Next Article in Journal
Antiviral Activity and Mechanisms of Seaweeds Bioactive Compounds on Enveloped Viruses—A Review
Previous Article in Journal
Marine Chitooligosaccharide Alters Intestinal Flora Structure and Regulates Hepatic Inflammatory Response to Influence Nonalcoholic Fatty Liver Disease
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Seaweed-Derived Phlorotannins: A Review of Multiple Biological Roles and Action Mechanisms

by
Fazlurrahman Khan
1,*,†,
Geum-Jae Jeong
2,†,
Mohd Sajjad Ahmad Khan
3,
Nazia Tabassum
4 and
Young-Mog Kim
1,2,*
1
Research Center for Marine Integrated Bionics Technology, Pukyong National University, Busan 48513, Korea
2
Department of Food Science and Technology, Pukyong National University, Busan 48513, Korea
3
Department of Basic Sciences, Deanship of Preparatory Year and Supporting Studies, Imam Abdulrahman Bin Faisal University, P.O. Box 1982, Dammam 34212, Saudi Arabia
4
Industry 4.0 Convergence Bionics Engineering, Pukyong National University, Busan 48513, Korea
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Mar. Drugs 2022, 20(6), 384; https://doi.org/10.3390/md20060384
Submission received: 11 May 2022 / Revised: 1 June 2022 / Accepted: 4 June 2022 / Published: 8 June 2022

Abstract

:
Phlorotannins are a group of phenolic secondary metabolites isolated from a variety of brown algal species belonging to the Fucaceae, Sargassaceae, and Alariaceae families. The isolation of phlorotannins from various algal species has received a lot of interest owing to the fact that they have a range of biological features and are very biocompatible in their applications. Phlorotannins have a wide range of therapeutic biological actions, including antimicrobial, antidiabetic, antioxidant, anticancer, anti-inflammatory, anti-adipogenesis, and numerous other biomedical applications. The current review has extensively addressed the application of phlorotannins, which have been extensively investigated for the above-mentioned biological action and the underlying mechanism of action. Furthermore, the current review offers many ways to use phlorotannins to avoid certain downsides, such as low stability. This review article will assist the scientific community in investigating the greater biological significance of phlorotannins and developing innovative techniques for treating both infectious and non-infectious diseases in humans.

1. Introduction

Marine organisms, particularly brown seaweeds, have garnered a lot of interest across the world for their potential use in treating a variety of infectious and non-infectious diseases [1,2]. Several compounds have been isolated from brown seaweeds, including phlorotannins, fucoxanthin, fucoidan, alginate, fucoidan, and laminarin [3]. Almost all secondary metabolites have been found to possess various biological activities. Among these, phlorotannins, which are members of the polyphenolic group of secondary metabolites, has been found in abundance in brown algae belonging to the Fucaceae, Sargassaceae, and Alariaceae families [4]. Some red and green algae also contain phlorotannins, albeit to a lesser extent than brown algae [5]. Despite problems in isolating and purifying phlorotannins from brown seaweeds, more than 150 phlorotannins have been identified [6]. The different sources from which phlorotannins have been isolated and identified are given in Table 1 and Table 2. Figure 1 depicts the chemical structures of these phlorotannins. The use of traditional antibiotics and other pharmaceuticals to treat infectious diseases caused by bacterial, viral, and fungal pathogens [7,8,9], as well as non-infectious diseases such as cancer, diabetes, inflammation, obesity, and adipogenesis, has been shown to be detrimental [10,11,12]. Microbial infections caused by viruses, bacteria, and fungus have become one of the most serious threats to human health across the world [13,14,15].
Furthermore, microbial pathogens such as bacteria and fungi evolved many resistance mechanisms against commonly used antibiotics and antifungals [16,17]. Multi-drug resistant strains of microorganisms from various hospital and clinical environments are also rapidly growing [18,19,20]. With rising antimicrobial resistance, the characterization of microbial pathogens and innovative strategies to combat microbial diseases are in great demand. Recent and previous research has found that phlorotannins have antibacterial, antifungal, and antiviral effects against a wide variety of microbial pathogens [21,22,23]. Non-infectious disorders, including cancer, diabetes, inflammation, and obesity, are also on the rise and creating a strong need for newer therapeutic drugs [24,25,26]. The current review addresses the biological properties of several types of phlorotannins in treating both infectious and non-infectious diseases. Furthermore, the underlying mechanism involved in treating diseases using phlorotannins have also been explored. Additionally, the current limitations and future perspectives associated with the isolation and application of phlorotannins have been thoroughly discussed.

2. Phlorotannin against Pathogenic Bacterial and Fungal Species

Antibiotics are used to treat the majority of bacterial pathogens, but the development of antibiotic resistance in pathogens has recently grown substantially, directly affecting human mortality. Antibiotic resistance is mostly caused by the indiscriminate use of antibiotics for pathogen treatment, long-term antibiotic therapy for patients, and patient discontinuation of antibiotics before finishing treatment [27]. According to reports, Acinetobacter baumannii, Enterococcus faecium, Enterobacter spp., Klebsiella pneumoniae, Pseudomonas aeruginosa, and Staphylococcus spp. are the major bacterial pathogens involved in the development of antibiotic resistance [28]. Antibiotic resistance can be developed or acquired through processes such as (1) lower intracellular concentration of antibiotics owing to pathogen efflux, (2) genetic change of the antibiotic target, and (3) antibiotic deactivation by hydrolysis or modification [29]. This is comparable to the situation in fungal pathogens, where studies have shown that antifungal drugs are successful in the treatment of common fungal pathogens such as Candida spp., Cryptococcus spp., and Aspergillus spp. However, fungi develop resistance as a result of the indiscriminate use of antifungal drugs, resulting in increased fungal disease mortality [30]. Furthermore, the majority of antifungal agents utilized in the treatment of human diseases include polyenes, triazoles, and echinocandins, but they are toxic to patients [31]. Aside from intrinsic and acquired resistance mechanisms, bacterial and fungal species have evolved phenotypic adaptive resistance mechanisms known as biofilm formation, which is an assembly of cells surrounded by self-produced polymeric substances [32]. Biofilms act as a structural and functional barrier to the entry of antibacterial and antifungal drugs into the microbial cell. Furthermore, these biofilms contained a small population of cells (persister cells) that are metabolically inactive and have been linked to persistent and recurring infection. As a result, there is an urgent need to develop a strategy or find novel molecules to battle antibiotic-resistant bacteria and fungi. Several studies have proven the antibacterial action of phlorotannins isolated from various seaweeds. Phlorotannins have an antibacterial effect by suppressing oxidative phosphorylation, altering target microorganism cell permeability, and triggering death by interacting with target microbial proteins [33]. The phenolic aromatic ring and hydroxyl groups of various phlorotannins have been shown to bind to microbial proteins and induce cell lysis [34]. The susceptibility of Gram-positive and Gram-negative bacteria to phlorotannins varies owing to structural differences in the cell membrane and cell wall. Gram-negative bacteria are often less susceptible to phlorotannin than Gram-positive bacteria [35]. Nagayama et al. [21] exhibited the bactericidal action of phlorotannins derived from Ecklonia kurome for the first time, where phlorotannins showed excellent activity against MRSA. Several types of phlorotannins isolated from Eisenia bicyclis showed antibacterial activity against Listeria monocytogenes at 16–256 µg/mL [36]. Furthermore, the combination of phlorotannins and streptomycin demonstrated strong synergistic activity against aminoglycoside antibiotic-resistant clinical strains of L. monocytogenes (fractional inhibitory concentration index, ∑FIC < 0.5). Cox et al. [37] discovered that extracts prepared from several seaweeds inhibited E. faecalis, L. monocytogenes, P. aeruginosa, and Salmonella abony. Brown seaweed, for instance, outperformed red and green seaweed in antibacterial activity. Another research group found that phlorotannins inhibited quorum sensing (QS) activity in Chromobacterium violaceum 12472 by preventing the synthesis of purple pigment [38]. Similarly, phlorotannins reduced P. aeruginosa virulence by inhibiting motility and decreasing the production of virulence factors such as extracellular protease, hemolysis, and pyocyanin. Eom et al. [39] conducted the first investigation on the anti-MRSA (methicillin-resistant S. aureus) activity of phlorotannins from E. bicyclis. The anti-MRSA activity of phlorofucofuroeckol-A was associated with a MIC value of 32 µg/mL. Furthermore, the combinations of phlorofucofuroeckol-A and β-lactam antibiotics such as ampicillin, penicillin, and oxacillin have shown strong synergistic effects. Choi et al. [40] discovered phlorotannin activity against the acne-causing bacteria Propionibacterium acnes. Dieckol and phlorofucofuroeckol-A, in particular, were found to be very active against acne-causing bacteria when isolated from E. cava. Lee et al. [41] also discovered that the compounds isolated from E. bicyclis have strong antibacterial action against acne-causing bacteria. Furthermore, fucofuroeckol-A combined with erythromycin and lincomycin had a synergistic impact on P. acnes. Ethanolic extracts of E. cava were found to have effective antibacterial activity towards S. aureus, MRSA, and Salmonella spp. [42]. Eckol from E. cava substantially reduced the MIC value of ampicillin against MRSA and Salmonella spp. Mittal et al. [43] discovered that phloroglucinol and related compounds exhibited antibacterial action against various Gram-positive bacteria. Furthermore, the combination of phlorodipropanophenone and oxycycline had a strong synergistic bactericidal effect. Kim et al. assessed the antibacterial activity of E. cava phlorotannin against fish infectious diseases caused by marine bacterial pathogens [44]. The methanol extract of E. cava was shown to have potent antibacterial action against several pathogens such as Streptococcus iniae, Edwardsiella tarda, S. parauberis, V. harveyi, V. scophthalm, and Vibrio anguillarum. Phlorotannin isolated from E. arborea has shown antibacterial action against V. parahaemolyticus, indicating its potential as an antibiotic to combat acute hepatopancreatic necrosis disease caused by this pathogen [45]. Furthermore, dieckol derived from E. stolonifera has shown antibacterial efficacy against MRSA and methicillin-susceptible S. aureus (MSSA) [46]. In addition, the application of dieckol reversed the resistance properties developed against ampicillin and penicillin by MRSA. Wei et al. [47] studied the mechanism of phlorotannins derived from Sargassum thunbergii against V. parahaemolyticus. Their findings revealed that phlorotannin was able to suppress bacterial growth and damaged the cytoplasmic membrane. Some studies have also found that phlorotannins damage the cell membranes of anaerobic bacteria, resulting in extracellular and intracellular repercussions [48]. According to Lopes et al. [22] phlorotannin isolated from seaweed has an inhibitory effect against yeast and dermatophytes. Dieckol extracted from E. cava showed fungicidal activity towards Trichophyton rubrum (200 μM) [49], resulting in the loss of cytoplasmic membrane integrity. According to Kim et al. [50] fucofuroeckol-A has significant antifungal action towards C. albicans with the MIC value of 512 μg/mL. Fucofuroeckol-A enhanced fluconazole’s antifungal activity against fluconazole-resistant C. albicans. Figure 2 depicts a diagrammatic explanation of the mechanism involved in the antimicrobial activity of several forms of phlorotannins against bacterial and fungal pathogens. Table 1 shows several other examples of phlorotannins with antibacterial and antifungal activities. Table 1 summarizes a number of additional phlorotannins having potent antimicrobial activity.

3. Phlorotannins as a Natural Adjuvant of Antiviral Therapy

Viral diseases are one of the major issues caused by several viruses, including the human immunodeficiency virus [51], hepatitis B and C virus, and coronavirus (COVID-19). All of these have a severe negative impact on human life. Antiviral agents exhibit a variety of action mechanisms, including inhibition of viral fusion and mRNA entry, antagonism of chemokine co-receptors and cyclophilins, and inhibition of proteases and reverse transcriptase within the host cell (Figure 3). Viruses are acellular and have a simple structure with a protein coat, nucleic acid, viral enzyme, and a lipid envelope, where they replicate after infecting host cells [52]. The viral life cycle inside host cells is extremely complicated, requiring several phases from invasion through maturity. As a result, one of the key problems in designing antiviral agents with selective toxicity is the lack of these characteristics [53]. Although various antiviral medications have been identified, the existing antiviral agents have significant drawbacks, such as the development of resistance, limited bioavailability, and high toxicity [54]. Besednova et al. [55] demonstrated that compounds with antiviral activities obtained from marine species do not cause resistance owing to a variety of mechanisms, including antiviral, immunostimulatory, anti-inflammatory, and antioxidant capabilities. One of the compounds obtained from marine organisms, phlorotannin, has high antioxidant capabilities, particularly due to the presence of the phenol group. These properties have been shown to prevent viral infection as well as interfere with virus adherence, penetration, and replication [56]. According to Artan et al. [57], 6,6′-bieckol derived from E. cava prevented HIV-1 induced fusion formation (EC50 1.72 μM), viral p24 antigen production (EC50 1.26 μM), and lytic effect (EC50 1.23 μM), thereby exhibiting antiviral activities. Moreover, at concentrations that suppressed HIV-1 replication, 6,6′-bieckol did not exhibit cytotoxicity. At non-cytotoxic doses, 8, 4′′′-dieckol derived from E. cava also inhibited HIV-1 lytic effect and viral p24 antigen production [57]. According to the findings, 8,4′′′-dieckol preferentially inhibited the reverse transcriptase activity of HIV-1 (Figure 3). Ahn et al. [58] discovered that 8,4”-dieckol and 8,8′-bieckol inhibited HIV-1 protease and reverse transcriptase. Furthermore, these drugs inhibited HIV-1 reverse transcriptase’s RNA-dependent DNA synthesis activity against dUTP/dTTP in a non-competitive manner. In a cell-free-based system, phlorotannins from E. cava competitively suppressed SARS-CoV-2 3CLpro [23]. These compounds showed an inhibitory effect on 3CLpro hydrolysis (IC50 2.7~164.7 μM). Ryu et al. [59] showed that the ethanol extract of E. cava exhibited inhibitory activity against neuraminidase action, thereby inhibiting the influenza virus. In particular, the ethanol extract effectively inhibited neuraminidase by synergizing with oseltamivir.

4. Anti-Inflammatory and Immunomodulatory Properties of Phlorotannins

Inflammation is the body’s reaction to harmful stimuli, such as an infection or injury, which leads to the development of symptoms such as seizure, pain, fever, and swelling. The inflammatory response is an immunological reaction to exogenous and endogenous signals that serve as the host’s defensive systems against potentially harmful stimuli [32]. An inflammatory response is necessary for survival, and it plays an important role in cellular physiology; nevertheless, uncontrolled and excessive inflammation leads to a variety of chronic diseases. [60]. Inflammatory drugs include both steroidal and nonsteroidal, which are synthetic drugs used to treat inflammatory disorders that have potentially dangerous side effects and are not safe for humans [61]. In recent years, many investigations on the isolation and identification of natural anti-inflammatory drugs have been carried out. As a result, naturally occurring organisms have components that chemically resemble steroid structures and have been reported to exhibit a high potential for anti-inflammatory action [62]. Natural compounds with anti-inflammatory potential have antioxidant and radical scavenging characteristics, as well as the capacity to control the activity of pro-inflammatory enzymes such as phospholipases A2 (PLA2s), cyclooxygenase, and lipoxygenases (LOX) [37]. They also control the production of pro-inflammatory factors and genes [63]. As a result, these discoveries may open the path for the discovery of natural anti-inflammatory drugs. Fucosterol isolated from E. bicyclis reduced LPS-induced nitric oxide (NO) generation at non-toxic doses [64]. Furthermore, fucosterol inhibited tert-butyl hydroperoxide-induced reactive oxygen species (ROS) production and the expression of COX-2i and NOS. Several researchers used RAW 264.7 cells activated by lipopolysaccharide (LPS) to assess anti-inflammatory activity. Phlorotannins isolated from Fucales were found to be non-toxic at concentrations ranging from 31.25–500 μg/mL [65]. In RAW 264.7 cells activated by LPS, phloroglucinol reduced the generation of inflammatory mediators such as interleukin-6 (IL-6), interleukin-1β (IL-1β), tumor necrosis factor-α (TNF-α), and prostaglandin E2 (PGE2) in RAW 264.7 cells stimulated by LPS [66]. Furthermore, phloroglucinol inhibited the production of metalloproteinase, which is involved in chronic inflammation in HT 1080 cells. In macrophages activated by LPS, phlorofucofuroeckol-B isolated from Myagropsis myagroides blocked the NF-κB pathway via suppression of the Akt-ERK pathway (Figure 4) [67]. Gonçalves-Fernández et al. [68] confirmed the cytotoxic activity against murine cell lines by identifying fractions of low molecular size and medium polarity among fractionated phlorotannins. At a concentration of 100 µM, the fraction reduced ATDC-5 cell proliferation by 50%. In another study, phlorotannins isolated from the brown macroalga Padina tetrastromatica were shown to have strong anti-MRSA, anti-inflammatory, and antioxidant activities [69]. It was discovered, in particular, that the content and structure of phlorotannins differed depending on environmental and geographic conditions. Yu et al. [70] discovered that phlorofucofuroeckol-B isolated from E. stolonifera suppressed IL-1β, IL-6, and TNF-α (Figure 4). Furthermore, phlorofucofuroeckol-B suppressed the generation of NO and prostaglandins and downregulated the LPS-induced expression of inducible cyclooxygenase-2 and NO synthase in LPS-stimulated BV-2. The compound 6,6′-Bieckol decreased nitric oxide and prostaglandin production and suppressed LPS-induced expression of inducible cyclooxygenase-2 and NO synthase (Figure 4) [71]. Importantly, 6,6′-bieckol inhibited the expression of IL-6 and TNF-α.

5. Phlorotaninns Having Anticancer Potential

Cancer is a non-infectious disease caused by uncontrolled cell division or malignant cell growth. Normal cells are still being attacked and destroyed by abnormal cells. Death can occur if these systems are not managed [72]. Therapeutics for many forms of cancer have been developed over the years. However, because normal and cancer cells are so similar, anticancer drugs are likely to have negative effects on normal cells. Furthermore, because anticancer drugs are very toxic, cancer patients usually have side effects after therapeutic procedures [73,74]. These problems explain why research is being conducted to discover new therapeutic agents to replace existing ones. Many metabolites isolated from organisms are being studied for anticancer properties. According to studies, the majority of natural anticancer compounds have no side effects and control the growth of cancer cells [74]. As a result, one of the most significant strategies for cancer control is the quest for effective natural anticancer drugs. The anticancer mechanism of phlorotannins encompasses several pathways that are linked to the development of cancer and the triggering of cell death [75,76,77]. Phlorotannins prevent cancer via increasing cytotoxic T lymphocytes, dendritic cells, the epithelial-mesenchymal transition process, matrix metalloproteinase, phagocyte release, and decreasing SLUG and VEGF expression. Furthermore, phlorotannins modulate apoptosis induction by increasing the expression of apoptosis antigen 1 (APO-1), B-cell lymphoma 2 (Bcl-2) protein, caspase-3, -7, -9, cysteinyl aspartate specific proteinase (casp), and down-regulating the expression of the protein kinase B (AKT) pathway, B-cell lymphoma-extra-large (Bcl-xL) protein, extracellular signal-related kinase (ERK) pathway, FLICE (FADD-like IL-1β-converting enzyme)-inhibitory protein (FLIP), nuclear factor-kB (NF-κB), phosphoinositide 3-kinase (PI3K) pathway, and X-linked inhibitor of apoptosis (XIAP) (Figure 5). Dieckol extracted from E. cava was shown to be cytotoxic to A2780 and SKOV3 ovarian cancer cells, as reported by Ahn et al. [78]. Dieckol promoted apoptosis in SKOV3 cells and reduced tumor growth in SKOV3-bearing mouse models without causing severe toxicity. Furthermore, phloroglucinol from E. cava inhibited multiplication and triggered apoptosis in MCF-7 human cancer cells [79]. Phloroglucinol induced an increase in apoptosis-related gene expression while decreasing anti-apoptotic gene expression. Another study found that phloroglucinol effectively suppressed breast cancer cells by downregulating SLUG without generating cytotoxicity [80]. Kang et al. [81] discovered that phloroglucinol decreased cell viability and promoted apoptosis in HT-29 colon cancer cells. Phloroglucinol induced apoptosis-related alterations such as Bcl-2 family protein modification, cytochrome c release, and caspase-3 and -8 activation (Figure 5). Phlorotannins isolated from Cystoseira sedoides induced apoptosis in more than half of MCF-7 breast cancer cells, with the IC50 value of 78 μg/mL [82]. Furthermore, in N-nitrosodiethylamine-induced hepatocarcinogenesis, dieckol inhibited lipid peroxidation, hepatic cell damage, and increased antioxidant defense mechanisms (Figure 5) [83].

6. Antioxidant Abilities of Phlorotannins in Modulating Oxidative Stress

Oxidation is a key metabolic process in which electrons or hydrogen are transferred from a material to an oxidizing agent. Free radicals are produced during oxidative processes, which damage or destroy cells. Excess-free radical generation destroys enzymes like superoxide, dismutase, catalase, and peroxidase, resulting in a variety of human disorders, including diabetes and cancer [84,85]. The consequences of excessive oxidation highlight the need for a wide range of antioxidants obtained from natural sources. By preventing oxidation, antioxidants serve as a defensive mechanism against the consequences of excessive oxidation. These antioxidants found in the body have been linked to protection from various diseases, including cancer, aging, and Alzheimer’s [51,86]. Various in-vitro methods are applied to determine the antioxidant activity present in natural sources. Scavenging assays utilizing 2,2-diphenyl-1-picrylhydrazyl (DPPH) and 2,2′-azino-bis-3-ethylbenzthiazoline-6-sulphonic acid radical are used to evaluate the radical scavenging activity of antioxidants. The β-carotene-linoleate model systems and the thiobarbituric acid reactive substances tests are widely used to assess the antioxidant activity of foods and organic compounds containing lipids. The reducing power of antioxidants is measured using the ferric antioxidant power reduction and cupric ion reducing antioxidant capacity assays. These assays are commonly used to identify prospective antioxidant sources [87,88]. Sathya et al. [89] found that the phlorotannins from C. trinodis had significant radical-scavenging activity against superoxide and DPPH radicals and were as effective as ascorbic acid and α-tocopherol. Previously, it was attempted to determine the relationship between the number of hydroxyl groups in phlorotannins and antioxidant activity. The results showed the antioxidant activity of phlorotannins to be dependent on the structure (substitution pattern) of phlorotannins [90]. Moreover, Ahn et al. [91] demonstrated that phlorotannins isolated from E. cava have better free radical scavenging and DNA damage inhibition properties. Nakai et al. [92] conducted a study to enhance the radical scavenging activity of S. ringgoldianum. As a result, it was shown that oligomerization of phloroglucinol was required for the improvement of radical scavenging activity. A recent study found a relationship between phlorotannin content and total antioxidant activity, indicating that phlorotannins play a role in determining total antioxidant activity [93]. Lee et al. [94] discovered that phlorofucofuroeckol-A exhibited excellent antioxidant properties against AAPH-induced ROS generation and lipid peroxidation in both in vitro and in vivo systems. Similarly, it was demonstrated that phlorotannins isolated from E. cava protected neurons against H2O2-induced neurotoxicity and significantly reduced intracellular ROS production, lipid peroxidation, and apoptosis generated by AAPH [95,96].

7. Phlorotannins with Anti-Adipogenesis, Antidiabetic, and Anti-Obesity Activities

Obesity, diabetes, dyslipidemia, myocardial infarction, stroke, and hypertension are all symptoms of metabolic syndrome [97]. Metabolic syndrome is becoming more common in modern life and is regarded as a major health issue. According to Blüher [98], metabolic syndrome and obesity are closely associated. Obesity is defined by a rise in adipose tissue mass and an increase in fat cell size and number. Adipogenic transcription factors that control enzymes involved in lipid metabolism induce adipogenesis [99]. Transcription factors such as CCAAT/enhancer-binding protein (C/EBPα) and peroxisome proliferator-activated receptor γ (PPARγ), in particular, have a direct influence on adipogenesis [100,101]. Adipocyte hypertrophy and hyperplasia can be induced during adipogenesis by expressing adipogenesis-specific genes [102]. The model utilized to assess adipogenesis and adipocyte differentiation are 3T3-L1 cells [103]. Despite the fact that several anti-obesity drugs have been developed to treat obesity, current anti-obesity treatments have substantial adverse effects. Existing anti-obesity drugs have been linked to major psychological and cardiovascular problems [104]. Given the negative effects of anti-obesity drugs, there is a need for research on natural compounds having anti-obesity efficacy. Hu et al. [105] discovered that phlorotannins had an anti-obesity action by inhibiting pancreatic lipase and protein tyrosine phosphatase 1B and inducing adipocyte differentiation and apoptosis. Dioxynodehydroeckol and fucofuroeckol A isolated from E. bicyclis were shown to have substantial antidiabetic action against α-glucosidase (IC50 value of 131.34, 93.33 nmol/L) and α-amylase (IC50 value of 42.91, 472.7 µmol/L) [106]. Furthermore, fucofuroeckol A and dioxinodehydroeckol inhibited α-glucosidase in a non-competitive manner. Another investigation found that among the E. cava phlorotannins, dieckol had the greatest inhibitory effect against α-glucosidase (IC50 value of 10.8 µmol/L) and α-amylase (IC50 value of 124.9 µmol/L) [107]. Gheda et al. [108] found that providing phlorotannins from C. compressa to diabetic rats lowered their serum glucose levels, total serum cholesterol, total triglycerides, liver malondialdehyde, α-amylase, and glucosidase activity. Moreover, the histological investigation revealed that the diabetic group’s islets of Langerhans and certain necrotic regions were substantially damaged. In contrast, sections of diabetic rats treated with phlorotannin exhibited a considerable improvement and recovery in the size of the islets of Langerhans. Fucus vesiculosus phlorotannin extract inhibited pancreatic lipase, α-amylase, and notably, µ-glucosidase [109]. Ryu et al. [110] discovered that isophloroglucin A isolated from Ishige okamurae had an IC50 value of 54.97 μM in α-glucosidase inhibition. It showed the greatest σ-glucosidase inhibitory activity among the phlorotannin groups isolated from I. okamurae. The IC50 values of phlorofucofuroeckol A for α-glucosidase and α-amylase from E. cava were 19.52 and 6.34 μM, respectively [111]. The enzyme activity was inhibited by the hydroxyl group of phlorofucofuroeckol A, which is bound to it. Kang et al. [112] administered dieckol (10 and 20 mg/kg body weight) to diabetic mice for 14 days and observed that the blood sugar levels, serum insulin levels, body weight, and thiobarbituric acid reactive substrates in the dieckol-administered group were significantly decreased compared to the saline-administered group. Furthermore, the dieckol-administered group showed enhanced activity of several antioxidant enzymes e.g., catalase, glutathione peroxidase, and superoxide dismutase. Kim et al. [113] reported that dioxynodehydroeckol reduced lipid accumulation by inhibiting PPARγ, CEBP/α, and sterol regulatory element-binding protein 1 (SREBP). In addition, dioxynodehydroeckol suppressed fatty acid synthase, fatty acid-binding protein, lipoprotein lipase, acyl-CoA synthetase 1, and fatty acid transport protein. In a tyrosinase inhibitor experiment, dieckol from E. cava outperformed a commercial tyrosinase inhibitor (kojic acid) [114]. Interestingly, E. stolonifera eckol, phlorofucofuroeckol A, and phloroglucinol, inhibited lipid accumulation in 3T3-L1 cells, as well as the expression of key adipocyte marker genes such as C/EBPα and PPARγ [100]. Liu et al. [115] administered dieckol extract (50 mg/kg/day) to nonalcoholic fatty liver induced rats for 4 weeks. Significant improvement was observed in the plasma lipid profile, visceral fat index, liver index, and liver fat accumulation in the dieckol extract administration group. Ko et al. [116] reported that dieckol of E. cava exhibited an adipogenesis inhibitory activity that downregulates the expression of SREBP1, PPARγ, and C/EBPα. Figure 6 depicts the diagrammatic explanation of phlorotannin’s activity as an antidiabetic, anti-obesity, and anti-adipogenesis agent via the activation and inhibition of various enzymes and signaling molecules. Table 2 lists a number of other phlorotannins with various biological functions.

8. Limitations Related to the Phlorotannin Isolation, Purification, Characterization, and Application

Phlorotannins isolated from seaweed are compounds with high biological activity and have been found to be effective as a natural antibacterial, antiviral, antioxidant, anticancer, and anti-obesity agent. However, several constraints must be overcome in order to take full advantage of these properties. Because certain seaweeds contain heavy metals such as Cu, Hg, As, and Cd, unpolished seaweed can be lethal to humans [117,118,119]. To address this issue, a high level purification technique is required. Although there is a rising interest in bioactive compounds found in seaweed, it is difficult to validate comprehensive in-vivo processes due to a lack of bioavailability research [120,121,122]. Furthermore, in vitro investigations make it difficult to identify a compound’s genuine potential in formulation development [35]. As a result, there is a need for a study to clarify the mechanism of phlorotannins in vivo and clinically. Methanol, ethyl acetate, ethanol, and acetone are the most common solvents used for phlorotannin extraction. These solvents are not safe for animals or humans, and researchers should be aware of the maximum residual limit of solvents in phlorotannins [123,124]. High-temperature extraction of phlorotannins or microwave extraction would dissolve more compounds and increase extraction yield. The heat generated at high temperatures and the high output of microwaves may reduce antioxidants in phlorotannins, hence it is critical to investigate proper extraction conditions [125,126]. Enzyme-assisted extraction of phlorotannins is a safe and ecologically friendly approach. However, due to the cost of enzymes, lack of substrate-specific enzymes, and limitations in maintaining bioreactor conditions, using an appropriate extraction method is required [127,128]. Several experiments have been carried out to identify the structural characteristics of purified phlorotannins, but the drawback is that there is no available library for comparison with the standard [129,130]. The spectrophotometer-based methods, such as the Folin-Ciocâlteu assay, FRAP assay, and ORAC assay, have a low specificity because non-phenolic substances are prone to overestimate the results [131,132]. Although GC-MS technology is capable of detecting pure compounds, its limits come when the analytical technique is more sensitive than the assay employed to evaluate biological activity [133]. This might lead to the loss of trace quantities of biologically active compounds, emphasizing the need to isolate pure molecules. Furthermore, the solvent employed in the extraction should be evaluated critically at each stage of the cycle process. The technique should be clarified fundamentally, including the rationale of the maximum residual limit of solvent within the pure compound. It might be argued that optimizing extraction processes would result in high-quality phlorotannin from various seaweed sources as nutraceuticals for improving human health. However, because most research on the biological use of phlorotannins have used extracts, efforts must be made to develop a standard purification technique for the active component present in the extracts.
Table 1. Different types of phlorotannins exhibiting antimicrobial activities and their action mechanisms.
Table 1. Different types of phlorotannins exhibiting antimicrobial activities and their action mechanisms.
Pure Phlorotannin or ExtractsSourcesPurification MethodologyMicrobial PathogensActive ConcentrationTypes of the Testing MethodAction MechanismReference
Antibacterial
  • 8,8′-Bieckol
  • Eckol
  • Dieckol
  • Phloroglucinol
  • Phlorofucofuroeckol A
Ecklonia kuromeSilica acid chromatography
  • MRSA
  • Escherichia coli
  • Bacillus cereus
  • Campylobacter jejuni
  • Salmonella Typhimurium
  • S. enteritidis
  • Vibrio parahaemolyticus
  • MBC (>6.35 to 0.13 µmol/mL)
  • MBC (>6.35 to 0.27 µmol/mL)
  • MBC (>6.35 to 0.54 µmol/mL)
  • MBC (0.79 to 0.03 µmol/mL)
  • MBC (>6.35 to 0.27 µmol/mL)
  • MBC (>6.35 to 0.27 µmol/mL)
  • MBC (>6.35 to 0.27 µmol/mL)
Broth microdilution methodND[21]
  • Ethyl acetate fraction of methanolic extract
  • Phlorofucofuroeckol
  • Eckol
  • 7-Phloroeckol
  • Dieckol
  • Dioxynodehydroeckol
Eisenia bicyclisFolin-Ciocâlteu methodListeria monocytogenes
  • MIC (128 to 256 µg/mL)
  • MIC (32 to 128 µg/mL)
  • MIC (128 to 256 µg/mL)
  • MIC (64 to 128 µg/mL)
  • MIC (64 to 128 µg/mL)
  • MIC (64 to 128 µg/mL)
Micro-dilution methodND[36]
Crude methanolic extract
  • Laminaria digitata
  • L. saccharina
  • Himanthalia elongata
  • Palmaria palmata
  • Chondrus crispus
  • Enteromorpha spirulina
Folin-Ciocâlteu methodL. monocytogenes, S. abony, Enterococcus faecalis, and P. aeruginosa2.21% to 100% bacterial inhibitionTwo-fold dilution methodND[37]
PhlorotanninsHizikia fusiformeND
  • Aeromonas hydrophila
  • Chrobacterium violaceum
  • E. coli
  • Staphylococcus aureus
  • P. aeruginosa
  • V. parahaemolyticus
  • MIC 0.1943 g/mL for C. violaceum
  • MIC 0.0972 g/mL for others
Two-fold dilution method
  • Inhibited QS activity
  • Decreased bacterial motility
  • Inhibited the extracellular protease, pyocyanin, and hemolysin
  • Inhibited biofilm formation
[38]
  • Ethyl acetate fraction of methanolic extraction
  • Eckol
  • Fucofuroeckol
  • 7-Phloroeckol
  • Dioxynodehydroeckol
  • Dieckol
E. bicyclisNDMRSAMIC (32 to 64 µg/mL)Two-fold dilution methodND[39]
  • Acetone fraction of methanolic extraction
  • Dieckol
  • Phlorofucofuroeckol A
E. cavaSephadex LH-20 column chromatographyPropinibacterium acnesMIC (39 µg/mL)Broth microdilution assayND[40]
  • Ethyl acetate fraction of ethanolic extraction
  • Eckol
E. cavaSilica gel column chromatography
  • MRSA
  • Salmonella sp.
MIC (125 to 250 µg/mL)Broth microdilution methodND[42]
  • Ethyl acetate fraction of methanolic extract
  • Phlorotannins
E. cavaReversed-phase column chromatography
  • Edwardsiella tarda
  • Streptococcus parauberis
  • S. iniae
  • V. anguillarum
  • V. harveyi
  • V. scophthalmi
MIC 128 to 256 µg/mLMicro broth dilution methodND[44]
  • Eckol
  • Dieckol
E. arboreaColumn chromatographyV. parahaemolyticus
  • MBC 350 to 5.23 mg/g for eckol
  • MBC 350 to 1.67 mg/g for dieckol
Broth dilution methodDevelop antibiotic agents for medicated shrimp feed additive[45]
DieckolE. stoloniferaSephadex LH-20 column chromatographyMRSAMIC 32 to 64 μg/mLTwo-fold dilution methodND[46]
Dieckol Phlorofucofuroeckol-AE. bicyclisND
  • S. aureus
  • S. epidermidis
  • P. acnes
MIC 128 −256 μg/mLMicro-dilution methodND[41]
PhlorotanninSargassum thunbergiiNDV. parahaemolyticus900 μg/mL inhibited biofilm formationMicro-dilution methodND[47]
  • Phlorotannin extraction
  • Phloroglucinol
  • Tetrafucol A
  • Tetraphlortol B
  • Eckol
Laminaria digitataElectron micrographMixed bacterial cultureBiofilm inhibitionBatch assayND[48]
Crude phlorotanninsCymbella spp.Thin-layer chromatography
  • Corynebacterium diphtheriae
  • E. coli
  • Klebsiella pneumonia
  • Proteus mirabilis
  • S. aureus
  • S. Typhimurium
  • P. aeruginosa
MIC value of 1.56, 1.56, 3.12, >3.12, 3.12, 1.56, 1.56 mg/mL respectivelyMicro-dilution methodInactivated microbial adhesions, enzymes, and cell envelope transport proteins[134]
Antifungal
  • Acetone extracted crude extract
  • Phlorotannins
  • Cystoseira usneoides
  • C. nodicaulis
  • Fucus spiralis
Crude extractionCandida albicans ATCC 10231MIC 15.6,31.3,31.3 mg/mL for C. nodicaulis, C. usneoides, and F. spiralis, respectivelyBroth microdilution method
  • Affected the ergosterol composition of the cell membrane
  • Increased the mitochondrial dehydrogenase
  • Inhibited dimorphic transition of fungi
[22]
  • Methanolic extraction
  • Dieckol
E. cavaSilica-gel chromatographyTrichophyton rubrumMIC 200 µMMicro broth dilution assayChanged cytoplasmic integrity[49]
  • Ethylacetate fraction of methanolic extract
  • Fucofuroeckol A
E. bicyclisRP-18 open column chromatography and Sephadex LH-20 C. albicansMIC 512 µg/mLBroth microdilution assay
  • Induced ROS species
  • Disrupted fungi cell wall
[50]
Anti-viral
  • Ethylacetate extract of methnol extract
  • Phloroglucinol derivatives
  • 6,6′-Bieckol
E. cavaSilica-gel chromatographyHIV-1EC50 1.72 µM (syncytia production)
EC50 1.26 µM (antigen production)
Western blot analysis
Cell viability assay
  • Induced syncytia production
  • Inhibited viral p24 antigen production
  • Inhibited lytic effect
[57]
8,4′-DieckolE. cavaSilica-gel chromatographyHIV-191% Inhibition of reverse transcriptase at 50 µMReverse transcriptase assay
  • Induced syncytia production
  • Inhibited viral p24 antigen production
  • Inhibited lytic effect
[57]
  • 8,4′-Dieckol
  • 8,8′-Bieckol
  • Phlorofucofuroeckol A
E. cavaNDHIV-150% Inhibition of reverse transcriptase at 0.51 µMReverse transcriptase assayInhibited reverse transcriptase enzyme activity[58]
  • Ethyl acetate fraction of ethanolic extract
  • Triphloretol
  • Eckol
  • Dioxynodehydroeckol
  • Dieckol
  • 2-Phloroeckol
  • 7-Phloroeckol
  • Phlorofucofuroeckol A
  • Fucodiphloroethol
E. cavaSilica-gel chromatographySARS-CoV 3CLIC50 2.7 to >200 µM
  • Cell-free /based analysis
  • SARS-CoV 3CLpro trans-cleavage assay
Inhibited 3CLpro hydrolysis[23]
PhlorofucofuroeckolE. cavaSilica-gel column chromatographyInfluenza A virusIC50 4.5 µMChemiluminescent neuraminidase inhibition assayND[59]
Table 2. Different types of phlorotannins exhibiting anticancer, anti-inflammatory, antidiabetic, antioxidant, anti-obesity, anti-adipogenesis, and other biological activities.
Table 2. Different types of phlorotannins exhibiting anticancer, anti-inflammatory, antidiabetic, antioxidant, anti-obesity, anti-adipogenesis, and other biological activities.
Pure Phlorotannin or ExtractsSourcesPurification MethodologyBiological ActivityActive ConcentrationTypes of the Testing MethodAction MechanismReference
Anticancer
  • 6,6′-Bieckol
  • 8,8-Bieckol
  • Dieckol
  • 7-Phloroeckol
  • Phlorofucofuroeckol
Ecklonia cavaNDOvarian cancer cells undergo apoptosis
  • IC50 84.3 and 99.6 µg/mL against A2780 and SKOV3 cells for ethanolic extract
  • IC50 77.31 and 92.7 µg/mL against A2780 and SKOV3 cells for dieckol
  • IC50 77.31 to 137.77 and 96.3 to >200 µg/mL against A2780 and SKOV3 cells for others
  • PI staining
  • MTT assay
  • PI and Annexin double staining
  • Western blot analysis and flow cytometry and
  • SKOV3 tumor xenograft model
  • Inhibited tumor xenograft growth
  • Cytotoxicity effect on ovarian cancer cells (A2780 and SKOV3)
  • Caused mitochondria disfunction
  • Suppressed the levels of anti-apoptosis protein
  • Induced ROS
  • Reversed the caspase activation
[78]
  • Phloroglucinol derivatives
  • Dioxynodehydroeckol
  • 1-(3′,5′-Dihydroxyphenoxy)-7-(2″,4″,6-trihydroxyphenoxy)-2,4,9-trihydroxydibenzo-1,4-dioxin
E. cavaSilica-gel column chromatographyHuman breast cancer cells
  • At 100 µM, it inhibited MCF-7 cells growth by 64%
  • Inhibited proliferation of MDA-MB-231 cells 55% at 100 µM
Cell proliferation assay
  • Anti-proliferative activity on MCF-7 cancer cells
  • Downregulated NF-κB and depended on activated genes
  • Induced caspase activity
  • Induced the cleavage of DNA repair enzyme
  • Induced pro-apoptotic gene
  • Suppressed anti-apoptotic gene
[79]
PhloroglucinolBrown seaweedsNDMDA-MB231 breast cancer cellsIC50 50 µM of migratory and invasive ability of MDA-MB231 breast cancer cells
  • Western blot analysis
  • Activated RAS affinity precipitation assay
  • Invasion and migration assay
  • SLUG was suppressed by inhibiting P13K/AKT and RAS/RAF-1/ERK signaling
  • Reduced cancer cell’s metastatic ability
[80]
PhloroglucinolSeaweedsNDHuman colon cancer cells HT-2912.5 µg/mL caused fragmented nuclei and cell shrinkage
  • Western blot analysis
  • Cell cycle analysis
  • mRNA expression analysis
  • 4′-6′-Diamidino-2-phenylindole staining assay
  • Downregulated the expression of Ras, Raf, and mitogen-activated protein kinase
  • Downregulated the phosphorylation of the extracellular-signal-regulated kinase
  • Decreased mammalian target of rapamycin and expression of its downstream effectors p70S6 kinase
  • Decreased elF4b and RPS6 translation initiation factor
[135]
PhlorotanninsCystoseira sedoidesReversed-phase column chromatographyMCF-7 cells (human breast cancer cells)In MCF-7 cells, the IC50 for inducing apoptosis was 78 µg/mLDouble Annexin V-FITC/PI testPrevented spheroid growth[82]
DieckolE. cavaSephadex LH-20 column chromatographyProtective efficacy against N-nitrosdiethylamine -induced rat hepatocarcinogenesisAlkaline phosphatase, lactate dehydrogenase, transaminase, gamma-glutamyl transferase, total bilirubin, and a-fetoprotein activities increased in NDEA-induced rats given dieckol water (10–40 mg/kg body weight) for 15 weeksSerum marker enzymes analysis
  • It prevents hepatic cell damage and lipid peroxidation
  • NDEA-induced hepatocarcinogenesis enhances the enzymatic and non-enzymatic antioxidant defense system
[83]
IsololiolideC. tamariscifoliaReverse phase preparative HPLCAnti-proliferative activity
  • IC50 32.36 µM of cytotoxic activity against gastric cancer cells
  • IC50 23.59 µM of cytotoxic activity against colon cancer cell line
  • IC50 13.15 µM of cytotoxic activity against human hepatocellular carcinoma cells
MTT colorimetric assay
  • Expression of proteins in the apoptotic cascade
  • Induced apoptosis through the modulation of apoptosis-related proteins.
[136]
Anti-inflammatory
  • Ethyl acetate fraction of methanolic extraction
  • Phloroglucinol (PG)
  • Phlorofucofuroeckol A (PFA)
  • Eckol (EK)
  • 7-Phloroeckol (7-PE)
  • Dieckol (DE)
Eisenia bicyclisChromatography on silica gel columnCell viability and NO production (LPS-induced RAW264.7 cells)
  • >10 µg/mL cytotoxicity for PG and PFA
  • 24.5% and 66.2% inhibition of NO production for PG and PFA
  • IC50 52.86, 51.42, and 26.87 µg/mL for EK, DE, 7-PE
MTT assay
  • Inhibited NF-κB -related inflammatory gene expression via ROS inhibition
  • Inhibited LPS-induced nitric oxide and butyl hydroperoxide-induced ROS production
[64]
PhloroglucinolE. cavaNDInhibition of oxidative stress and inflammation
  • Inhibition of protein oxidation (90% at 10 µM)
  • Inhibition of TNF-α, IL-1β (30% at 10 µM)
  • Inhibition of PGE2 (40% at 10 µM)
  • MTT assay
  • Membrane protein oxidation assay
  • Enzyme immunoassay
  • Western blot analysis
  • Inhibited oxidative stress
  • Inhibited the production of IL-6, TNF-α, IL-1β, and PGE2
  • Matrix metalloproteinase expression was reduced
[66]
PhlorotanninsBifurcaria bifurcataReversed-phase column chromatographyCytotoxic effect on ATDC-5 mouse model cell lines50% inhibition of cell growth in ATDC-5 cells at 100 µMMTT assayND[68]
PhlorotanninsPadina tetrastromaticaReversed-phase column chromatographyEffect on THP-1 cell viabilityFrom 1.5 to 50.0 µg/mLMTT assay
  • Anti-MRSA potential
  • Enhances the high-glucose-induced pro-MMP-9 expression
[69]
PhloroglucinolBrown algaeNDAnti-inflammatory effect and oxidative stress on RAW264.7 and HT1080 cellsFrom 1 to 100 µMMTT assay
  • Inhibitory effects on oxidative stress and the production of inflammatory mediators such as IL-6, TNF-a, IL-1β, and PGE2 in RAW264.7 cells stimulated by LPS
  • Decreased the expression of matrix metalloproteinase in HT1080 cells
[66]
Phlorotannins
  • Fucus guiry
  • F. serratus
  • F. sprialis
  • F. vesiculosus-wild
  • F. vesiculous-aquaculture
Aqueous acetone extractionAnti-inflammatory and toxicity capability in RAW 264.7 macrophages and cell-free systems
  • IC50 82.10, 110.16, 362.42, 364.84, > 500 μg DE/mL for LOX inhibition, respectively
  • IC50 451.91, 1214.73, 801.97, 1330.61, 2072.32 μg DE/mL for NO scavenging, respectively
  • IC25 97.73, 77.04, 95.86, 56.52, 317.41 μg DE/mL for NO reduction, respectively
  • Nontoxic at 31.25–500 μg/mL in RAW 264.7 macrophages stimulated with bacterial LPS
MTT assayFunction in inflammatory conditions, acting on both enzymatic and non-enzymatic inflammatory targets[65]
Phlorofucofuroeckol B (PFF-B)E. stoloniferaHigh-performance chromatographyPFF-B inhibits the generation of inflammatory mediators induced by LPS
  • Decreased secretion of pro-inflammatory cytokines, including TNF-α, IL-1β, and IL-6
  • Decreased expression of pro-inflammatory proteins such as COX-2 and inducible NO synthase
  • MTS assay
  • NF-κB promoter/luciferase assay
  • By limiting the breakdown of the inhibitor κb-α, the activation of nuclear factor kappaB was prevented.
  • Inhibited the phosphorylation of Akt, ERK, and JNK
[70]
6,6′-BieckolE. cavaNDLPS-stimulated macrophage RAW 264.7 cells have anti-inflammatory effects
  • Inhibited NO and PGE2 production at concentrations of 100, 200 μM
  • Suppressed the LPS-induced expression of COX-2 and inducible-NOS and at the mRNA and protein levels
  • TNF-α and IL-6 mRNA expression were reduced
  • MTT assay
  • PGE2 assay
  • Chromatin immunoprecipitation assay
Downregulation of COX-2, iNOS, and pro-inflammatory cytokines in LPS-stimulated macrophages via the NF-κB pathway[71]
Phlorofucofuroeckol BMyagropsis myagroidesHigh-performance chromatographyAnti-inflammatory activity
  • Inhibited LPS-induced PGE2 and NO production
  • Inhibited COX-2 and iNOS
  • Pro-inflammatory cytokines were inhibited, as well as NF-κB activation and translocation
  • MTS assay
  • Transient transfection and luciferase assay
In LPS-stimulated macrophage cells, the NF-κB pathway was inhibited by inhibiting the phosphorylation of ERKs and Akt[67]
Antidiabetic
  • Ethyl acetate fraction of methanolic extraction
  • Dioxynodehydroeckol (DDE)
  • Fucofuroeckol (FFE)
E. bicyclisChromatography and nuclear magnetic resonanceAntidiabetic activity
  • IC50 131.34 nmol/L for α-glucosidase and 42.91 µmol/L for α-amylase for FFE
  • IC50 93.33 nmol/L for α-glucosidase and 472.7 µmol/L for α-amylase for DDE
Enzymatic inhibitory assayInhibition of α-amylase and α-glucosidase enzyme activities[106]
  • Phlorofucofuroeckol
  • Dieckol
  • 6-6′-Bieckol
  • 7-Phloroeckol
  • Fucodiphloroetho
E. cavaColumn chromatography using silica gelantidiabetic activityIC50 10.75 to 49.49 µmol/mL for α-glucosidase and >500 to 124.98 µmol/L for α-amylaseEnzymatic inhibitory assayInhibition of α-amylase and α-glucosidase enzymes activities[107]
PhlorotanninsC. compressaUPLC-MS/MSAntidiabetic activityAfter four weeks of diabetes induction, diabetics were treated with 60 mg/kg of phlorotannin extract.MTT assay
  • Reduced serum glucose, malondialdehyde, glucosidase, and α-amylase activity in the liver
  • Reduced damage in pancreatic β cells
[108]
Ethylacetate fraction of acetone extractF. vesiculosusMass spectroscopy (UHPLC-MS)Antidiabetic and anti-obesity activity
  • IC50 for α-amylase 2.8 µg/mL
  • IC50 for α-glucosidase 0.82 µg/mL
  • IC50 for pancreatic lipase 0.82 µg/mL
α-amylase, α-glucosidase, pancreatic lipase inhibitory assay
  • Inhibited α-amylase, α-glucosidase, and pancreatic lipase enzymes
  • Delayed carbohydrates and lipid digestion
[109]
Ishophloroglucin AIshige okamuraeSemipreparative HPLC columnAnti-α-glucosidase activityIC50 value of 54.97 µM in α-glucosidase inhibitionα-Glucosidase inhibitory assayInhibited α-glucosidase[110]
Phlorofucofuroeckol AE. cavaElectrospray ionization-multistage tandem mass spectrometry and HPLCAntidiabetic activity
  • IC50 for α-amylase 6.34 µM
  • IC50 for α-glucosidase 19.52 µM
  • α-Glucosidase inhibitory assay
  • α-Amylase inhibitory assay
Inhibition of α-amylase and α-glucosidase enzymes activities[111]
DieckolE. cavaReversed-phase HPLCAntidiabetic activity
  • Dieckol administration reduced blood glucose levels, serum insulin levels, and body weight
  • Increased activity of anti-oxidant enzymes in liver tissues, including superoxide dismutase, catalase, and glutathione peroxidase
  • Serum glucose content reduced
  • Lipid peroxidation production
  • Catalase and superoxide dismutase activities
Activates both the AMPK and PKB signaling cascades[112]
Antioxidant
Dichloromethane fraction of methanolic extraction
  • C. trinodis (Forsskal)
  • C. Agardh
Silica-gel column chromatographyAntioxidant activity69.62% radical scavenging activityDPPH radical scavengingND[89]
  • Dichloromethane fraction of ethanolic extraction
  • 974-A
  • 974B
  • Phloroglucinol
  • Dieckol
E. curomeSequential chromatography on two reverse phase columnAntioxidant activityIC50 10, 11, 110, 10 µM respectivelyDPPH radical scavengingReduced intracellular reactive oxygen species[90]
  • Ethylacetate fraction of methanolic extract
  • Phloroglucinol
  • Eckol
  • Dieckol
E. cavaSephadex LH-20 column chromatographyAntioxidant activity
  • 90% radical scavenging activity for eckol at 0.25 to 1 mg/mL
  • 100% hydroxyl radical scavenging activity for dieckol at 0.25 mg/mL
  • 90% alkyl radical scavenging activity for eckol at 0.05 mg/mL
  • 68.96% to 50% DNA damage values for all three at 25 µg/mL
  • DPPH radical assay
  • Hydroxyl radical assay
  • Alkyl radical assay
  • Superoxide radical assay
  • Comet assay (Protecting effects against H2O2-mediate DNA damage
  • Inhibited free radical activities
  • Damaged DNA
[91]
  • Ethanolic extract
  • Phlorotannins
Sargassum inggoldianmMatrix-assisted laser desorption ionization time-of-flight mass spectroscopyAntioxidant activityIC50 1.0 µg/mLElectron spin resonance spectrometryShowed superoxide anion radical scavenging activity[92]
PhlorotanninsS. dupplicatumSephadex LH-20 column chromatographyAntioxidant activity
  • Total antioxidant activity 11.17 ± 0.28 mg ascorbic acid equivalent/g DW
  • Reducing power activity by 11.09 ± 0.24 mg FeSO4 equivalent/g DW
  • Antioxidant activity
  • Reducing power activity
ND [93]
Phlorofucofuroeckol-AE. cavaCentrifugal partition chromatographyAntioxidant activity
  • Scavenging activity against alkyl radicals, with an IC50 value of 3.9 µM
  • With an IC50 value of 21.4 µM, it has a scavenging action against hydroxyl radicals
  • With an IC50 value of 10.3 µM, it has a scavenging action against DPPH radicals
  • Assay for alkyl radical scavenging capacity
  • Assay for hydroxyl radical scavenging capacity
  • DPPH assay
  • Antioxidant and lipid peroxidation protection
  • Malondialdehyde inhibition in AAPH-stimulated Vero cells
[94]
  • Dieckol
  • Eckol
  • Eckstolonol
  • Phloroglucinol
  • Triphloroethol A
E. cavaHPLCNeuroprotective against H2O2-induced cellular damage in HT22 neuronal cells from the murine hippocampus
  • ROS levels were reduced to 75.22%, 69%, 67.32%, 77.63%, 77.73%, respectively (treatment of H2O2-treated cells at a concentration of 50 M)
  • The apoptotic sub-G1 DNA content was reduced to 9.55%, 13.03%, 5.55%, 6.61%, 3.03%, respectively (pre-treatment with 50 M phlorotannins)
  • Scavenging efficacy on ROS production
  • Neuroprotective effects in H2O2-treated HT22 cells
  • Remove intracellular ROS and inhibit ROS accumulation
  • Inhibited H2O2-induced Ca2+ release
[96]
  • Dieckol
  • Eckol
  • Eckstolonol
  • Phloroglucinol
  • Triphloroethol A
E. cavaSephadex LH-20 column chromatographyThe ability of phlorotannins to scavenge ROS in AAPH-induced zebrafish embryosReduced intracellular ROS buildup to DCF-DA intensity of 1568, 2346, 1703, 1540, and 2262, respectively (50 µM phlorotannins + 25 mM AAPH)ROS generation in AAPH-induced zebrafish embryosAntioxidant efficacy against AAPH-mediated toxicity[95]
Anti-obesity
  • Ethylacetate fraction of methanol extraction
  • 6-6′-Bieckol
  • Dieckol
  • Phlorofucofuroeckol A
E. cavaSephadex LH-20 column chromatographyInhibition of adipogenesisInhibited lipid accumulation 60%, 40% 20% at 100 µL respectivelyOil-Red O staining methodDownregulated adipogenic specific genes (SREBP-1, C/EBPα, FABP4, and PPARγ)[116]
  • Ethylacetate fraction of ethanolic extraction
  • Dioxynodehydroeckol
E. cavaSilica-gel column chromatographyInhibition of adipogenesisInhibited adipogenesis 20 µMOil-Red O staining method
  • Reduced lipid accumulation
  • Downregulated the expression of adipogenic specific genes of 3T3-L1 (C/EBPα, SREBP1, and PPARγ)
  • Activated and modulated AMPK signaling pathway
[113]
  • Diethyl ether fraction of methanolic extract
  • Dieckol
  • Eckol
E. cavaSephadex LH-20 column chromatographyInhibition of melanogenesis
(UV-B radiation-induced cell damage protection effect)
92.7% (dieckol) and 62.4% (eckol) inhibitory effect on tyrosinase at 100 µM
  • Tyrosinase inhibition assay
  • DCFH-DA, MTT, comet assay
  • Inhibited tyrosinase
  • Inhibited melamine syntheses
[114]
Anti-adipogenesis
  • Ethylacetate fraction of methanolic extraction
  • Phloroglucinol
  • Eckol
  • Dieckol
  • Dioxynodehydroeckol
  • Phlorofucofuroeckol A
E. stoloniferaSephadex LH-20 column chromatographyInhibition of adipogenesisInhibited adipogenesis (12.5–100 µM)Oil Red O stainingDownregulated adipogenic specific genes of 3T3-L1 (C/EBPα and PPARγ)[100]
DieckolLaminaria japonicaSilica gel resin absorptionAnti-fatty liver activityBody weight gain, plasma lipid profiles, visceral fat index, liver index, and hepatic fat deposition were improved in high-fat diet mice given a dieckol-enriched extract (50 mg/kg/day) for four weeksHistopathological analysis
  • Advantageous effects on hepatic lipid metabolism
  • Stimulation of hepatic fatty acid β-oxidation
[115]
DieckolE. cavaReverse-phase high-performance liquid chromatographyInhibitory effect on adipogenesis
  • Inhibition of adipogenesis
  • Peroxisome proliferator-activated receptor, CCAAT/enhancer-binding proteins, fatty acid-binding protein, and SREBP1 expression were all reduced
MTT assayActivated AMP-activated protein kinase[116]
Other biological activities
  • Ethyl acetate fraction of methanolic extract
  • Phloroglucinol
E. cavaSephadex LH-20 column chromatographyCytoprotective effect against oxidative stress-induced cell damage in (V79-4) Chinese hamster lung fibroblast
  • 65% DPPH radical scavenging activity at 10 µg/mL
  • 70% H2O2 scavenging activity at 10 µg/mL
  • 26% hydroxy radicals scavenging activity at 10 µg/mL
  • 73% Intracellular reactive oxygen scavenging activity at 10 µg/mL
  • Radical scavenging activity assay
  • MTT assay
  • Western blot analysis
  • Scavenge 1,1-diphenyl-2-picrylhydrazyl (DPPH) radical
  • Scavenge H2O2, hydroxy radicals
  • Prevented lipid peroxidation
  • Reduced intracellular reactive oxygen species
  • Increased the activity of catalase
  • Increased phosphorylation of the extracellular signal-regulated kinase
[137]
  • Ethyl acetate fraction of methanolic extract
  • Phloroglucinol
E. cavaSephadex LH-20 column chromatographyRadioprotective effect of cells against γ-ray radiation-induced oxidative damage
  • Phloroglucinol-treated cells reduced irradiation DNA damage by 27%
  • Cell survival rose to 66% in irradiated cells at 10 µg/mL phloroglucinol
  • Laser scanning microscopy
  • Colorimetric assay
  • Lipid peroxidation assay
  • Western blot analysis
  • Comet assay
  • Reduced the amount of radiation-induced intracellular ROS
  • Reduced cellular component (lipid, DNA, and protein) damage
  • Reduced the radiation-induced loss of mitochondrial membrane action potential
  • Reduced the active levels of caspase-3 and 9
[138]
PhloroglucinolSeaweedsNDUVB-induced photoaging of human HaCaT keratinocytes
  • 10 µg/mL protected HaCaT keratinocytes against UVB-induced cytotoxicity
  • 10 µg/mL inhibited the accumulation of UVB-induced MMP-1 mRNA and protein at 48 h
  • UV/visible light absorption analysis
  • Intracellular ROS measurement
  • Western blot analysis
  • Human active MMP-1 Fluorokine E fluorescent assay
  • Chromatin immunoprecipitation assay
  • Upregulated MMP-1 mRNA and protein activity
  • Augmented intracellular Ca2+ level
  • Phosphorylation of mitogen-activated protein kinases
  • Enhanced the activator protein-1 (AP-1) binding to the MMP-1 promotor
[139]
PhlorotanninsE. radiataHigh-performance counter-current chromatographyNeuroprotective activity
  • In neuronal PC-12 cells, it was nontoxic up to 50 µM
  • Cholinesterase inhibitory activity at IC50 value of 41 µM
  • MTT assay
  • Acetylcholinesterase inhibitory assay
  • Elicited neuroprotective activity against amyloid β protein, Aβ1–42
  • Inhibited Aβ1–42 aggregation, AChE activity, and ROS formation in PC-12 cells
  • Neuroprotective effects through multiple pathways
[140]
PhlorotanninsE. cavaSephadex LH-20 column chromatographyMG63 cell survival and calcium deposition on polycaprolactone (PCL/Ph) micro nanofibres
  • Increased cell survivability on PCL/Ph micro nanofibers as phlorotannin content increased
  • Increased calcium mineralization on the PCL/Ph micro nanofibers after 14 days
MTT assayEnhanced bone tissue growth[116]
DioxynodehydroeckolE. cavaSilica-gel column chromatographyUVB-induced apoptosis prevention in human keratinocyte (HaCaT) cellsReduced by 1.83% at 20 µM of DHE compared to 13.31% in cells exposed to UVBFlowcytometry following Annexin V and PI labelingND[141]
DieckolE. cavaSephadex LH-20 with MeOHAnti-proliferative and anti-angiogenic effect on EA. hy926 cell lines induced with vascular endothelial growth factor
  • Nontoxic up to 100 µM in EA. hy926 cells
  • Inhibition activity of AP-N enzyme with an IC50 value of 52.80 µM
  • Dieckol at the concentration of 10, 50, and 100 µM inhibited vascular endothelial growth factor-induced cell proliferation.
  • MTT assay
  • Aminopeptidase-N enzyme assay
  • Cell proliferation assay
Cell migration was inhibited by lowering the level of protein and gene expression of matrix metalloproteinases such as MMP-2 and -9[142]

9. Conclusions and Future Perspectives

In conclusion, the present review paper describes the many forms of phlorotannins produced from various marine algae sources. These phlorotannins, which are polyphenolic compounds derived from brown algae, have a wide range of biological characteristics, including antimicrobial, antidiabetic, anticancer, and anti-inflammatory properties (Table 1 and Table 2). The vast spectrum of biological activities associated with phlorotannins is projected to increase their favorable health value in the food, pharmaceutical, and cosmeceutical industries. Despite the prospective uses of phlorotannins, the successful development of a series of products produced from brown algal polyphenols as nutraceuticals have remained an unmet goal. Nutritional epidemiology studies demonstrate a strong link between consuming polyphenols and modifying the molecular pathways of carcinogenesis, as well as lowering cancer cell growth. Despite its extensive use as an antimicrobial, anticancer, antidiabetic, and anti-inflammatory agent, phlorotannin’s low stability and solubility restrict its application. In recent trends, nanotechnology, particularly nanocarriers and nanoformulation, are evolving to overcome the limits associated with natural products and allow targeted and controlled drug delivery. To treat microbial infections, the phlorotannin must be loaded onto the nanocarrier (metallic or nonmetallic) in such a way that the nanocarriers are biocompatible and have inherent antimicrobial capabilities allowing for the synergistic killing of the microbial pathogens [143].
Furthermore, combination therapy has emerged as a viable strategy for revitalizing antibiotics or antifungals in order to fight microbial pathogens in a synergistic manner. The reported chemical structures for some of the phlorotannins (Figure 1) can be employed as scaffolds for the synthesis of the derivatives by grafting some of the active molecules on the phlorotannin to produce broad-spectrum and effective biological effects. Studying the structure-activity relationships of phlorotannins is essential in order to explain the molecular action mechanisms in various biological roles, since various antimicrobial resistance signaling proteins and transcription factors, such as efflux pumps, biofilms, virulence, reverse transcriptase, protease, integrase, and so on, have been thoroughly described in microbial pathogens. As a result, elucidating the molecular docking-interaction of phlorotannins with known microbial proteins relevant to pathogenesis and resistance mechanisms will be more intriguing in future studies. Aside from gene expression studies of microbial virulence, biofilms, antimicrobial resistance genes, and genes implicated in cancer signaling pathways, inflammation, diabetes, and obesity would provide more supporting research to validate the results of the in vitro studies. In addition, to confirm the in vitro antimicrobial activity of phlorotannins, an in vivo investigation employing an animal model organism is required.

Author Contributions

F.K.: Conceptualization, literature search, writing—original draft, writing—review and editing. G.-J.J.: literature search, writing—original draft and editing, M.S.A.K.; literature search, writing, and editing, N.T.: literature search, writing, and editing, and Y.-M.K.: supervision, funding, writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by Basic Science Research Program through the National Research Foundation of Korea (NRF) grant funded by the Ministry of Education (2021R1A6A1A03039211 and 2022R1A2B5B01001998).

Institutional Review Board Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ghosh, S.; Sarkar, T.; Pati, S.; Kari, Z.A.; Edinur, H.A.; Chakraborty, R. Novel Bioactive Compounds from Marine Sources as a Tool for Functional Food Development. Front. Mar. Sci. 2022, 9, 832957. [Google Scholar] [CrossRef]
  2. Bamunuarachchi, N.I.; Khan, F.; Kim, Y.M. Antimicrobial Properties of Actively Purified Secondary Metabolites Isolated from Different Marine Organisms. Curr. Pharm. Biotechnol. 2021, 22, 920–944. [Google Scholar] [CrossRef] [PubMed]
  3. Lomartire, S.; Gonçalves, A.M.M. An Overview of Potential Seaweed-Derived Bioactive Compounds for Pharmaceutical Applications. Marine Drugs 2022, 20, 141. [Google Scholar] [CrossRef] [PubMed]
  4. Singh, I.P.; Sidana, J. Chapter 5—Phlorotannins. In Functional Ingredients from Algae for Foods and Nutraceuticals; Domínguez, H., Ed.; Woodhead Publishing: Cambridge, UK, 2013; pp. 181–204. [Google Scholar]
  5. Yang, H.; Zeng, M.; Dong, S.; Liu, Z.; Li, R. Anti-proliferative activity of phlorotannin extracts from brown algae Laminaria japonica Aresch. Chin. J. Oceanol. Limnol. 2010, 28, 122–130. [Google Scholar] [CrossRef]
  6. Isaza Martínez, J.H.; Torres Castañeda, H.G. Preparation and Chromatographic Analysis of Phlorotannins. J. Chromatogr. Sci. 2013, 51, 825–838. [Google Scholar] [CrossRef] [Green Version]
  7. Yang, B.; Fang, D.; Lv, Q.; Wang, Z.; Liu, Y. Targeted Therapeutic Strategies in the Battle Against Pathogenic Bacteria. Front. Pharmacol. 2021, 12, 673239. [Google Scholar] [CrossRef]
  8. Spitzer, M.; Robbins, N.; Wright, G.D. Combinatorial strategies for combating invasive fungal infections. Virulence 2017, 8, 169–185. [Google Scholar] [CrossRef] [Green Version]
  9. Meganck, R.M.; Baric, R.S. Developing therapeutic approaches for twenty-first-century emerging infectious viral diseases. Nat. Med. 2021, 27, 401–410. [Google Scholar] [CrossRef]
  10. Ahmad, B.; Serpell, C.J.; Fong, I.L.; Wong, E.H. Molecular Mechanisms of Adipogenesis: The Anti-adipogenic Role of AMP-Activated Protein Kinase. FFront. Mol. Biosci. 2020, 7, 76. [Google Scholar] [CrossRef]
  11. Hurt, R.T.; Kulisek, C.; Buchanan, L.A.; McClave, S.A. The obesity epidemic: Challenges, health initiatives, and implications for gastroenterologists. Gastroenterol. Hepatol. 2010, 6, 780–792. [Google Scholar]
  12. Chakraborty, S.; Rahman, T. The difficulties in cancer treatment. Ecancermedicalscience 2012, 6, ed16. [Google Scholar] [PubMed]
  13. Yang, X.; Yang, H.; Zhou, G.; Zhao, G.-P. Infectious Disease in the Genomic Era. Annu. Rev. Genom. Hum. Genet. 2008, 9, 21–48. [Google Scholar] [CrossRef] [PubMed]
  14. Reedy, J.L.; Bastidas, R.J.; Heitman, J. The Virulence of Human Pathogenic Fungi: Notes from the South of France. Cell Host Microbe 2007, 2, 77–83. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Heise, M.T. Viral Pathogenesis. In Reference Module in Biomedical Sciences; Elsevier: Amsterdam, The Netherlands, 2014. [Google Scholar]
  16. Peterson, E.; Kaur, P. Antibiotic Resistance Mechanisms in Bacteria: Relationships Between Resistance Determinants of Antibiotic Producers, Environmental Bacteria, and Clinical Pathogens. Front. Microbiol. 2018, 9, 2928. [Google Scholar] [CrossRef] [PubMed]
  17. Dhingra, S.; Rahman, N.A.A.; Peile, E.; Rahman, M.; Sartelli, M.; Hassali, M.A.; Islam, T.; Islam, S.; Haque, M. Microbial Resistance Movements: An Overview of Global Public Health Threats Posed by Antimicrobial Resistance, and How Best to Counter. Front. Public Health 2020, 8, 535668. [Google Scholar] [CrossRef]
  18. van Duin, D.; Paterson, D.L. Multidrug-Resistant Bacteria in the Community: Trends and Lessons Learned. Infect. Dis. Clin N. Am. 2016, 30, 377–390. [Google Scholar] [CrossRef] [Green Version]
  19. Arendrup, M.C.; Patterson, T.F. Multidrug-Resistant Candida: Epidemiology, Molecular Mechanisms, and Treatment. J. Infect. Dis. 2017, 216 (Suppl. 3), S445–S451. [Google Scholar] [CrossRef] [Green Version]
  20. Villa, T.G.; Feijoo-Siota, L.; Rama, J.L.R.; Ageitos, J.M. Antivirals against animal viruses. Biochem. Pharmacol. 2017, 133, 97–116. [Google Scholar] [CrossRef]
  21. Nagayama, K.; Iwamura, Y.; Shibata, T.; Hirayama, I.; Nakamura, T. Bactericidal activity of phlorotannins from the brown alga Ecklonia kurome. J. Antimicrob. Chemother. 2002, 50, 889–893. [Google Scholar] [CrossRef] [Green Version]
  22. Lopes, G.; Pinto, E.; Andrade, P.B.; Valentão, P. Antifungal activity of phlorotannins against dermatophytes and yeasts: Approaches to the mechanism of action and influence on Candida albicans virulence factor. PLoS ONE 2013, 8, e72203. [Google Scholar] [CrossRef]
  23. Park, J.-Y.; Kim, J.H.; Kwon, J.M.; Kwon, H.-J.; Jeong, H.J.; Kim, Y.M.; Kim, D.; Lee, W.S.; Ryu, Y.B. Dieckol, a SARS-CoV 3CL(pro) inhibitor, isolated from the edible brown algae Ecklonia cava. Bioorganic Med. Chem. 2013, 21, 3730–3737. [Google Scholar] [CrossRef] [PubMed]
  24. Klil-Drori, A.J.; Azoulay, L.; Pollak, M.N. Cancer, obesity, diabetes, and antidiabetic drugs: Is the fog clearing? Nat. Rev. Clin. Oncol. 2017, 14, 85–99. [Google Scholar] [CrossRef] [PubMed]
  25. Scully, T.; Ettela, A.; LeRoith, D.; Gallagher, E.J. Obesity, Type 2 Diabetes, and Cancer Risk. Front. Oncol. 2021, 10, 615375. [Google Scholar] [CrossRef] [PubMed]
  26. Zatterale, F.; Longo, M.; Naderi, J.; Raciti, G.A.; Desiderio, A.; Miele, C.; Beguinot, F. Chronic Adipose Tissue Inflammation Linking Obesity to Insulin Resistance and Type 2 Diabetes. Front. Physiol. 2020, 10, 1607. [Google Scholar] [CrossRef] [PubMed]
  27. Ediriweera, M.K.; Tennekoon, K.H.; Samarakoon, S.R. A review on ethnopharmacological applications, pharmacological activities, and bioactive compounds of Mangifera indica (Mango). Evid.-Based Complementary Altern. Med. 2017, 2017, 6949835. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Boucher, H.W.; Talbot, G.H.; Bradley, J.S.; Edwards, J.E.; Gilbert, D.; Rice, L.B.; Scheld, M.; Spellberg, B.; Bartlett, J. Bad bugs, no drugs: No ESKAPE! An update from the Infectious Diseases Society of America. Clin. Infect. Dis. 2009, 48, 1334. [Google Scholar] [CrossRef] [Green Version]
  29. Zacchino, S.A.; Butassi, E.; Di Liberto, M.; Raimondi, M.; Postigo, A.; Sortino, M.J.P. Plant phenolics and terpenoids as adjuvants of antibacterial and antifungal drugs. Phytomedicine 2017, 37, 27–48. [Google Scholar] [CrossRef]
  30. Da Silva, C.M.; da Silva, D.L.; Modolo, L.V.; Alves, R.B.; de Resende, M.A.; Martins, C.V.; de Fátima, Â.J. Schiff bases: A short review of their antimicrobial activities. J. Adv. Res. 2011, 2, 1–8. [Google Scholar] [CrossRef] [Green Version]
  31. Li, T.; Li, L.; Du, F.; Sun, L.; Shi, J.; Long, M.; Chen, Z.J.M. Activity and Mechanism of Action of Antifungal Peptides from Microorganisms: A Review. Molecules 2021, 26, 3438. [Google Scholar] [CrossRef]
  32. Dedeurwaerdere, S.; Friedman, A.; Fabene, P.F.; Mazarati, A.; Murashima, Y.L.; Vezzani, A.; Baram, T.Z. Finding a better drug for epilepsy: Antiinflammatory targets. Epilepsia 2012, 53, 1113–1118. [Google Scholar] [CrossRef]
  33. Shannon, E.; Abu-Ghannam, N. Antibacterial derivatives of marine algae: An overview of pharmacological mechanisms and applications. Mar. Drugs 2016, 14, 81. [Google Scholar] [CrossRef] [PubMed]
  34. Shrestha, S.; Zhang, W.; Smid, S.D. Phlorotannins: A review on biosynthesis, chemistry and bioactivity. Food Biosci. 2021, 39, 100832. [Google Scholar] [CrossRef]
  35. Besednova, N.N.; Andryukov, B.G.; Zaporozhets, T.S.; Kryzhanovsky, S.P.; Kuznetsova, T.A.; Fedyanina, L.N.; Makarenkova, I.D.; Zvyagintseva, T.N. Algae polyphenolic compounds and modern antibacterial strategies: Current achievements and immediate prospects. Biomedicines 2020, 8, 342. [Google Scholar] [CrossRef] [PubMed]
  36. Kim, H.-J.; Dasagrandhi, C.; Kim, S.-H.; Kim, B.-G.; Eom, S.-H.; Kim, Y.-M. In vitro Antibacterial Activity of Phlorotannins from Edible Brown Algae, Eisenia bicyclis Against Streptomycin-Resistant Listeria monocytogenes. Indian J. Microbiol. 2018, 58, 105–108. [Google Scholar] [CrossRef]
  37. Cox, S.; Abu-Ghannam, N.; Gupta, S. An assessment of the antioxidant and antimicrobial activity of six species of edible Irish seaweeds. Int. Food Res. J. 2010, 17, 205–220. [Google Scholar]
  38. Tang, J.; Wang, W.; Chu, W. Antimicrobial and Anti-Quorum Sensing Activities of Phlorotannins From Seaweed (Hizikia fusiforme). Front. Cell. Infect. Microbiol. 2020, 10, 586750. [Google Scholar] [CrossRef]
  39. Eom, S.-H.; Kim, D.-H.; Lee, S.-H.; Yoon, N.-Y.; Kim, J.H.; Kim, T.H.; Chung, Y.-H.; Kim, S.-B.; Kim, Y.-M.; Kim, H.-W.; et al. In vitro Antibacterial Activity and Synergistic Antibiotic Effects of Phlorotannins Isolated from Eisenia bicyclis Against Methicillin-Resistant Staphylococcus aureus. Phytother. Res. 2013, 27, 1260–1264. [Google Scholar] [CrossRef]
  40. Choi, J.-S.; Lee, K.; Lee, B.-B.; Kim, Y.-C.; Kim, Y.D.; Hong, Y.-K.; Cho, K.K.; Choi, I.S. Antibacterial activity of the phlorotannins dieckol and phlorofucofuroeckol-A from Ecklonia cava against Propionibacterium acnes. Bot. Sci. 2014, 92, 425–431. [Google Scholar] [CrossRef]
  41. Lee, J.-H.; Eom, S.-H.; Lee, E.-H.; Jung, Y.-J.; Kim, H.-J.; Jo, M.-R.; Son, K.-T.; Lee, H.-J.; Kim, J.H.; Lee, M.-S.; et al. In vitro antibacterial and synergistic effect of phlorotannins isolated from edible brown seaweed Eisenia bicyclis against acne-related bacteria. Algae 2014, 29, 47–55. [Google Scholar] [CrossRef]
  42. Choi, J.G.; Kang, O.H.; Brice, O.O.; Lee, Y.S.; Chae, H.S.; Oh, Y.C.; Sohn, D.H.; Park, H.; Choi, H.G.; Kim, S.G.; et al. Antibacterial activity of Ecklonia cava against methicillin-resistant Staphylococcus aureus and Salmonella spp. Foodborne Pathog. Dis. 2010, 7, 435–441. [Google Scholar] [CrossRef]
  43. Mittal, N.; Tesfu, H.H.; Hogan, A.M.; Cardona, S.T.; Sorensen, J.L. Synthesis and antibiotic activity of novel acylated phloroglucinol compounds against methicillin-resistant Staphylococcus aureus. J. Antibiot. 2019, 72, 253–259. [Google Scholar] [CrossRef] [PubMed]
  44. Kim, J.-H.; Kim, S.-B.; Hwang, H.-J.; Kim, Y.-M.; Lee, M.-S. Antibacterial Property of Ecklonia cava Extract against Marine Bacterial Pathogens. J. Food Hyg. Saf. 2016, 31, 380–385. [Google Scholar] [CrossRef]
  45. González-Colunga, D.; Antunes-Ricardo, M.; Gutiérrez-Uribe, J.A.; Cruz-Suárez, L.E. Bioactivity-guided identification of anti-AHPND (acute hepatopancreatic necrosis disease) metabolites of Ecklonia arborea. J. Appl. Phycol. 2019, 31, 3189–3199. [Google Scholar] [CrossRef]
  46. Lee, D.-S.; Kang, M.-S.; Hwang, H.-J.; Eom, S.-H.; Yang, J.-Y.; Lee, M.-S.; Lee, W.-J.; Jeon, Y.-J.; Choi, J.-S.; Kim, Y.-M. Synergistic effect between dieckol from Ecklonia stolonifera and β-lactams against methicillin-resistant Staphylococcus aureus. Biotechnol. Bioprocess Eng. 2008, 13, 758–764. [Google Scholar] [CrossRef]
  47. Wei, Y.; Liu, Q.; Xu, C.; Yu, J.; Zhao, L.; Guo, Q. Damage to the Membrane Permeability and Cell Death of Vibrio parahaemolyticus Caused by Phlorotannins with Low Molecular Weight from Sargassum thunbergii. J. Aquat. Food Prod. Technol. 2016, 25, 323–333. [Google Scholar] [CrossRef]
  48. Hierholtzer, A.; Chatellard, L.; Kierans, M.; Akunna, J.C.; Collier, P.J. The impact and mode of action of phenolic compounds extracted from brown seaweed on mixed anaerobic microbial cultures. J. Appl. Microbiol. 2013, 114, 964–973. [Google Scholar] [CrossRef]
  49. Lee, M.H.; Lee, K.B.; Oh, S.M.; Lee, B.H.; Chee, H.Y. Antifungal activities of dieckol isolated from the marine brown alga Ecklonia cava against Trichophyton rubrum. J. Korean Soc. Appl. Biol. Chem. 2010, 53, 504–507. [Google Scholar] [CrossRef]
  50. Kim, K.-H.; Yu, D.; Eom, S.-H.; Kim, H.-J.; Kim, D.-H.; Song, H.-S.; Kim, D.-M.; Kim, Y.-M. Fucofuroeckol-A from edible marine alga Eisenia bicyclis to restore antifungal activity of fluconazole against fluconazole-resistant Candida albicans. J. Appl. Phycol. 2018, 30, 605–609. [Google Scholar] [CrossRef]
  51. Balakrishnan, D.; Kandasamy, D.; Nithyanand, P. A review on antioxidant activity of marine organisms. Int. J. Chem. Technol. Res. 2014, 6, 3431–3436. [Google Scholar]
  52. Ma, L.; Yao, L.J.M. Antiviral effects of plant-derived essential oils and their components: An updated review. Molecules 2020, 25, 2627. [Google Scholar] [CrossRef]
  53. Kausar, S.; Said Khan, F.; Ishaq Mujeeb Ur Rehman, M.; Akram, M.; Riaz, M.; Rasool, G.; Hamid Khan, A.; Saleem, I.; Shamim, S.; Malik, A. A review: Mechanism of action of antiviral drugs. Int. J. Immunopathol. Pharmacol. 2021, 35. [Google Scholar] [CrossRef] [PubMed]
  54. Riccio, G.; Ruocco, N.; Mutalipassi, M.; Costantini, M.; Zupo, V.; Coppola, D.; de Pascale, D.; Lauritano, C.J.B. Ten-year research update review: Antiviral activities from marine organisms. Biomolecules 2020, 10, 1007. [Google Scholar] [CrossRef] [PubMed]
  55. Besednova, N.N.; Andryukov, B.G.; Zaporozhets, T.S.; Kryzhanovsky, S.P.; Fedyanina, L.N.; Kuznetsova, T.A.; Zvyagintseva, T.N.; Shchelkanov, M.Y. Antiviral effects of polyphenols from marine algae. Biomolecules 2021, 9, 200. [Google Scholar] [CrossRef] [PubMed]
  56. Sansone, C.; Brunet, C.; Noonan, D.M.; Albini, A.J.A. Marine algal antioxidants as potential vectors for controlling viral diseases. Antioxidants 2020, 9, 392. [Google Scholar] [CrossRef]
  57. Karadeniz, F.; Kang, K.-H.; Park, J.W.; Park, S.-J.; Kim, S.-K. Anti-HIV-1 activity of phlorotannin derivative 8,4‴-dieckol from Korean brown alga Ecklonia cava. Biosci. Biotechnol. Biochem. 2014, 78, 1151–1158. [Google Scholar] [CrossRef]
  58. Ahn, M.J.; Yoon, K.D.; Min, S.Y.; Lee, J.S.; Kim, J.H.; Kim, T.G.; Kim, S.H.; Kim, N.G.; Huh, H.; Kim, J. Inhibition of HIV-1 reverse transcriptase and protease by phlorotannins from the brown alga Ecklonia cava. Biol. Pharm. Bull. 2004, 27, 544–547. [Google Scholar] [CrossRef] [Green Version]
  59. Ryu, Y.B.; Jeong, H.J.; Yoon, S.Y.; Park, J.Y.; Kim, Y.M.; Park, S.J.; Rho, M.C.; Kim, S.J.; Lee, W.S. Influenza virus neuraminidase inhibitory activity of phlorotannins from the edible brown alga Ecklonia cava. J. Agric. Food Chem. 2011, 59, 6467–6473. [Google Scholar] [CrossRef]
  60. Afroz, M.; Zihad, S.N.K.; Uddin, S.J.; Rouf, R.; Rahman, M.S.; Islam, M.T.; Khan, I.N.; Ali, E.S.; Aziz, S.; Shilpi, J.A.; et al. A systematic review on antioxidant and antiinflammatory activity of Sesame (Sesamum indicum L.) oil and further confirmation of antiinflammatory activity by chemical profiling and molecular docking. Phytother. Res. 2019, 33, 2585–2608. [Google Scholar] [CrossRef]
  61. Beg, S.; Swain, S.; Hasan, H.; Barkat, M.A.; Hussain, M.S.J.P.R. Systematic review of herbals as potential anti-inflammatory agents: Recent advances, current clinical status and future perspectives. Pharmacogn. Rev. 2011, 5, 120. [Google Scholar] [CrossRef] [Green Version]
  62. Patel, S.S.; Savjani, J.K. Systematic review of plant steroids as potential antiinflammatory agents: Current status and future perspectives. J. Phytopharm. 2015, 4, 121–125. [Google Scholar] [CrossRef]
  63. Barbosa, M.; Lopes, G.; Andrade, P.B.; Valentão, P. Technology, Bioprospecting of brown seaweeds for biotechnological applications: Phlorotannin actions in inflammation and allergy network. Trends Food Sci. Technol. 2019, 86, 153–171. [Google Scholar] [CrossRef]
  64. Jung, H.A.; Jin, S.E.; Ahn, B.R.; Lee, C.M.; Choi, J.S. Anti-inflammatory activity of edible brown alga Eisenia bicyclis and its constituents fucosterol and phlorotannins in LPS-stimulated RAW264.7 macrophages. Food Chem. Toxicol. 2013, 59, 199–206. [Google Scholar] [CrossRef] [PubMed]
  65. Barbosa, M.; Lopes, G.; Ferreres, F.; Andrade, P.B.; Pereira, D.M.; Gil-Izquierdo, Á.; Valentão, P. Phlorotannin extracts from Fucales: Marine polyphenols as bioregulators engaged in inflammation-related mediators and enzymes. Algal Res. 2017, 28, 1–8. [Google Scholar] [CrossRef]
  66. Kim, M.-M.; Kim, S.-K. Effect of phloroglucinol on oxidative stress and inflammation. Food Chem. Toxicol. 2010, 48, 2925–2933. [Google Scholar] [CrossRef] [PubMed]
  67. Joung, E.J.; Lee, M.S.; Choi, J.W.; Kim, J.S.; Shin, T.; Jung, B.M.; Kim, J.I.; Kim, H.R. Anti-inflammatory effects of phlorofucofuroeckol B-rich ethyl acetate fraction obtained from Myagropsis myagroides on lipopolysaccharide-stimulated RAW 264.7 cells and mouse edema. Int. Immunopharmacol. 2012, 14, 471–480. [Google Scholar] [CrossRef] [PubMed]
  68. Gonçalves-Fernández, C.; Sineiro, J.; Moreira, R.; Gualillo, O. Extraction and characterization of phlorotannin-enriched fractions from the Atlantic seaweed Bifurcaria bifurcata and evaluation of their cytotoxic activity in murine cell line. J. Appl. Phycol. 2019, 31, 2573–2583. [Google Scholar] [CrossRef]
  69. Nair, D.; Vanuopadath, M.; Balasubramanian, A.; Iyer, A.; Ganesh, S.; Anil, A.N.; Vikraman, V.; Pillai, P.; Bose, C.; Nair, B.G.; et al. Phlorotannins from Padina tetrastromatica: Structural characterisation and functional studies. J. Appl. Phycol. 2019, 31, 3131–3141. [Google Scholar] [CrossRef]
  70. Yu, D.-K.; Lee, B.; Kwon, M.; Yoon, N.; Shin, T.; Kim, N.-G.; Choi, J.-S.; Kim, H.-R. Phlorofucofuroeckol B suppresses inflammatory responses by down-regulating nuclear factor κB activation via Akt, ERK, and JNK in LPS-stimulated microglial cells. Int. Immunopharmacol. 2015, 28, 1068–1075. [Google Scholar] [CrossRef]
  71. Yang, Y.-I.; Shin, H.-C.; Kim, S.H.; Park, W.-Y.; Lee, K.-T.; Choi, J.-H. 6,6′-Bieckol, isolated from marine alga Ecklonia cava, suppressed LPS-induced nitric oxide and PGE2 production and inflammatory cytokine expression in macrophages: The inhibition of NFκB. Int. Immunopharmacol. 2012, 12, 510–517. [Google Scholar] [CrossRef]
  72. Hamed, I.; Özogul, F.; Özogul, Y.; Regenstein, J.M. Marine bioactive compounds and their health benefits: A review. Compr. Rev. Food Sci. Food Saf. 2015, 14, 446–465. [Google Scholar] [CrossRef]
  73. Bratchkova, A.; Kroumov, A.D. Microalgae as producers of biologically active compounds with antibacterial, antiviral, antifungal, antialgal, antiprotozoal, antiparasitic and anticancer activity. Acta Microbiol. Bulg. 2020, 36, 79–89. [Google Scholar]
  74. Manganyi, M.C.; Ateba, C.N.J.M. Untapped potentials of endophytic fungi: A review of novel bioactive compounds with biological applications. Microorganisms 2020, 8, 1934. [Google Scholar] [CrossRef] [PubMed]
  75. Jose, G.M. Biological Responses of Algal Derived Sulfated Polysaccharides: An Emphasis on Cancer Prophylaxis. Trends Biomater. Artif. Organs 2015, 29, 64–85. [Google Scholar]
  76. Hussain, E.; Wang, L.-J.; Jiang, B.; Riaz, S.; Butt, G.Y.; Shi, D.-Y. A review of the components of brown seaweeds as potential candidates in cancer therapy. RSC Adv. 2016, 6, 12592–12610. [Google Scholar] [CrossRef]
  77. Meng, W.; Mu, T.; Sun, H.; Garcia-Vaquero, M. Phlorotannins: A review of extraction methods, structural characteristics, bioactivities, bioavailability, and future trends. Algal Res. 2021, 60, 102484. [Google Scholar] [CrossRef]
  78. Ahn, J.-H.; Yang, Y.-I.; Lee, K.-T.; Choi, J.-H. Dieckol, isolated from the edible brown algae Ecklonia cava, induces apoptosis of ovarian cancer cells and inhibits tumor xenograft growth. J. Cancer Res. Clin. Oncol. 2015, 141, 255–268. [Google Scholar] [CrossRef] [PubMed]
  79. Kong, C.-S.; Kim, J.-A.; Yoon, N.-Y.; Kim, S.-K. Induction of apoptosis by phloroglucinol derivative from Ecklonia Cava in MCF-7 human breast cancer cells. Food Chem. Toxicol. 2009, 47, 1653–1658. [Google Scholar] [CrossRef]
  80. Kim, R.-K.; Suh, Y.; Yoo, K.-C.; Cui, Y.-H.; Hwang, E.; Kim, H.-J.; Kang, J.-S.; Kim, M.-J.; Lee, Y.Y.; Lee, S.-J. Phloroglucinol suppresses metastatic ability of breast cancer cells by inhibition of epithelial-mesenchymal cell transition. Cancer Sci. 2015, 106, 94–101. [Google Scholar] [CrossRef] [Green Version]
  81. Kang, M.-H.; Kim, I.-H.; Nam, T.-J. Phloroglucinol induces apoptosis via apoptotic signaling pathways in HT-29 colon cancer cells. Oncol. Rep. 2014, 32, 1341–1346. [Google Scholar] [CrossRef] [Green Version]
  82. Abdelhamid, A.; Lajili, S.; Elkaibi, M.A.; Ben Salem, Y.; Abdelhamid, A.; Muller, C.D.; Majdoub, H.; Kraiem, J.; Bouraoui, A. Optimized Extraction, Preliminary Characterization and Evaluation of the in vitro Anticancer Activity of Phlorotannin-Rich Fraction from the Brown Seaweed, Cystoseira sedoides. J. Aquat. Food Prod. Technol. 2019, 28, 892–909. [Google Scholar] [CrossRef]
  83. Sadeeshkumar, V.; Duraikannu, A.; Ravichandran, S.; Fredrick, W.S.; Sivaperumal, R.; Kodisundaram, P. Protective effects of dieckol on N-nitrosodiethylamine induced hepatocarcinogenesis in rats. Biomed. Pharmacother. 2016, 84, 1810–1819. [Google Scholar] [CrossRef] [PubMed]
  84. Antolovich, M.; Prenzler, P.D.; Patsalides, E.; McDonald, S.; Robards, K. Methods for testing antioxidant activity. Analyst 2002, 127, 183–198. [Google Scholar] [CrossRef] [PubMed]
  85. Fubini, B.; Hubbard, A. Reactive oxygen species (ROS) and reactive nitrogen species (RNS) generation by silica in inflammation and fibrosis. Free. Radic. Biol. Med. 2003, 34, 1507–1516. [Google Scholar] [CrossRef]
  86. Nigam, V.; Sodhi, J.S. Some medicinal plants with antioxidant activity—A review. Int. J. Pharm. Biol. Sci. 2014, 4, 173–178. [Google Scholar]
  87. Diniz do Nascimento, L.; De Moraes, A.A.B.; Da Costa, K.S.; Galúcio, J.M.P.; Taube, P.S.; Costa, C.M.L.; Cruz, J.N.; de Aguiar Andrade, E.H.; De Faria, L.J.G. Bioactive Natural Compounds and Antioxidant Activity of Essential Oils from Spice Plants: New Findings and Potential Applications. Biomolecules 2020, 10, 988. [Google Scholar] [CrossRef] [PubMed]
  88. Chaves, N.; Santiago, A.; Alías, J.C. Quantification of the Antioxidant Activity of Plant Extracts: Analysis of Sensitivity and Hierarchization Based on the Method Used. Antioxidants 2020, 9, 76. [Google Scholar] [CrossRef] [Green Version]
  89. Sathya, R.; Kanaga, N.; Sankar, P.; Jeeva, S. Antioxidant properties of phlorotannins from brown seaweed Cystoseira trinodis (Forsskål) C. Agardh. Arab. J. Chem. 2017, 10, S2608–S2614. [Google Scholar] [CrossRef] [Green Version]
  90. Yotsu-Yamashita, M.; Kondo, S.; Segawa, S.; Lin, Y.-C.; Toyohara, H.; Ito, H.; Konoki, K.; Cho, Y.; Uchida, T. Isolation and Structural Determination of Two Novel Phlorotannins from the Brown Alga Ecklonia kurome Okamura, and Their Radical Scavenging Activities. Marine Drugs 2013, 11, 165–183. [Google Scholar] [CrossRef] [Green Version]
  91. Ahn, G.-N.; Kim, K.-N.; Cha, S.-H.; Song, C.-B.; Lee, J.; Heo, M.-S.; Yeo, I.-K.; Lee, N.-H.; Jee, Y.-H.; Kim, J.-S.; et al. Antioxidant activities of phlorotannins purified from Ecklonia cava on free radical scavenging using ESR and H2O2-mediated DNA damage. Eur. Food Res. Technol. 2007, 226, 71–79. [Google Scholar] [CrossRef]
  92. Nakai, M.; Kageyama, N.; Nakahara, K.; Miki, W. Phlorotannins as Radical Scavengers from the Extract of Sargassum ringgoldianum. Mar. Biotechnol. 2006, 8, 409–414. [Google Scholar] [CrossRef]
  93. Boi, V.N.; Trang, N.T.M.; Cuong, D.X.; Ha, H.T. Antioxidant Phlorotannin from Brown Algae Sargassum dupplicatum: Enzyme-assissted Extraction and Purification. World J. Food Sci. Technol. 2020, 4, 62–68. [Google Scholar] [CrossRef]
  94. Lee, J.-H.; Ko, J.-Y.; Oh, J.-Y.; Kim, E.-A.; Kim, C.-Y.; Jeon, Y.-J. Evaluation of phlorofucofuroeckol-A isolated from Ecklonia cava (Phaeophyta) on anti-lipid peroxidation in vitro and in vivo. Algae 2015, 30, 313–323. [Google Scholar] [CrossRef] [Green Version]
  95. Kang, M.-C.; Cha, S.H.; Wijesinghe, W.A.J.P.; Kang, S.-M.; Lee, S.-H.; Kim, E.-A.; Song, C.B.; Jeon, Y.-J. Protective effect of marine algae phlorotannins against AAPH-induced oxidative stress in zebrafish embryo. Food Chem. 2013, 138, 950–955. [Google Scholar] [CrossRef] [PubMed]
  96. Kang, S.-M.; Cha, S.-H.; Ko, J.-Y.; Kang, M.-C.; Kim, D.; Heo, S.-J.; Kim, J.-S.; Heu, M.S.; Kim, Y.-T.; Jung, W.-K.; et al. Neuroprotective effects of phlorotannins isolated from a brown alga, Ecklonia cava, against H2O2-induced oxidative stress in murine hippocampal HT22 cells. Environ. Toxicol. Pharmacol. 2012, 34, 96–105. [Google Scholar] [CrossRef]
  97. Kang, X.; Liang, H.; Luo, Y.; Li, Z.; He, F.; Han, X.; Zhang, L. Anti-adipogenesis and metabolism-regulating effects of heat-inactivated Streptococcus thermophilus MN-ZLW-002. Lett. Appl. Microbiol. 2021, 72, 677–687. [Google Scholar] [CrossRef]
  98. Blüher, M. Obesity: Global epidemiology and pathogenesis. Nat. Rev. Endocrinol. 2019, 15, 288–298. [Google Scholar] [CrossRef]
  99. Tsai, Y.-C.; Yang, B.-C.; Peng, W.-H.; Lee, Y.-M.; Yen, M.-H.; Cheng, P.-Y. Heme oxygenase-1 mediates anti-adipogenesis effect of raspberry ketone in 3T3-L1 cells. Phytomedicine 2017, 31, 11–17. [Google Scholar] [CrossRef]
  100. Jung, H.A.; Jung, H.J.; Jeong, H.Y.; Kwon, H.J.; Ali, M.Y.; Choi, J.S. Phlorotannins isolated from the edible brown alga Ecklonia stolonifera exert anti-adipogenic activity on 3T3-L1 adipocytes by downregulating C/EBPα and PPARγ. Fitoterapia 2014, 92, 260–269. [Google Scholar] [CrossRef]
  101. Karadeniz, F.; Ahn, B.-N.; Kim, J.; Seo, Y.; Jang, M.-S.; Nam, K.-H.; Kim, M.; Lee, S.-H.; Kong, C.-S. Phlorotannins suppress adipogenesis in pre-adipocytes while enhancing osteoblastogenesis in pre-osteoblasts. Arch. Pharmacal Res. 2015, 38, 2172–2182. [Google Scholar] [CrossRef]
  102. Seo, Y.-J.; Kim, K.-J.; Koh, E.-J.; Choi, J.; Lee, B.-Y. Anti-adipogenesis mechanism of pterostilbene through the activation of heme oxygenase-1 in 3T3-L1 cells. Phytomedicine 2017, 33, 7–13. [Google Scholar] [CrossRef]
  103. Guo, L.; Li, K.; Kang, J.S.; Kang, N.J.; Son, B.G.; Choi, Y.W. Strawberry fermentation with Cordyceps militaris has anti-adipogenesis activity. Food Biosci. 2020, 35, 100576. [Google Scholar] [CrossRef]
  104. Kong, C.S.; Kim, H.; Seo, Y. Edible Brown Alga Ecklonia cava Derived Phlorotannin-Induced Anti-Adipogenic Activity in vitro. J. Food Biochem. 2015, 39, 1–10. [Google Scholar] [CrossRef]
  105. Hu, X.; Tao, N.; Wang, X.; Xiao, J.; Wang, M. Marine-derived bioactive compounds with anti-obesity effect: A review. J. Funct. Foods 2016, 21, 372–387. [Google Scholar] [CrossRef]
  106. Eom, S.H.; Lee, S.H.; Yoon, N.Y.; Jung, W.K.; Jeon, Y.J.; Kim, S.K.; Lee, M.S.; Kim, Y.M. α-Glucosidase-and α-amylase-inhibitory activities of phlorotannins from Eisenia bicyclis. J. Sci. Food Agric. 2012, 92, 2084–2090. [Google Scholar] [CrossRef]
  107. Lee, S.-H.; Yong, L.; Karadeniz, F.; Kim, M.-M.; Kim, S.-K. α-Glucosidase and α-amylase inhibitory activities of phloroglucinal derivatives from edible marine brown alga, Ecklonia cava. J. Sci. Food Agric. 2009, 89, 1552–1558. [Google Scholar] [CrossRef]
  108. Gheda, S.; Naby, M.A.; Mohamed, T.; Pereira, L.; Khamis, A. Antidiabetic and antioxidant activity of phlorotannins extracted from the brown seaweed Cystoseira compressa in streptozotocin-induced diabetic rats. Environ. Sci. Pollut. Res. 2021, 28, 22886–22901. [Google Scholar] [CrossRef]
  109. Catarino, M.D.; Silva, A.; Mateus, N.; Cardoso, S.M. Optimization of phlorotannins extraction from Fucus vesiculosus and evaluation of their potential to prevent metabolic disorders. Marine Drugs 2019, 17, 162. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Ryu, B.; Jiang, Y.; Kim, H.-S.; Hyun, J.-M.; Lim, S.-B.; Li, Y.; Jeon, Y.-J. Ishophloroglucin A, a novel phlorotannin for standardizing the anti-α-glucosidase activity of Ishige okamurae. Marine Drugs 2018, 16, 436. [Google Scholar] [CrossRef] [Green Version]
  111. You, H.-N.; Lee, H.-A.; Park, M.-H.; Lee, J.-H.; Han, J.-S. Phlorofucofuroeckol A isolated from Ecklonia cava alleviates postprandial hyperglycemia in diabetic mice. Eur. J. Pharmacol. 2015, 752, 92–96. [Google Scholar] [CrossRef]
  112. Kang, M.-C.; Wijesinghe, W.A.J.P.; Lee, S.-H.; Kang, S.-M.; Ko, S.-C.; Yang, X.; Kang, N.; Jeon, B.-T.; Kim, J.; Lee, D.-H.; et al. Dieckol isolated from brown seaweed Ecklonia cava attenuates type II diabetes in db/db mouse model. Food Chem. Toxicol. 2013, 53, 294–298. [Google Scholar] [CrossRef]
  113. Kim, S.-K.; Kong, C.-S. Anti-adipogenic effect of dioxinodehydroeckol via AMPK activation in 3T3-L1 adipocytes. Chem.-Biol. Interact. 2010, 186, 24–29. [Google Scholar] [CrossRef] [PubMed]
  114. Heo, S.-J.; Ko, S.-C.; Cha, S.-H.; Kang, D.-H.; Park, H.-S.; Choi, Y.-U.; Kim, D.; Jung, W.-K.; Jeon, Y.-J. Effect of phlorotannins isolated from Ecklonia cava on melanogenesis and their protective effect against photo-oxidative stress induced by UV-B radiation. Toxicol. Vitr. 2009, 23, 1123–1130. [Google Scholar] [CrossRef] [PubMed]
  115. Liu, Y.; Zhang, D.; Liu, G.-M.; Chen, Q.; Lu, Z. Ameliorative effect of dieckol-enriched extraction from Laminaria japonica on hepatic steatosis induced by a high-fat diet via β-oxidation pathway in ICR mice. J. Funct. Foods 2019, 58, 44–55. [Google Scholar] [CrossRef]
  116. Ko, S.-C.; Lee, M.; Lee, J.-H.; Lee, S.-H.; Lim, Y.; Jeon, Y.-J. Dieckol, a phlorotannin isolated from a brown seaweed, Ecklonia cava, inhibits adipogenesis through AMP-activated protein kinase (AMPK) activation in 3T3-L1 preadipocytes. Environ. Toxicol. Pharmacol. 2013, 36, 1253–1260. [Google Scholar] [CrossRef]
  117. Filippini, M.; Baldisserotto, A.; Menotta, S.; Fedrizzi, G.; Rubini, S.; Gigliotti, D.; Valpiani, G.; Buzzi, R.; Manfredini, S.; Vertuani, S. Heavy metals and potential risks in edible seaweed on the market in Italy. Chemosphere 2021, 263, 127983. [Google Scholar] [CrossRef] [PubMed]
  118. Rajaram, R.; Rameshkumar, S.; Anandkumar, A. Health risk assessment and potentiality of green seaweeds on bioaccumulation of trace elements along the Palk Bay coast, Southeastern India. Mar. Pollut. Bull. 2020, 154, 111069. [Google Scholar] [CrossRef] [PubMed]
  119. Chen, Q.; Pan, X.-D.; Huang, B.-F.; Han, J.-L. Distribution of metals and metalloids in dried seaweeds and health risk to population in southeastern China. Sci. Rep. 2018, 8, 3578. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Cassani, L.; Gomez-Zavaglia, A.; Jimenez-Lopez, C.; Lourenço-Lopes, C.; Prieto, M.A.; Simal-Gandara, J. Seaweed-based natural ingredients: Stability of phlorotannins during extraction, storage, passage through the gastrointestinal tract and potential incorporation into functional foods. Food Res. Int. 2020, 137, 109676. [Google Scholar] [CrossRef] [PubMed]
  121. Corona, G.; Ji, Y.; Anegboonlap, P.; Hotchkiss, S.; Gill, C.; Yaqoob, P.; Spencer, J.P.; Rowland, I. Gastrointestinal modifications and bioavailability of brown seaweed phlorotannins and effects on inflammatory markers. Br. J. Nutr. 2016, 115, 1240–1253. [Google Scholar] [CrossRef] [Green Version]
  122. Rajha, H.N.; Paule, A.; Aragonès, G.; Barbosa, M.; Caddeo, C.; Debs, E.; Dinkova, R.; Eckert, G.P.; Fontana, A.; Gebrayel, P. Recent Advances in Research on Polyphenols: Effects on Microbiota, Metabolism, and Health. Mol. Nutr. Food Res. 2022, 66, 2100670. [Google Scholar] [CrossRef]
  123. Rajan, D.K.; Mohan, K.; Zhang, S.; Ganesan, A.R. Dieckol: A brown algal phlorotannin with biological potential. Biomed. Pharmacother. 2021, 142, 111988. [Google Scholar] [CrossRef] [PubMed]
  124. Michalak, I.; Chojnacka, K. Algae as production systems of bioactive compounds. Eng. Life Sci. 2015, 15, 160–176. [Google Scholar] [CrossRef]
  125. Stengel, D.B.; Connan, S. Marine algae: A source of biomass for biotechnological applications. In Natural Products from Marine Algae; Springer: New York, NY, USA; Humana Press: New York, NY, USA, 2015; pp. 1–37. [Google Scholar]
  126. Dang, T.T.; Bowyer, M.C.; Van Altena, I.A.; Scarlett, C.J. Optimum conditions of microwave-assisted extraction for phenolic compounds and antioxidant capacity of the brown alga Sargassum vestitum. Sep. Sci. Technol. 2018, 53, 1711–1723. [Google Scholar] [CrossRef]
  127. Kadam, S.U.; Tiwari, B.K.; O’Donnell, C.P. Application of novel extraction technologies for bioactives from marine algae. J. Agric. Food Chem. 2013, 61, 4667–4675. [Google Scholar] [CrossRef] [PubMed]
  128. Thiyagarasaiyar, K.; Goh, B.-H.; Jeon, Y.-J.; Yow, Y.-Y. Algae metabolites in cosmeceutical: An overview of current applications and challenges. Mar. Drugs 2020, 18, 323. [Google Scholar] [CrossRef]
  129. Quitério, E.; Soares, C.; Ferraz, R.; Delerue-Matos, C.; Grosso, C. Marine Health-Promoting Compounds: Recent Trends for Their Characterization and Human Applications. Foods 2021, 10, 3100. [Google Scholar] [CrossRef]
  130. Corsetto, P.A.; Montorfano, G.; Zava, S.; Colombo, I.; Ingadottir, B.; Jonsdottir, R.; Sveinsdottir, K.; Rizzo, A.M. Characterization of antioxidant potential of seaweed extracts for enrichment of convenience food. Antioxidants 2020, 9, 249. [Google Scholar] [CrossRef] [Green Version]
  131. Santos, S.A.; Félix, R.; Pais, A.; Rocha, S.M.; Silvestre, A.J. The quest for phenolic compounds from macroalgae: A review of extraction and identification methodologies. Biomolecules 2019, 9, 847. [Google Scholar] [CrossRef] [Green Version]
  132. Brglez Mojzer, E.; Knez Hrnčič, M.; Škerget, M.; Knez, Ž.; Bren, U. Polyphenols: Extraction methods, antioxidative action, bioavailability and anticarcinogenic effects. Molecules 2016, 21, 901. [Google Scholar] [CrossRef]
  133. Buedenbender, L.; Astone, F.A.; Tasdemir, D. Bioactive molecular networking for mapping the antimicrobial constituents of the baltic brown alga fucus vesiculosus. Mar. Drugs 2020, 18, 311. [Google Scholar] [CrossRef]
  134. Al-Mola, H.F. Antibacterial activity of crude extracts and phlorotannin isolated from the diatom Cymbella spp. J. Pharm. Res 2009, 2, 304–308. [Google Scholar]
  135. Kang, M.-H.; Kim, I.-H.; Nam, T.-J. Phloroglucinol induces apoptosis through the regulation of insulin-like growth factor 1 receptor signaling pathways in human colon cancer HT-29 cells. Int. J. Oncol. 2014, 45, 1036–1042. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Vizetto-Duarte, C.; Custódio, L.; Gangadhar, K.N.; Lago, J.H.; Dias, C.; Matos, A.M.; Neng, N.; Nogueira, J.M.; Barreira, L.; Albericio, F.; et al. Isololiolide, a carotenoid metabolite isolated from the brown alga Cystoseira tamariscifolia, is cytotoxic and able to induce apoptosis in hepatocarcinoma cells through caspase-3 activation, decreased Bcl-2 levels, increased p53 expression and PARP cleavage. Phytomedicine 2016, 23, 550–557. [Google Scholar] [PubMed]
  137. Kang, K.A.; Lee, K.H.; Chae, S.; Zhang, R.; Jung, M.S.; Ham, Y.M.; Baik, J.S.; Lee, N.H.; Hyun, J.W. Cytoprotective effect of phloroglucinol on oxidative stress induced cell damage via catalase activation. J. Cell. Biochem. 2006, 97, 609–620. [Google Scholar] [CrossRef]
  138. Kang, K.A.; Zhang, R.; Chae, S.; Lee, S.J.; Kim, J.; Kim, J.; Jeong, J.; Lee, J.; Shin, T.; Lee, N.H.; et al. Phloroglucinol (1,3,5-trihydroxybenzene) protects against ionizing radiation-induced cell damage through inhibition of oxidative stress in vitro and in vivo. Chem.-Biol. Interact. 2010, 185, 215–226. [Google Scholar] [CrossRef]
  139. Piao, M.J.; Zhang, R.; Lee, N.H.; Hyun, J.W. Phloroglucinol Attenuates Ultraviolet B Radiation-Induced Matrix Metalloproteinase-1 Production in Human Keratinocytes via Inhibitory Actions against Mitogen-Activated Protein Kinases and Activator Protein-1. Photochem. Photobiol. 2012, 88, 381–388. [Google Scholar] [CrossRef]
  140. Shrestha, S.; Johnston, M.R.; Zhang, W.; Smid, S.D. A phlorotannin isolated from Ecklonia radiata, Dibenzodioxin-fucodiphloroethol, inhibits neurotoxicity and aggregation of β-amyloid. Phytomed. Plus 2021, 1, 100125. [Google Scholar] [CrossRef]
  141. Ryu, B.; Ahn, B.-N.; Kang, K.-H.; Kim, Y.-S.; Li, Y.-X.; Kong, C.-S.; Kim, S.-K.; Kim, D.G. Dioxinodehydroeckol protects human keratinocyte cells from UVB-induced apoptosis modulated by related genes Bax/Bcl-2 and caspase pathway. J. Photochem. Photobiol. B Biol. 2015, 153, 352–357. [Google Scholar] [CrossRef]
  142. Li, Y.X.; Li, Y.; Je, J.Y.; Kim, S.K. Dieckol as a novel anti-proliferative and anti-angiogenic agent and computational anti-angiogenic activity evaluation. Environ. Toxicol. Pharmacol. 2015, 39, 259–270. [Google Scholar] [CrossRef]
  143. Khan, F.; Oh, D.; Chandika, P.; Jo, D.M.; Bamunarachchi, N.I.; Jung, W.K.; Kim, Y.M. Inhibitory activities of phloroglucinol-chitosan nanoparticles on mono- and dual-species biofilms of Candida albicans and bacteria. Colloids Surf. B Biointerfaces 2022, 211, 112307. [Google Scholar] [CrossRef]
Figure 1. Chemical structures of phlorotannins.
Figure 1. Chemical structures of phlorotannins.
Marinedrugs 20 00384 g001
Figure 2. Mechanism of antibacterial and antifungal activities of different types of phlorotannins.
Figure 2. Mechanism of antibacterial and antifungal activities of different types of phlorotannins.
Marinedrugs 20 00384 g002
Figure 3. Mechanism of antiviral activities of phlorotannins.
Figure 3. Mechanism of antiviral activities of phlorotannins.
Marinedrugs 20 00384 g003
Figure 4. Mechanisms of anti-inflammatory activity of several types of phlorotannins.
Figure 4. Mechanisms of anti-inflammatory activity of several types of phlorotannins.
Marinedrugs 20 00384 g004
Figure 5. The anticancer activities of different types of phlorotannins.
Figure 5. The anticancer activities of different types of phlorotannins.
Marinedrugs 20 00384 g005
Figure 6. Phlorotannins with anti-adipogenesis, antidiabetic, and anti-obesity activities.
Figure 6. Phlorotannins with anti-adipogenesis, antidiabetic, and anti-obesity activities.
Marinedrugs 20 00384 g006
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Khan, F.; Jeong, G.-J.; Khan, M.S.A.; Tabassum, N.; Kim, Y.-M. Seaweed-Derived Phlorotannins: A Review of Multiple Biological Roles and Action Mechanisms. Mar. Drugs 2022, 20, 384. https://doi.org/10.3390/md20060384

AMA Style

Khan F, Jeong G-J, Khan MSA, Tabassum N, Kim Y-M. Seaweed-Derived Phlorotannins: A Review of Multiple Biological Roles and Action Mechanisms. Marine Drugs. 2022; 20(6):384. https://doi.org/10.3390/md20060384

Chicago/Turabian Style

Khan, Fazlurrahman, Geum-Jae Jeong, Mohd Sajjad Ahmad Khan, Nazia Tabassum, and Young-Mog Kim. 2022. "Seaweed-Derived Phlorotannins: A Review of Multiple Biological Roles and Action Mechanisms" Marine Drugs 20, no. 6: 384. https://doi.org/10.3390/md20060384

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

Article Metrics

Back to TopTop