholar 

  206. Senthilkumar D, Jayanthi S (2016) Partial characterization and anticancer activities of purified glycoprotein extracted from green seaweed Codium decorticatum. J Funct Food 25:323–332

    Article  CAS  Google Scholar 

  207. Pragna Lakshmi T, Vajravijayan S, Mondal M, Natarajan S, Gunasekaran K, Krishna R (2018) A novel guaiane sesquiterpene derivative, guai-2-en-10α-ol, from Ulva fasciata Delile inhibits EGFR/PI3K/Avt signaling and induces cytotoxicity in triple-negative breast cancer cells. Mol Cell Biochem 438:123–139

    Article  CAS  PubMed  Google Scholar 

  208. Ropellato J, Carvalho MM, Ferreira LG, Noseda MD, Zuconelli CR, Gonçalves AG, Ducatti DRB, Kenski JCN, Nasato PL, Winnischofer SMB, Duarte MER (2015) Sulfated heterorhamnans from the green seaweed Gayralia oxysperma: partial depolymerization, chemical structure and antitumor activity. Carbohydr Polym 117:476–485

    Article  CAS  PubMed  Google Scholar 

  209. Thangam R, Senthilkumar D, Suresh V, Sathuvan M, Sivasubramanian S, Pazhanichamy K, Gorlagunta PK, Kannan S, Gunasekaran P, Rengasamy R, Sivaraman J (2014) Induction of ROS-dependent mitochondrial-mediated intrinsic apoptosisin MDA-MB-231 cells by glycoprotein from Codium decorticatum. J Agric Food Chem 62:3410–3421

    Article  CAS  PubMed  Google Scholar 

  210. Namvar F, Baharara J, Mahdi AA (2014) Antioxidant and anticancer activities of selected Persian Gulf algae. Ind J Clin Biochem 29:13–20

    Article  CAS  Google Scholar 

  211. Kim YM, Kim IH, Nam TJ (2013) Capsosiphon fulvescens glycoprotein reduces AGS gastric cancer cell migration by downregulating transforming growth factor-β1 and integrin expression. Int J Oncol 43:1059–1065

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  212. Kim YM, Kim IH, Nam TJ (2013) Capsosiphon fulvescens glycoprotein inhibits AGS gastric cancer cell proliferation by downregulating Wnt-1 signaling. Int J Oncol 43:1395–1401

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  213. Yun EJ, Yu S, Kim Y-A, Liu J-J, Kang NJ, Jin Y-S, Kim KH (2021) In vitro prebiotic and anti-colon cancer activities of agar-derived sugars from red seaweeds. Mar Drugs 19:213

    Article  PubMed  PubMed Central  Google Scholar 

  214. Prasedya ES, Ardiana N, Padmi H, Ilhami BTK, Martyasari NWR, Sunarwidhi AL, Nikmatullah A, Widyastuti S, Sunarpi H, Frediansyah A (2021) The antiproliferative and apoptosis-inducing effects of the red macroalgae Gelidium latifolium extract against melanoma cells. Molecules 26:6568

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  215. Pham TNA, Le B, Yang SH (2021) Anticancer activity of the potential Pyropia yezoensis galactan fractioned in human prostate cancer cells. Biotechnol Bioproc E 26:63–70

    Article  CAS  Google Scholar 

  216. He D, Yan L, Ma X, Cheng Y, Wu S, Zuo J, Park E-J, Liu J, Wu M, Choi J-I, Tong H (2020) Gamma-irradiation degraded sulfated polysaccharide from a new red algal strain Pyropia yezoensis Sookwawon 124 with in vitro antiproliferative activity. Oncol Lett 20:91

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  217. He D, Wu S, Yan L, Zuo J, Cheng Y, Wang H, Liu J, Zhang X, Wu M, Choi J-I, Tong H (2019) Antitumor bioactivity of porphyran extracted from Pyropia yezoensis Chonsoo2 on human cancer cell lines. J Sci Food Agric 99:6722–6730

    Article  CAS  PubMed  Google Scholar 

  218. Eduardas C, Begun MA, Tiasto VA, Belousov AS, Vikhareva VV, Mikhailova VA, Kalitnik AA (2020) In vitro antitumor and immunotropic activity of carrageenans from red algae Chondrus armatus and their low-moleclular weight degradation products. J Biomed Mater Res A 108:254–266

    Article  CAS  Google Scholar 

  219. Calvo GH, Cosenza VA, Sáenz DA, Navarro DA, Stortz CA, Céspedes MA, Mamone LA, Casas AG, Di Venosa GM (2019) Disaccharides obtained from carrageenans as potential antitumor agents. Sci Rep 9:6654

    Article  PubMed  PubMed Central  Google Scholar 

  220. Ghannam A, Murad H, Jazzara M, Odeh A, Allaf AW (2018) Isolation, structural characterization, and antiproliferative activity of phycocolloids from the red seaweed Laurencia papillosa on MCF-7 human breast cancer cells. Int J Biol Macromol 108:916–926

    Article  CAS  PubMed  Google Scholar 

  221. Martins RM, Nedel F, Guimarães VBS, da Silva AF, Colepicolo P, de Pereira CMP, Lund RG (2018) Macroalgae extracts from Antarctica have antimicrobial and anticancer potential. Front Microbiol 9:412

    Article  PubMed  PubMed Central  Google Scholar 

  222. Anand J, Sathuvan M, Babu GV, Sakthivel M, Palani P, Nagaraj S (2018) Bioactive potential and composition analysis of sulfated polysaccharide from Acanthophora spicifera (Vahl) Borgeson. Int J Biol Macromol 111:1238–1244

    Article  CAS  PubMed  Google Scholar 

  223. Alves C, Pinteus S, Rodrigues A, Horta A, Pedrosa R (2018) Algae from Portuguese Coast presented high cytotoxicity and antiproliferative effects on an in vitro model of human colorectal cancer. Pharmacognosy Res 10:24–30

    CAS  PubMed  PubMed Central  Google Scholar 

  224. Sae-lao T, Luplertlop N, Janvilisri T, Tohtong R, Bates DO, Wongprasert K (2017) Sulfated galactans from the red seaweed Gracilaria fisheri exerts anti-migration effect on cholangiocarcinoma cells. Phytomedicine 36:59–67

    Article  CAS  PubMed  Google Scholar 

  225. Tarhouni-Jabberi S, Zakraoui O, Ioannou E, Riahi-Chebbi I, Haoues M, Roussis V, Kharrat R, Essafi-Benkhadir K (2017) Mertensene, a halogenated monoterpene, induces G2/M cell cycle arrest and caspase dependent apoptosis of human colon adenocarcinoma HT29 cell line through the modulation of ERK-1/-2, AKT and NF-Κb signaling. Mar Drugs 15:221

    Article  PubMed  PubMed Central  Google Scholar 

  226. Poupard N, Badarou P, Fasani F, Groult H, Bridiau N, Sannier F, Bordenave-Juchereau S, Kieda C, Piot J-M, Grillon C, Fruitier-Arnaudin I, Maugard T (2017) Assessment of heparanase-mediated angiogenesis using microvascular endothelial cells: Identification of λ-carrageenan derivative as a potent anti angiogenic agent. Mar Drugs 15:134

    Article  PubMed  PubMed Central  Google Scholar 

  227. Kang Y, Wang ZJ, Xie D, Sun X, Yang W, Zhao X, Xu N (2017) Characterization and potential antitumor activity of polysaccharide from Gracilariopsis lemaneiformis. Mar Drugs 15:100

    Article  PubMed  PubMed Central  Google Scholar 

  228. Suganya AM, Sanjivkumar M, Chandran MN, Palavesam A, Immanuel G (2016) Pharmacological importance of sulphated polysaccharide carrageenan from red seaweed Kappaphycus alvarezii in comparison with commercial carrageenan. Biomed Pharmacother 84:1300–1312

    Article  CAS  PubMed  Google Scholar 

  229. Murad H, Hawat M, Ekhtiar A, AlJapawe A, Abbas A, Darwish H, Sbenati O, Ghannam A (2016) Induction of G1-phase cell cycle arrest and apoptosis pathway in MDA-MB-231 human breast cancer cells by sulfated polysaccharide extracted from Laurencia papillosa. Cancer Cell Int 16:39

    Article  PubMed  PubMed Central  Google Scholar 

  230. Alves C, Pinteus S, Horta A, Pedrosa R (2016) High cytotoxicity and anti-proliferative activity of algae extracts on an in vitro model of human hepatocellular carcinoma. Springerplus 5:1339

    Article  PubMed  PubMed Central  Google Scholar 

  231. Ҫelenk FG, Ӧzkaya AB, Sukatar A (2016) Macroalgae of Izmir Gulf: Dictyotaceae exhibit high in vitro anti-cancer activity independent from their antioxidant capabilities. Cytotechnology 68:2667–2676

    Article  Google Scholar 

  232. Gao Y, Li Y, Niu Y, Ju H, Chen R, Li B, Song X, Song L (2021) Chemical characterization, antitumor, and immune-enhancing activities of polysaccharide from Sargassum pallidum. Molecules 26:7559

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  233. Fernando IPS, Sanjeewa KKA, Lee HG, Kim H-S, Vaas APJP, De Silva HIC, Nanayakkara CM, Abeytunga DTU, Lee D-S, Lee J-S, Jeon Y-J (2020) Fucoidan purified from Sargassum polycystum induces apoptosis through mitochondria-mediated pathway in HL-60 and MCF-7 cells. Mar Drugs 18:196

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  234. Song J-H, Won SY, Hwang B, Jung S, Choi C, Park S-S, Choi YH, Kim W-J, Moon S-K (2020) In vitro and in vivo antitumor efficacy of Hizikia fusiforme celluclast extract against bladder cancer. Nutrients 12:2159

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  235. Abu-Khudir R, Ismail GA, Diab T (2021) Antimicrobial, antioxidant, and anti-tumor activities of Sargassum linearifolium and Cystoseira crinite from Egyptian Meditterranean Coast. Nutr Cancer 73:829–844

    Article  CAS  PubMed  Google Scholar 

  236. Yan M-D, Lin H-Y, Hwang P-A (2019) The anti-tumor activity of brown seaweed oligo-fucoidan via IncRNA expression modulation in HepG2 cells. Cytotechnology 71:363–374

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  237. Lahrsen E, Liewert I, Alban S (2018) Gradual degradation of fucoidan from Fucus vesiculosus and its effect on structure, antioxidant and antiproliferative activities. Carbohydr Polym 192:208–216

    Article  CAS  PubMed  Google Scholar 

  238. Fan S, Yu G, Nie W, Jin J, Chen L, Chen X (2018) Antitumor activity and underlying mechanism of Sargassum fusiforme polysaccharides in CNE-bearing mice. Int J Biol Macromol 112:516–522

    Article  CAS  PubMed  Google Scholar 

  239. Fan S, Zhang J, Nie W, Zhou W, Jin L, Chen X, Lu J (2017) Antitumor effects of polysaccharide from Sargassum fusiforme against human hepatocellular carcinoma HepG2 cells. Food Chem Toxicol 102:53–62

    Article  CAS  PubMed  Google Scholar 

  240. Alwarsamy M, Gooneratne R, Ravichandran R (2016) Effect of fucoidan from Turbinaria conoides on human lung adenocarcinoma epithelial (A549) cells. Carbohydr Polym 152:207–213

    Article  CAS  PubMed  Google Scholar 

  241. Liu Y, Zheng J, Zhang Y, Wang Z, Yang Y, Bai M, Dai Y (2016) Fucoxanthin activates apoptosis via inhibition of PI3K/Akt/mTOR pathway and suppresses invasion and migration by restriction of p38-MMP-2/9 pathway in human glioblastoma cells. Neurochem Res 41:2728–2751

    Article  CAS  PubMed  Google Scholar 

  242. Yang G, Zhang Q, Kong Y, Xie B, Gao M, Tao Y, Xu H, Zhan F, Dai B, Shi J, Wu X (2015) Antitumor activity of fucoidan against diffuse large B cell lymphoma in vitro and in vivo. Acta Biochim Biophys Sin 47:925–931

    Article  CAS  PubMed  Google Scholar 

  243. Gomes DL, Telles CBS, Costa MSSP, Almeida-Lima J, Costa LS, Keesen TSL, Rocha HAO (2015) Methanolic extracts from brown seaweeds Dictyota cilliolata and Dictyota menstrualis induce apoptosis in human cervical adenocarcinoma HeLa cells. Molecules 20:6573–6591

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  244. Abu R, Jiang Z, Ueno M, Isaka S, Nakazono S, Okimura T, Cho K, Yamaguchi K, Kim D, Oda T (2015) Anti-metastatic effects of the sulfated polysaccharide ascophyllan isolated from Ascophyllum nodosum on B16 melanoma. Biochem Biophys Res Commun 458:727–732

    Article  CAS  PubMed  Google Scholar 

  245. Mondal AS, Sharma R, Das A, Warghane A (2022) In silico molecular characterization of Helicobacter pylori based on tandem repeat number and 16S rRNA gene. EJMA. https://doi.org/10.14744/ejma.2022.76476

  246. Ma HJ, Wang JL (2013) Quadruple therapy for eradication of Helicobacter pylori. World J Gastroenterol 19:931–935

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  247. Rehman A (2018) Inside Helicobacter pylori: a serious threat to humans. Avicenna J Clin Microbiol Infect 5:60463

    Article  Google Scholar 

  248. Gutiérrez-Zamorano C, González-Ávila M, Díaz-Blas G, Smith CT, González-Correa C, García-Cancino A (2019) Increased anti-Helicobacter pylori effect of the probiotic Lactobacillus fermentum UCO-979C strain encapsulated in carrageenan evaluated in gastric simulations under fasting conditions. Food Res Int 121:812–816

    Article  PubMed  Google Scholar 

  249. Díaz P, Valenzuela Valderrama M, Bravo J, Quest AFG (2018) Helicobacter pylori and Gastric Cancer: adaptive cellular mechanisms involved in disease progression. Front Microbiol 9:5

    Article  PubMed  PubMed Central  Google Scholar 

  250. Alzahrani S, Lina TT, Gonzalez J, Pinchuk IV, Beswick EJ, Reyes VE (2014) Effect of Helicobacter pylori on gastric epithelial cells. World J Gastroenterol 20:12767–12780

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  251. Loke MF, Lui SY, Ng BL, Gong M, Ho B (2007) Antiadhesive property of microalgal polysaccharide extract on the binding of Helicobacter pylori to gastric mucin. FEMS Immunol Med Microbiol 50:231–238

    Article  CAS  PubMed  Google Scholar 

  252. Ayala G, Escobedo-Hinojosa WI, Cruz-Herrera CFL, Romero I (2014) Exploring alternative treatments for Helicobacter pylori infection. World J Gastroenterol 20:1450–1469

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  253. MubarakAli D, Akshaya T, Sathya R, Irfan N (2022) Study on the interaction of algal peptides on virulence factors of Helicobacter pylori: In silico approach. Appl Biochem Biotechnol 194:37–53

    Article  CAS  PubMed  Google Scholar 

  254. Besednova NN, Zaporozhets TS, Somova LM, Kuznetsova TA (2015) Review: Prospects for the use of extracts and polysaccharides from marine algae to prevent and treat the diseases caused by Helicobacter pylori. Helicobacter 20:89–97

    Article  CAS  PubMed  Google Scholar 

  255. Hetz C (2012) The unfolded protein response: controlling cell fate decisions under ER stress and beyond. Nat Rev Mol Cell Biol 13:89–102

    Article  CAS  PubMed  Google Scholar 

  256. Akazawa Y, Isomoto H, Matsushima K, Kanda T, Minami H, Yamaghchi N, Taura N, Shiozawa K, Ohnita K, Takeshima F, Nakano M, Moss J, Hirayama T, Nakao K (2013) Endoplasmic reticulum stress contributes to Helicobacter pylori VacA-induced apoptosis. PLoS ONE 8:e82322

    Article  PubMed  PubMed Central  Google Scholar 

  257. Li N, Tang B, Jia YP, Zhu P, Zhuang Y, Fang Y, Li Q, Wang K, Zhang W-J, Guo G, Wang T-J, Feng Y-J, Qiao B, Mao X-H, Zou Q-M (2017) Helicobacter pylori CagA protein negatively regulates autophagy and promotes inflammatory response via c-Met-PI3K/Akt-mTOR signaling pathway. Front Cell Infect Microbiol 7:417

    Article  PubMed  PubMed Central  Google Scholar 

  258. Wei J, Gou Z, Wen Y, Luo Q, Huang Z (2020) Marine compounds targeting the PI3K/Akt signaling pathway in cancer therapy. Biomed Pharmacother 129:110484

    Article  CAS  PubMed  Google Scholar 

  259. de Azevedo JWV, de Medeiros Fernandes TAA, Fernandes JV Jr, de Azevedo JCV, Lanza DCF, Bezerra CM, Andrade VS, de Araújo JMG, Fernandes JV (2020) Biology and pathogenesis of human osteosarcoma. Oncol Lett 19:1099–1116

    PubMed  Google Scholar 

  260. Xie P, Fujii I, Zhao J, Shinohara M, Matsukura M (2016) A novel polysaccharide derived from algae extract induces apoptosis and cell cycle arrest in human gastric carcinoma MKN45 cells via ROS/JNK signaling pathway. Int J Oncol 49:1561–1568

    Article  CAS  PubMed  Google Scholar 

  261. Xie P, Horio F, Fujii I, Zhao J, Shinohara M, Matsukura M (2018) A novel polysaccharide derived from algae extract inhibits cancer progression via JNK, not via the p38 MAPK signaling pathway. Int J Oncol 52:1380–1390

    CAS  PubMed  PubMed Central  Google Scholar 

  262. Kim S-K, Karagozlu MZ (2011) Marine algae: Natural product source for gastrointestinal cancer treatment. In: Kim SK (ed) Advances in food and nutrition research. Elsevier, pp 225–233

    Google Scholar 

  263. Chua EG, Verbrugghe P, Perkins TT, Tay CY (2015) Fucoidans disrupt adherence of Helicobacter pylori to AGS cells in vitro. Evid Based Complement Altern Med 2015:120981

    Article  Google Scholar 

  264. Zhang X-Y, Zhang P-Y, Aboul-Soud MAM (2017) From inflammation to gastric cancer: Role of Helicobacter pylori. Oncol Lett 13:543–548

    Article  CAS  PubMed  Google Scholar 

  265. Liu BH, Lee YK (2003) Effect of total secondary carotenoids extracts from Chlorococcum sp on Helicobacter pylori-infected BALB/c mice. Int Immunopharmacol 3:979–986

    Article  CAS  PubMed  Google Scholar 

  266. Shih CM, Cheng SN, Wong CS, Kuo YL, Chou TC (2009) Antiinflammatory and antihyperalgesic activity of C-phycocyanin. Anesth Analg 108:1303–1310

    Article  CAS  PubMed  Google Scholar 

  267. Hamidi M, Kozani PS, Kozani PS, Pierre G, Michaud P, Delattre C (2020) Marine bacteria versus microalgae: Who is the best for biotechnological production of bioactive compounds with antioxidant properties and other biological applications? Mar Drugs 18:28

    Article  CAS  Google Scholar 

  268. Zhong D, Du Z, Zhou M (2021) Algae: a natural active material for biomedical applications. View 2:20200189

    Article  CAS  Google Scholar 

  269. Zhong D, Zhang D, Chen W, He J, Ren C, Zhang X, Kong N, Tao W, Zhou M (2021) Orally deliverable strategy based on microalgal biomass for intestinal disease treatment. Sci Adv. https://doi.org/10.1126/sciadv.abi9265

    Article  PubMed  PubMed Central  Google Scholar 

  270. Akolpoglu MB, Dogan NO, Bozuyuk U, Ceylan H, Kizilel S, Sitti M (2020) High-yield production of biohybrid microalgae for on-demand cargo delivery. Adv Sci (Weinh) 7:2001256

    Article  CAS  PubMed  Google Scholar 

  271. Wang X, Cai J, Sun L, Zhang S, Gong D, Li X, Yue S, Feng L, Zhang D (2019) Facile fabrication of magnetic microrobots based on Spirulina templates for targeted delivery and synergistic chemo-photothermal therapy. ACS Appl Mater Interfaces 11:4745–4756

    Article  CAS  PubMed  Google Scholar 

  272. Delalat B, Sheppard VC, Ghaemi SR, Rao S, Prestidge CA, McPhee G, Rogers ML, Donoghue JF, Pillay V, Johns TG, Kröger N, Voelcker NH (2015) Targeted drug delivery using genetically engineered diatom biosilica. Nat Commun 6:8791

    Article  CAS  PubMed  Google Scholar 

  273. Zhong D, Zhang D, Xie T, Zhou M (2020) Biodegradable microalgae-based carriers for targeted delivery and imaging-guided therapy toward lung metastasis of breast cancer. Small 16:2000819

    Article  CAS  Google Scholar 

  274. Muz B, de la Puente P, Azab F, Azab AK (2015) The role of hypoxia in cancer progression, angiogenesis, metastasis, and resistance to therapy. Hypoxia (Auckl) 3:83–92

    Article  PubMed  Google Scholar 

  275. Zhong D, Li W, Qi Y, He J, Zhou M (2020) Photosynthetic biohybrid nanoswimmers system to alleviate tumor hypoxia for FL/PA/MR imaging-guided enhanced radio-photodynamic synergetic therapy. Adv Funct Mater 30:1910395

    Article  CAS  Google Scholar 

  276. Li W, Zhong D, Hua S, Du Z, Zhou M (2020) Biomineralized biohybrid algae for tumor hypoxia modulation and cascade radio-photodynamic therapy. ACS Appl Mater Interfaces 12:44541–44553

    Article  CAS  PubMed  Google Scholar 

  277. Qiao Y, Yang F, Xie T, Du Z, Zhong D, Qi Y, Li Y, Li W, Lu Z, Rao J, Sun Y, Zhou M (2020) Engineered algae: A novel oxygen-generating system for effective treatment of hypoxic cancer. Sci Adv 6:eaba5996

  278. Teas J, Vena S, Cone DL, Irhimeh M (2013) The consumption of seaweed as a protective factor in the etiology of breast cancer: proof of principle. J Appl Phycol 25:771–779

    Article  CAS  PubMed  Google Scholar 

  279. Song M, Wu K, Meyerhardt JA, Ogino S, Wang M, Fuchs CS, Giovannucci EL, Chan AT (2018) Fiber intake and survival after colorectal cancer diagnosis. JAMA Oncol 4:71–79

    Article  PubMed  PubMed Central  Google Scholar 

  280. Minami Y, Kanemura S, Oikawa T, Suzuki S, Hasegawa Y, Nishino Y, Fujiya T, Miura K (2020) Associations of Japanese food intake with survival of stomach and colorectal cancer: a prospect patient cohort study. Cancer Sci 111:2558–2569

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  281. Yang YJ, Nam SJ, Kong G, Kim MK (2010) A case-control study on seaweed consumption and the risk of breast cancer. Br J Nutr 103:1345–1353

    Article  CAS  PubMed  Google Scholar 

  282. Song K, Xu L, Zhang W, Cai Y, Jang B, Oh J, Jin JO (2017) Laminarin promotes anti-cancer immunity by the maturation of dendritic cells. Oncotarget 8:38554–38567

    Article  PubMed  PubMed Central  Google Scholar 

  283. Sun J, Sun J, Song B, Zhang L, Shao Q, Liu Y, Yuan D, Zhang Y, Qu X (2016) Fucoidan inhibits CCL2 production through NF-κB pathway in M2 macrophages: a potential therapeutic strategy for cancer. Sci Rep 6:35855

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  284. Lai YH, Chiang CS, Hsu CH, Cheng HW, Chen SY (2020) Development and characterization of a fucoidan-based drug delivery system by using hydrophilic anticancer polysaccharides to simultaneously deliver hydrophobic anticancer drugs. Biomolecules 10:970

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  285. Youssouf L, Bhaw-Luximon A, Diotel N, Catan A, Giraud P, Gimié F, Lallemand L (2019) Enhanced effects of curcumin encapsulated in polycaprolactone-grafted oligocarrageenan nanomicelle, a novel nanoparticle drug delivery system. Carbohydr Polym 217:35–45

    Article  CAS  PubMed  Google Scholar 

  286. Kahya N, Golcu A, Erim FB (2019) Barium ion cross-linked alginate-carboxymethyl cellulose composites for controlled release of anticancer drug methotrexate. J Drug Deliv Sci Technol 54:101324

    Article  CAS  Google Scholar 

  287. Khan H, Chaudhary JP, Meena R (2019) Anionic carboxymethylagarose-based-pH-responsive smart superabsorbent hydrogels for controlled release of anticancer drugs. Int J Biol Macromol 124:1220–1229

    Article  CAS  PubMed  Google Scholar 

  288. Prabhu R, Ashik Mohammed M, Anjali R, Archunan G, Prabhu NM, Pugazhendhi A, Suganthy N (2019) Ecofriendly one pot fabrication of methyl gallate@ZIF-L nanoscale hybrid as pH responsive drug delivery system for lung cancer therapy. Process Biochem 84:39–52

    Article  Google Scholar 

  289. Moon C, Kim SH, Kim JC, Hyun JW, Lee NH, Park JW, Shin T (2008) Protective effect of phlorotannin components phloroglucinol and eckol on radiation-induced intestinal injury in mice. Phytother Res 22:238–242

    Article  CAS  PubMed  Google Scholar 

  290. Zhang R, Kang KA, Piao MJ, Ko DO, Wang ZH, Lee IK, Kim BJ, Jeong IY, Shin T, Park JW, Lee NH, Hyun JW (2008) Eckol protects V79–4 lung fibroblast cells against gamma-ray radiation-induced apoptosis via the scavenging of reactive oxygen species and inhibiting of the c-Jun NH(2)-terminal kinase pathway. Eur J Pharmacol 591:114–123

    Article  CAS  PubMed  Google Scholar 

  291. Park E, Ahn G, Yun JS, Kim MJ, Bing SJ, Kim DS (2010) Dieckol rescues mice from lethal irradiation by accelerating homopoiesis and curtailing immunosuppression. Int J Radiat Biol 86:848–859

    CAS  PubMed  Google Scholar 

  292. Jeong J, Yang W, Ahn M, Kim KC, Hyun JW, Kim SH, Moon C, Shin T (2011) Protective effect of the methanolic extract of Polyopes lancifolia (Harvey) kawaguchi et wang against ionizing radiation-induced mouse gastrointestinal injury. Korean J Vet Res 51:177–183

    Article  Google Scholar 

Download references

Acknowledgements

The authors would like to acknowledge the Department of Science and Technology (DST), New Delhi, India, for the DST INSPIRE Faculty grant.

Funding

NT received DST INSPIRE faculty funds from the Department of Science and Technology, New Delhi.

Author information

Authors and Affiliations

Authors

Contributions

RS, ASM, and NT have equally contributed to this review article. The microalgae and macroalgae sections are contributed by RS and ASM. NT has conceptualized the structure of the MS and critically reviewed it. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Nitin Trivedi.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

The authors declare no conflict of interest.

Competing interests

The authors have no competing interests to declare relevant to this article’s content.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Sharma, R., Mondal, A.S. & Trivedi, N. Anticancer potential of algae-derived metabolites: recent updates and breakthroughs. Futur J Pharm Sci 9, 44 (2023). https://doi.org/10.1186/s43094-023-00492-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s43094-023-00492-2

Keywords