Next Article in Journal
Research on Dynamic Reserve and Energy Arbitrage of Energy Storage System
Next Article in Special Issue
Cosmeceuticals and Thalassotherapy: Recovering the Skin and Well-Being after Cancer Therapies
Previous Article in Journal
Time Series Prediction of Dam Deformation Using a Hybrid STL–CNN–GRU Model Based on Sparrow Search Algorithm Optimization
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Dermo-Cosmetic Benefits of Marine Macroalgae-Derived Phenolic Compounds

by
Haresh S. Kalasariya
1 and
Leonel Pereira
2,*
1
Centre for Natural Products Discovery, School of Pharmacy and Biomolecular Sciences, Liverpool John Moores University, Byrom Street, Liverpool L3 3AF, UK
2
MARE—Marine and Environmental Sciences Centre, Department of Life Sciences, University of Coimbra, Calçada Martim de Freitas, 3000-456 Coimbra, Portugal
*
Author to whom correspondence should be addressed.
Appl. Sci. 2022, 12(23), 11954; https://doi.org/10.3390/app122311954
Submission received: 19 October 2022 / Revised: 14 November 2022 / Accepted: 18 November 2022 / Published: 23 November 2022
(This article belongs to the Special Issue Recent Developments and Emerging Trends in Marine Biotechnology)

Abstract

:
Marine macroalgae have an interesting profile of bioactive compounds and have gained tremendous attention in cosmeceuticals with negligible toxicity effects (cytotoxicity, reproductive toxicity, genotoxicity, mutagenicity, carcinogenicity, etc.) on humans and exhibit strong benefits for the skin. Among the diversified compounds, phenolic compounds are the group of phytochemicals found in high amounts with great structural diversity. Phlorotannin is the most studied polyphenol compound in brown algae, but besides there are some other phenolic compounds observed and studied in macroalgae such as terpenoids, bromophenols, mycosporine amino acids (MAAs), and flavonoids. These compounds are already characterized and studied for their full range of cosmeceutical benefits such as skin whitening, moisturizing, photoprotection, antiaging, antiwrinkle, anti-melanogenic, and antioxidant activities as well as in the treatment of pruritus (caused by acne, eczema, dermatitis, hives, psoriasis), photoaging, and skin pigmentation disorders (hypopigmentation due to the absence of melanocytes and hyperpigmentation caused by skin irritation or metabolic disorders). This review study mainly focuses on marine algae-derived phenolic compounds and their extraction, characterization, and skin cosmetic benefits described in the literature. The present study aims to provide a detailed insight into the phenolic compounds in marine algae.

1. Introduction

Cosmeceutical ingredients are active compounds that are used to improve the appearance of the human body and represent a new category of preparations placed between cosmetics and pharmaceuticals. Cosmeceutical formulations intend the improvements of skin health and beauty [1,2,3]. Globally, the cosmeceutical sector is growing each year due to increasing modern beauty trends. To meet consumer demand, industries are moving towards the excessive use of synthetic cosmetic ingredients in formulations listed as Hydroquinone (HQ), Phthalates, Para-aminobenzoic acid (PABA), Benzophenones, Butylated Hydroxyanisole (BHA), Butylated Hydroxytoluene (BHT), and Dibenzoylmethane (DBM). According to SCCS (Scientific Committee on Consumer Safety) opinion (SCCS/1564/15), the excessive use of synthetic ingredients in cosmeceutical formulations may lead to different types of toxicities such as acute toxicity, corrosion and irritation, skin sensation, dermal/percutaneous absorption, repeated dose toxicity, reproductive toxicity, mutagenicity/genotoxicity, carcinogenicity, and photoinduced toxicity on the skin as well as human health. Hydroxyanisole, widely used in skin-whitening creams, has reported many harmful effects such as ochronosis and potential mutagenicity [4,5,6]. Benzophenones, DBM, and PABA have shown allergic phototoxicities, dermatitis, and skin irritations [7,8]. Besides, BHA and BHT are applied in moisturization and lipstick preparations that cause allergic reactions, irritation, and corrosivity in the skin. Another ingredient, parabens, are highly carcinogenic and neurotoxic among other harmful health effects. Around 75 to 90 percent of commercially available products contain parabens, which are mostly used as a mixture in cosmetic formulations. Parabens have been reported to have a high risk of breast cancer and the development of malignant melanoma in women [9]. However, in the ACDS Contact Allergen Management Program (CAMP) report, about 19% of products contained different types of parabens, mainly methylparaben, ethylparaben, propylparaben, and butylparaben. According to them, these components have little allergenicity compared to other preservatives, with no adverse reactions, and low toxicity, safety, and cost [10]. Hafeez and Maibach [11] reported fewer sensitizing effects of parabens in commercial applications but very limited reports are often attributable to the use of parabens on damaged skin. Polyethylene glycol (PEG) is a genotoxic compound that irritates and causes systemic toxicities and skin damage. In skin cosmetics, PEGs function in three ways: as emollients (that soften and lubricate the skin), as emulsifiers (that help to mix water-based and oil-based ingredients properly), and as vehicles (that deliver ingredients deeper into the skin). In addition, the Agency for Toxic Substances and Disease Registry (ATSDR) Information Center, and the Centers for Disease Control and Prevention (CDC) reported the toxicity of dibutyl phthalate (DBP) in DNA damage in male reproductive cells [12]. Some previous studies have reported the harmful effects of cosmetic ingredients in animal studies, such as male genitalia disabilities that altered pregnancy outcomes as well as reduced sperm counts [13,14]. Moreover, the EC 1223/2009 regulation regarding the testing and marketing ban of cosmetic products suggests the prohibition of testing finished cosmetic products on animals and their marketing.
To overcome the toxicities of these formulations, consumers have changed their preference to natural skin care products in the last few years. As a result, industries have moved towards natural bioactive ingredients from various natural resources that are eco-friendly and less toxic [15,16]. Various natural resources can be used in skin cosmetic products such as terrestrial plants, fungi, marine algae, bacteria, animals, etc. [17,18,19,20,21]. Among them, marine macroalgae are widely utilized for their skin benefits nowadays. Marine macroalgae are also known as seaweed: eukaryotic, aquatic photosynthetic macroscopic, multicellular organisms that are ubiquitously found along the seacoast and in seawater. They belong to the Eukaryota domain and are classified into three major taxonomic groups, red algae, brown algae, and green algae, belonging to the Rhodophyta phylum, Ochrophyta phylum, and Chlorophyta phylum, respectively [22,23,24,25]. These different types of marine macroalgae are illustrated in Figure 1. There is an increasing demand for bioactive constituents in cosmetic and cosmeceutical applications from macroalgae. The applications of macroalgae-derived compounds to the cosmetic industry are based on their potential biological activities [26,27,28]. These are lipids, fatty acids, polysaccharides, vitamins, minerals, amino acids, phenolic compounds, proteins, pigments, etc., which have attracted attention for their skin cosmeceutical benefits [29,30,31].
Marine algae are one of the natural resources of phytochemical compounds, which confer potential biological activities [32,33]. Phenolic compounds are one of the bioactive compounds produced in seaweeds, are made of an aromatic ring with one or more hydroxyl groups, and their structures diversify from simple to complex, higher molecular weight compounds [34,35]. Many previous studies have been carried out in which phenolic compounds were isolated from marine algae and they include simple phenolic compounds or polyphenols such as flavonoids, phlorotannins, mycosporine-like amino acids (MAAs), bromophenols, and terpenoids [36]. The biological action of phenolic compounds is determined by the position of the hydroxyl groups, and the number of phenyl rings in the structure [37]. Brown algal species contain a high amount of phlorotannins whereas green and red algae mainly produce flavonoids, bromophenols, terpenoids, and mycosporine amino acids in response to environmental conditions [36,37,38,39,40]. Marine algae-derived phenolic compounds have a wide variety of applications such as enzyme inhibitory effects (for example, tyrosinase inhibition, elastase inhibition, collagenase inhibition, matrix metalloproteinase inhibition in photoprotection, inhibition of angiotensin-converting enzyme-1 (ACE-1), pro-inflammatory cyclooxygenase and lipoxygenase (COX-1, 2 and 5-LOX) as well as dipeptidyl peptidase-4 (DPP-4) inhibition, and hydroxymethyl glutaryl coenzyme A reductase (hMGCR) inhibition) antibacterial, antifungal, antioxidant, and anti-inflammatory properties, which can be very attractive when utilized in cosmetics and cosmeceutical product preparations [41,42,43,44,45,46,47]. In cosmetics, phlorotannin provides hyaluronidase activation, antiallergic, anti-wrinkle, anti-aging, skin whitening, photoprotection, and skin health improvement benefits [48,49]. This review represents marine algae-derived phenolic compounds, their chemical structures, together with their skin benefits and their potential in the skin cosmetics industries.
Phenolic compounds are one of the most researched marine macroalgae-derived biologically active compounds and are already utilized in various cosmeceutical preparations [50]. Normally, these phenolic compounds are not extracted because seaweed extracts contain a good number of phenolic compounds. According to Thomas and Kim [51], Nagayama et al. [52], and Hwang [53], phlorotannin is one of the marine algae-derived phenolic compounds with less toxicity than other natural antioxidant molecules and its anti-aging benefit is attracting the attention of researchers looking to use it as an ingredient in cosmetic formulations. Some marine algae extracts are rich in different phenolic compounds such as phlorotannin, phloroglucinol, eckol, dieckol, fucol, phlorethol, fuhalols, lignans, bromophenol, flavonoid, phenolic terpenoids, and mycosporine amino acids (MAAs). These phenolic compounds contribute to skin benefits, with antiaging, photoprotection, antiwrinkle, antiallergic, anti-inflammation, antioxidant, antimicrobial, antifungal, tyrosinase inhibition, anti-melanogenic, skin whitening, UVB protector, and antiacne properties, etc. [54,55,56]. Tang et al. [57], and Khanavi et al. [58] reported not only cytotoxicity but also the antibacterial effect of a phenolic fraction from the Ulva clathrata and Ulva flexuosa species that can be utilized for skin benefits. In addition, Lavoie et al. [59] identified the antibacterial activity of Cladophora socialis-derived phenolic compounds, such as 2,3,8,9-tetrahydroxybenzo[c]chromen-6-one, 3,4,30,40-tetrahydroxy-1,10-biphenyl, and cladophorol against methicillin-resistant Staphylococcus aureus. Based on a study carried out by Ko et al. [60], bromophenols, such as 5′-Hydroxyisoavrainvilleol found in green macroalga Avrainvillea nigricans, demonstrate promising antimicrobial activity. The green macroalga, Caulerpa sp., has been reported for various flavonoids, such as kaempferol and quercetin. These compounds have been studied and identified with antioxidant benefits [61]. In addition, Vidalia colensoi (as Osmundaria colensoi) derived from lanosol methyl ether, lanosol butenone, and rhodomelol has revealed antibacterial and antifungal activity against various bacterial and fungal pathogens. These effects proved to be bactericidal and bacteriostatic or fungicidal, fungistatic, and antiacne, and demonstrated a dose-dependent curve effect against the pathogenic organisms [62]. Mycosporine amino acids such as palythine, shinorine, asterina-330, Porphyra-334, palythinol, and usujirene have already been isolated and have antioxidant, photoprotective, and antiproliferative activity in the HeLa and HaCat cancer cell lines. In other studies, Lawrence et al. [63], Orfanoudaki et al. [64], and Becker et al. [65] reported the anti-inflammatory and immune-modulatory properties of mycosporine-like amino acids. These compounds can act as UV filters against photodamage. Moreover, Ecklonia cava-derived phlorotannin acts as an anti-UVB protective and reduces the photodamage effect provoked by UVB radiation [66]. Some other brown macroalgal species-derived phlorotannins such as dieckol, dioxinodehydroeckol, eckol, eckstolonol, phlorofucofuroeckol A, and 7-phloroeckol are being researched for skin-whitening and antiwrinkle properties, as well as tyrosinase and hyaluronidase inhibition [67,68,69,70,71,72,73]. Bak et al. [74] reported the hair growth-promoting activity of 7-phloroeckol isolated from E. cava.

2. Characterization and Types of Phenolic Compounds

The extraction and characterization of phenolic compounds from marine algae constitute interesting results, with those reported in the literature [75,76,77,78,79]. These compounds and their biological action are commonly correlated. However, some phenolic extracts have an interesting property but have not been fully characterized. Antioxidant activities have been reported in green seaweed-derived bromophenols and flavonoids. Farasat et al. [80], and Cho et al. [81] studied and proved the high radical scavenging activities of various green (Chlorophyta) species such as Ulva clathrata, U. compressa (formerly known as Enteromorpha compressa), U. intestinalis, U. linza, U. flexuosa, U. australis (formerly known as Ulva pertusa), Capsosiphon fulvescens, and Chaetomorpha moniligera. In their findings, antibacterial and cytotoxic effects on breast ductal carcinoma cell lines were verified in the phenolic fraction of U. clathrata and U. flexuosa [82,83]. In a more recent study, Lavoie et al. [84] reported C. socialis-derived phenolic compounds, such as 2,3,8,9-tetrahydrobenzo[c]chromen-6-one (Figure 2a), 3,4,3′,4′-tetrahydroxy-1,1′-biphenyl (Figure 2b), and cladophorol C (α-hydro-ω-[3,4-dihydroxyphenyl]octa[oxy(2-hydroxyphen-4-yl)]) (Figure 2c), have been identified with antibacterial activity against methicillin-resistant Staphylococcus aureus (MRSA).
The functions of phenol compounds in red marine algae have barely been studied, but they probably have multipurpose actions in cell life, such as antioxidant, chelation, and anti-infection actions, as well as cofactors or hormones [85]. However, some research is not with the isolated phenolic compounds but with an extract enriched in polyphenolics [85]. More than 8000 different structures of phenolic compounds are found in marine macroalgae, because of their importance in organisms’ growth, survival, and defense. The classification of phenolic compounds according to their chemical structures is depicted in Figure 3.
Moreover, the chemical structures of some phenolic compounds are illustrated in Figure 4. These compounds can be synthesized by various metabolic pathways such as the pentose phosphate pathway (PPP), and the phenylpropanoid and shikimate pathways.
Giada [86], and Vermerris and Nicholson [87] reported varieties of phenolic compound classification to comprise a large number of heterogeneous structures from simple to highly polymerized structures. A simple phenolic group is formed that reveals hydroxyl groups at different positions: ortho, meta, and para (1,2-, 2,3-, and 1,4-), respectively. Catechol, HQ, and phloroglucinol are examples of simple phenolic compounds found exclusively in macroalgae [88]. One of the previous studies showed catechol in twenty-seven Japanese green and red seaweeds [89]. It is quite common to find this simple phenol with a bromine substituent that is bromophenol. C6-CN phenolic compounds possess a basic C6-CN structure, where N is found between 1 and 3. Within this phenolic group, three subdivisions can be made in C6-C1, C6-C2, and C6-C3 and correspond to phenolic acids and aldehydes, which are characterized by having phenol with carboxylic group substitution [87]. Other examples are phenolic acid, hydroxybenzoic acids (C6-C1), acetophenones, phenylacetic acids, coumarins (C6-C2), coumarins, hydroxycinnamic acids, phenylpropanoids (C6-C3), naphthoquinones (C6-C4), xanthones (C6-C1-C6), stilbenoids, anthraquinones (C6-C2-C6), flavonoids, isoflavonoids (C6-C3-C6), lignans, neolignans ([C6-C3]2), lignins ([C6-C3]n), and condensed tannins ([C6-C3-C6]n). Gallic acid is one of the simplest phenolic acids used as a standard for total phenol content estimation. It can be found in high concentrations in the brown alga Halopteris scoparia. Other simple acids such as 4-hydroxybenzoic acid have been also reported in the brown alga Undaria pinnatifida [90]. C6-C2 are not very common in nature but the red alga Tichocarpus crinitus was studied for the C6-C3 category of phenolic compounds such as coumarins, isocoumarins, chromones, monolignols, hydroxycinnamic acids, and cinnamic aldehydes [87,91]. Moreover, Hartmann et al. [92] found the presence of coumarins in the green algae Dasycladus vermicularis. Other phenolic compounds include xanthonoids (C6-C1-C6), stilbenoids, anthraquinones, anthrones (C6-C2-C6), flavonoids (C6-C3-C6), and diarylheptanoids (C6-C7-C6). (C6-C3-C6) can be classified based on the arrangement of the C3 group that connects two benzene rings such as chalcones, aurones, and flavonoids. The last compound is further classified into different classes such as flavonols, flavones, isoflavones, anthocyanins, and flavanones. Cho et al. [81] found a high content of flavonoids in red algae, which was higher than in green and brown algae. In addition, Generalić Mekinić et al. [90] reported a good number of different flavonoids, catechin, epicatechin, gallate, and epigallocatechin, in brown algae species such as Eisenia bicyclis, Sargassum fusiforme, and Saccharina japonica.

2.1. Polyphenolic Compounds

Polyphenol is mainly of two types, phlorotannin, and phloroglucinol. The former is a polymer of phloroglucinol with an additional halogen or hydroxyl group whereas the latter contains an aromatic ring structure with three hydroxyl groups [93,94,95]. These can be subclassified into six different groups: (i) Eckols; (ii) Fucophlorethols; (iii) Fucols; (iv) Phlorethols; (v) Carmalols; and (vi) Fuhalols.

2.2. Lignans

Lignans are a type of phenolic compound, a dimer or oligomer, formed due to the union of monolignols, coniferyl alcohol, and sinapyl alcohol. Freile-Pelegrín and Robledo [96] reported the presence of lignans in calcified red marine algae Calliarthron cheilosporioides (Rhodophyta). Another polymeric phenol, lignin, is the most abundant organic polymer found in nature but not extensively studied in marine algae, which are structurally composed of monolignols (coniferyl alcohol, sinapyl alcohol), and lignan units randomly linked forming a polymeric network. Tannins are usually divided into three different chemical structures: hydrolyzable tannins, flavonoid-based tannins, and phlorotannins. The first one is derived from simple phenolic acids and their carbohydrate hydroxyl groups that are partially or completely esterified with phenolic groups. The second, flavonoid-based tannins, synthesize through flavins and catechins whereas the last, phlorotannins, are oligomers of phloroglucinol that are exclusively found in brown algae [97].

2.3. Phlorotannins

Phlorotannins, a group of compounds that majorly include dioxinodehydroeckol (eckostolonol), dieckol, eckol, phlorofucofuroeckol A, 7-phloroeckol, and fucofuroeckol A and 8,8′-bieckol, exhibit antioxidant-inhibitory effects against melanin synthesis, skin whitening (tyrosinase inhibition), and UV protection [98,99,100,101,102,103]. Kong et al. [104], Kim et al. [105], Ahn et al. [106], Lee et al. [107], and Li et al. [108] demonstrated the anti-proliferative, anti-inflammatory, and anti-adipogenic activities of Ecklonia cava (Phaeophyceae)-derived dioxinodehydroeckol, dieckol, and phlorofucofuroeckol.
Phlorotannins are the most deeply studied phenolic compounds from algae [109]. Their antioxidant power is 2 to 10 times higher when compared to ascorbic acid or tocopherol [110,111], which demonstrates their role as an anti-inflammatory agent [112]. They can act as an anti-UCB protector; Ryu et al. [113] suggested UVB protection by dioxinodehydroeckol from E. cava on the HaCat cells that reduce the provoked apoptosis. Moreover, phlorotannins such as dieckol, dioxinodehydroeckol, eckol, eckstolonol, phlorofucofuroeckol A, and 7-phloroeckol isolated from different marine algae are being researched in cosmetics as whiteners and antiwrinkle agents. They have been shown as promising tyrosinase inhibitors and hyaluronidase inhibitors [114,115,116,117,118,119,120]. In addition, Bak et al. [121] proved the hair growth-promoting activity of 7- E. cava-derived phloroeckol. Several reports have evaluated the effective antibacterial effect of phlorotannins, including from Ecklonia cava subsp. kurome (formerly Ecklonia kurome) (Phaeophyceae), against several food-borne pathogenic bacteria (e.g., methicillin-resistant Staphylococcus aureus (MRSA) strains, Campylobacter sp., and Streptococcus pyogenes) [122,123,124].

2.4. Bromophenols

Phenolic compounds such as bromophenol and benzoic acids have been fully isolated and characterized from red seaweeds [125]. Pérez et al. [126], Duan et al. [127], and Choi et al. [128] studied the antioxidant activity of Vertebrata constricta (formerly Polysiphonia stricta or P. urceolata) (Rhodophyta)-derived phenolic compounds, but that depends on the brominated units and degree of bromination. In the same study, Symphyocladia latiuscula-derived bromophenols reported antioxidant activity that was studied by DPPH assay. Colon et al. [129] found a cytotoxic effect on KB cells (human epithelial carcinoma cells) and the antimicrobial activity of Avrainvillea nigricans (Chlorophyta)-derived 5′-hydroxyisoavrainvilleol, which is an example of bromophenol. Moreover, Carte et al. [130] studied rawsonol, an example of bromophenol, isolated from the same genus, but from another species, A. rawsoni, which revealed an inhibitory effect in HMG-CoA reductase (a rate-controlling enzyme that produces cholesterol) activity. Besides, Estrada et al. [131] reported the antibacterial activity of brominated monoterpenoid quinol isolated from Cymopolia barbata (Chlorophyta) against S. aureus and Pseudomonas aeruginosa.

2.5. Flavonoids

Other classes of phenolic compounds have been investigated for varieties of applications in cosmetics. Tanna et al. [132] found the antioxidant activity of various flavonoids such as kaempferol and quercetin from Caleurpa spp. (Chlorophyta). Acanthophora spicifera (Rhodophyta)-derived flavonoid demonstrates a mixture of chlorogenic acid (69.64%), caffeic acid (12.86%), vitexin-rhamnose (12.35%), quercetin (1.41%), and catechol (0.59%), and this flavonoid-enriched extract has revealed antioxidant activity [133,134]. These molecules are multi-active components that play a role in UV radiation absorption, the neutralization of ROSs, and the inhibition of radical reactions, etc., which makes them important contributors to cosmeceuticals [135]. This antioxidant activity becomes helpful to overcome photo-induced skin aging. Ultraviolet light produces reactive oxygen species (ROS) in cells that initiate the intracellular and extracellular oxidative stresses that are responsible for wrinkle formation and atypical pigmentation on the skin [136].

2.6. Phenolic Terpenoids

Makkar and Chakraborty [137] studied a chromene-based phenolic compound from Gracilaria opuntia (Rhodophyta) that has been reported to have antioxidant activity in in vitro assays. Pillai et al. [138] reported the role of antioxidants in the prevention of extracellular matrix damage, the activation of MMPs, and inhibition of their expression. These molecules scavenge and quench radical oxygen species (ROS). Freile-Pelegrín and Robledo [96] found diterpenes and sesquiterpenes more commonly in red macroalgae as well as in Sargassaceae and Rhodomelaceae. J. Chappell and R. M. Coates [139] showed the role of sesquiterpene patchoulol as an extremely popular fragrance agent in colognes and perfumes. Ruberto and Baratta [140] found the significant lipid oxidation efficacy of oxygenated sesquiterpenes, which contributes to its role as an antioxidant that may act as a eustressor.

2.7. Mycosporine-like Amino Acid

Various marine algal species such as Asparagopsis armata, Chondrus crispus, Mastocarpus stellatus, Palmaria palmata, Gelidium sp., Pyropia sp. (formerly known as Porphyra sp.), Gracilaria cornea, Solieria chordalis, Grateloupia lanceola, and Curdiea racovitzae (Rhodophyta) have been investigated for this exclusive class of phenolic compounds. This class of compounds is more commonly found free in the intracellular space and around cell organelles sensitive to ultraviolet rays. Mycosporine-like amino acids (MAAs) are formed by cyclohexenone or cycloheximide chromophore conjugated to imino alcohol or an amino acid residue [141,142]. Various MAAs (palythine, shinorine, asterina-330, Porphyra-334, palythinol, and usujirene) have already been studied that have high antioxidant, photoprotection, and anti-proliferative (HeLa cancer cell line, human cervical adenocarcinoma cell line) and HaCat (human immortalized keratinocyte) activity [143,144]. Recent studies have reported other important activities such as anti-inflammatory, and photoprotective activities (an alternative to the synthetic UV-R filters in sunscreens). Thus, MAAs seems to be a special focus on a specific area and application that can be applied to humans. Based on the literature, the different types of marine algae-derived phenolic compounds and their skin cosmetic benefits are tabulated in the below Table 1.

3. Extraction of Phenolic Compounds

There are several extraction techniques available for obtaining phenolic compounds; two general techniques are found: conventional and nonconventional extraction techniques. The conventional techniques include simple solid solvent extraction, whereas nontraditional techniques include microwave-assisted extraction, subcritical CO2 extraction, ultrasound-assisted extraction, and pressurized liquid extraction, among others. The extraction and characterization of phenolic compounds from marine algae reported an interesting result as in Figure 5 [44,45,46,47,48].
The most important step is to select an appropriate extraction method, since many procedures of extraction are available nowadays. Traditional methods include heat-assisted extraction or maceration, percolation, and Soxhlet extraction as reported by Aires [233]. One of the classical methods is maceration, in which the components are extracted by submerging marine algae in an appropriate solvent or solvent combinations [234]. On a large scale, at the industrial level, ethanol is preferred as a solvent for extraction because of its economic benefit [235]. This procedure is widely applicable in current practice. In this method, methanol, ethanol, acetone, water, and ethyl ethanoate in different proportions are commonly utilized for extraction. The selection can be done based on polarity. Due to the hydrophilic nature of these compounds, hydroalcoholic solvent is the most effective for this process. Some previous studies have mentioned the combination of solvents, with acids such as citric acid, tartaric acid, or HCl potentially improving the extraction of phenolic compounds [236,237]. In traditional procedures, Soxhlet extraction provides better results of extraction in terms of yield, although this technique also presents some demerits such as the degradation of temperature-sensitive compounds as some phenolic acids, tannins, and anthocyanins require a large number of solvents and are time-consuming. Besides, this classical Soxhlet extraction method is a continuous process; the solvent can easily be recycled, and less time and less solvent are used than in maceration and percolation [238]. Moreover, the used extract of the selected algae is constantly being heated at the boiling point of the solvent and it may damage some temperature-sensitive components, which may affect further analysis [239]. Santos-Buelga et al. [236] reported the use of the Soxhlet method in the extraction of lipophilic compounds. Moreover, heat-assisted extraction can be divided into two steps. The first step is the faster step and the second one is slower. The faster method consists of a compound’s transference from the matrix surface to the solvent whereas the slower method consists of diffusion from the matrix to the solvent. Extraction is mainly affected by the type of sample, type of solvent, temperature value, and time. The main disadvantage is that it requires filtration, decantation, or clarification to separate the solid parts when the extraction is done. It also requires a large number of solvents and takes a long time. Hence, these classical methods are not efficient and environmentally friendly due to the high requirements of the organic solvents [240]. With advancements, new techniques have evolved to improve the efficacy and accuracy of extraction.
A non-conventional technique, Pressurized Liquid Extraction (PLE), also known as extraction with pressurized solvent, includes high pressure (1 to 15 MPa), short processing time, and temperature ranges of about 50 to 200 °C using a low volume of nontoxic solvent and thus being considered a green technology. Otero et al. [241] observed a highest extraction yield of 37% for 80 °C and 52% for 160 °C using diluted ethanol from the brown alga Laminaria ochroleuca (Phaeophyceae) at 100 bars. Microwave-assisted extraction is mainly used for the extraction of polyphenols and polysaccharides. This method can be performed in open (at atmospheric pressure) or closed (higher than atmospheric pressure) vessels. In this method, electromagnetic waves cause changes in cell structures. Two mechanisms, ionic conduction, and dipole rotation, transform electromagnetic energy into calorific energy [242,243]. There are many affecting factors such as the type of extraction, frequency of microwave, solid-to-solvent ratio, temperature, pressure, and time. Besides, the demerit is that high microwave power and elevated temperature may degrade phenolic compounds [244]. Cikoš et al. [243] showed the merits of ultrasound-assisted extraction (UAE) for the extraction of phenolic compounds from algae including low temperature, short times, and low amounts of solvent. However, the ultrasonication time can increase the temperature, which may affect the stability of the phenolic compounds. Besides, there are some other affecting factors such as frequency, power, time, temperature, and solid: solvent ratio [245]. This method applies ultrasound waves with a frequency between 20 kHz and 100 kHz, which creates bubbles due to pressure differences. Then, the bubbles collapse and cavitation occurs, causing the near liquid–solid interface breakdown of particles with the release of bioactive compounds to the matrix. The subcritical water extraction (SWE) method requires an application of water at a higher temperature (100–374 °C) than its boiling point under high pressure (10–60 bar) to maintain its liquid state for 5–10 min. In this technique, pressure, time, temperature, and the selection of the solvent are affecting factors for extraction [246,247]. In the SC-CO2 method, CO2 is a nontoxic gas used as a supercritical fluid, so the fluid behaves like a liquid simultaneously, which makes extraction easier. Ethanol can be used to change the polarity of CO2, while extraction, low temperature, and low pressure are used to degrade the phenolic compounds. Moreover, varieties of seaweed species were studied and explored for a great variety of biochemicals and their skin benefits.

4. Commercial Availability of Seaweed-Based Cosmetic Products

Marine algae have received more attention recently in cosmetics. Several skin cosmetic products are available in the market, some of them prepared by using algal extract, whereas some contain extracted bioactive compounds from potential marine algae. Nowadays, several cosmetic companies are using marine algae extracts and compounds in cosmetic preparations, as an active ingredient or as an excipient, gelling, thickening, preservative, additive, aroma, or fragrance agent [248]. For example, red alga Gracilaria sp. extracts are integrated into different products, such as A-Gel, Sealaria (Kfar Hess, Israel; https://www.sealaria.com/our-products/, accessed on 23 September 2022), facial masks by Balinique (Miami, FL, USA; https://www.gsg-creative.com/cases/balinique, accessed on 23 September 2022), and hydrating creams by Thalasso (Rosa Graf, Stamford, USA; https://skincare.rosagraf.com/product-category/thalasso/, accessed on 23 September 2022) [249]. Helioguard® 365 (Mibelle Biochemistry; https://mibellebiochemistry.com/helioguardtm-365, accessed on 21 September 2022) is a cosmetic ingredient complex that was formulated by using mycosporine-like amino acids derived from the red alga Porphyra umbilicalis (Rhodophyta), which has a powerful UV-protective capacity. This product proved suitable to use daily and boasts photoaging benefits. Besides, a product prepared by OSEA Malibu (Los Angeles, USA; https://oseamalibu.com/products/undaria-algae-oil, accessed on 24 September 2022), Undaria algae body oil, contains Undaria pinnatifida powder, which confers antioxidant benefits and improves skin nourishment and firmness. “Hyaluronic Sea Serum” is prepared by mixing Codium fragile (Chlorophyta) extract with other natural extracts and ingredients to improve hydration, minimize lines, and enhance firmness (shorturl.at/lq148). There are some other products such as W2 SPF 50 PA+++ Red Seaweed (Life Essentials Personal Care Pvt Ltd., Haryana, India; shorturl.at/lrS28), which contains red alga extract that prevents dehydration of the skin and provides photoprotection benefits; Seaweed Cleansing Soap (Mario Badescu Skin Care Inc., New York, NY 10022 https://www.mariobadescu.com/product/seaweed-cleansing-soap, accessed on 24 September 2022), which contains seaweed grains and Bladderwrack (Seaweed) extract that contribute to nourishment, creamy cleansing, and soothing and gentle mineral exfoliation benefits; Seaweed Oil Control Gel Cream (The Body Shop International Limited, West Sussex, UK; https://www.thebodyshop.com/en-gb/face/moisturisers/seaweed-oil-control-gel-cream/p/p000181, accessed on 23 September 2022), which controls shine, hydration, skin protection, and nourishment; Sea Algae Daily Repair Serum (Prolixr, India; https://prolixr.in/products/sea-algae-daily-repair-serum, accessed on 23 September 2022), which replenishes moisturization and improves skin elasticity; and in addition other products, Seaweed Oil-Control Gel Cream (The Body Shop International Limited, UK), prepared by using the extract of the brown alga Fucus vesiculosus with other natural ingredients that help to maintain oil balance and excess sebum for a matte and shine-free complexion. Likewise, Sea Algae Daily Repair Ace Serum (FURR, Pee Safe, India; https://furr.in/products/furr-daily-repair-face-serum, accessed on 18 October 2022) strengthens the elastin tissues, revitalizes the skin and reduces shine. Another market-available product is Universal Face Oil by MARA Beauty (Queenstown, New Zealand; https://themarabeauty.com/products/algae-moringa-universal-face-oil, accessed on 18 October 2022) that is made by mixing algae plus moringa. In this product, algae play a proprietary role that enhances the natural hyaluronic acid synthesis and is loaded with phytonutrients and fatty acids to improve the plumpness, firmness, and smoothness of the skin. Moreover, Green Confertii Extract-NS (Gyeonggi-do, Republic of Korea; https://cosmetics.specialchem.com/product/i-the-garden-of-naturalsolution-green-confertii-extract-ns, accessed on 18 October 2022) contains an extract of Ulva compressa (formerly Enteromorpha compressa) (Chlorophyta), which is rich in bioactive compounds, polysaccharides, flavonoids, tannins, and acrylic acid. This extract possesses antioxidant, antiallergic, and antimicrobial activity.

5. Conclusions

The macroalgae-derived phenolic compounds are scarce and further exploration will create a good library of bioactive chemicals and enhance the possibility of discovering new compounds in different types of skin cosmetic preparations. Hence, phycological research, mainly isolation, extraction, and the characterization of seaweed species, will improve the cosmetic market commercially. The main focus is on the concentration of bioactive compounds present in macroalgal species, which creates a real problem at the formulation level. Polyphenolic compounds and other classes of chemical compounds are the attention of ongoing research. Their study is very limited and there is a lack of clarity about the in vivo effects of seaweed-derived phenolic compounds and their interaction with human cells. These problems can be overcome by using various methodologies and determination methods that evaluate at a deeper level to make them safer. More studies and research are needed on the characterization of phytochemicals, extraction, characterization, and in vitro and in vivo study for toxicity by diversified methodology. Overall, new research studies are required to analyze and fully understand their biological benefits in cosmetic formulation and on the skin to make the cosmetic sector sustainable.
Ultimately, many more seaweed species will require study and characterization for their application in cosmetic formulation and to understand their skin benefits. Due to their tremendous number of applications and various biological benefits, marine macroalgae are gaining attention and becoming increasingly attractive in the exploration of the skin cosmetic properties of their natural bioactive extracts and formulations.

Author Contributions

Conception and design of the idea: H.S.K. and L.P.; Design of tables and figures: H.S.K. and L.P.; Writing and bibliographical research: H.S.K. and L.P.; Supervision and manuscript revision: H.S.K. and L.P. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financed by national funds through the FCT—Foundation for Science and Technology, I.P., within the scope of the project LA/P/0069/2020 granted to the Associate Laboratory ARNET, UIDB/04292/2020 granted to MARE—Marine and Environmental Sciences Centre.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Martin, K.I.; Glaser, D.A. Cosmeceuticals: The new medicine of beauty. Mo. Med. 2011, 108, 60. [Google Scholar] [PubMed]
  2. Dureja, H.; Kaushik, D.; Gupta, M.; Kumar, V.; Lather, V. Cosmeceuticals: An emerging concept. Indian J. Pharmacol. 2005, 37, 155. [Google Scholar] [CrossRef] [Green Version]
  3. Draelos, Z.D. The cosmeceutical realm. Clin. Dermatol. 2008, 26, 627–632. [Google Scholar] [CrossRef] [PubMed]
  4. Yin, S.N.; Hayes, R.B.; Linet, M.S.; Li, G.L.; Dosemeci, M.; Travis, L.B.; Zhang, Z.N.; Li, D.G.; Chow, W.H.; Wacholder, S.; et al. An expanded cohort study of cancer among benzene-exposed workers in China. Benzene Study Group. Environ. Health Perspect. 1996, 104 (Suppl. S6), 1339–1341. [Google Scholar] [CrossRef] [PubMed]
  5. Briganti, S.; Camera, E.; Picardo, M. Chemical and instrumental approaches to treat hyperpigmentation. Pigment. Cell Res. 2003, 16, 101–110. [Google Scholar] [CrossRef]
  6. Zhang, L.; Robertson, M.L.; Kolachana, P.; Davison, A.J.; Smith, M.T. Benzene metabolite, 1,2,4-benzenetriol, induces micronuclei and oxidative DNA damage in human lymphocytes and HL60 cells. Environ. Mol. Mutagen. 1993, 21, 339–348. [Google Scholar] [CrossRef]
  7. Fernández-Álvarez, M.; Llompart, M.; Sánchez-Prado, L.; García-Jares, C.; Lores, M. Photochemical behavior of UV filter combinations. In Cosmetics: Types, Allergies and Applications; Nova Science Publishers, Inc.: Hauppauge, NY, USA, 2010; p. 1. [Google Scholar]
  8. Knowland, J.; McKenzie, E.A.; McHugh, P.J.; Cridland, N.A. Sunlight-induced mutagenicity of a common sunscreen ingredient. FEBS Lett. 1993, 324, 309–313. [Google Scholar] [CrossRef] [Green Version]
  9. Kerdudo, A.; Burger, P.; Merck, F.; Dingas, A.; Rolland, Y.; Michel, T.; Fernandez, X. Development of a natural ingredient–Natural preservative: A case study. Comptes Rendus. Chim. 2016, 19, 1077–1089. [Google Scholar] [CrossRef] [Green Version]
  10. Mowad, C.M. Allergic contact dermatitis caused by parabens: 2 case reports and a review. Am. J. Contact Dermat. 2000, 11, 53–56. [Google Scholar] [CrossRef]
  11. Hafeez, F.; Maibach, H. An overview of parabens and allergic contact dermatitis. Skin Ther. Lett. 2013, 18, 5–7. [Google Scholar]
  12. Barrett, J. Chemical Exposures: The Ugly Side of Beauty Products. Environ. Health Perspect. 2005, 113, A24. [Google Scholar] [CrossRef] [PubMed]
  13. Khan, A.D.; Alam, M.N. Cosmetics and their associated adverse effects: A review. J. Appl. Pharm. Sci. Res. 2019, 2, 1–6. [Google Scholar] [CrossRef] [Green Version]
  14. Warbanski, M. The ugly side of the beauty industry. Herizons 2007, 21, 24–28. [Google Scholar]
  15. Kaličanin, B.; Velimirović, D. A study of the possible harmful effects of cosmetic beauty products on human health. Biol. Trace Elem. Res. 2016, 170, 476–484. [Google Scholar] [CrossRef] [PubMed]
  16. Pangestuti, R.; Shin, K.H.; Kim, S.K. Anti-photoaging and potential skin health benefits of seaweeds. Mar. Drugs 2021, 19, 172. [Google Scholar] [CrossRef] [PubMed]
  17. Kim, A.R.; Shin, T.S.; Lee, M.S.; Park, J.Y.; Park, K.E.; Yoon, N.Y.; Kim, J.S.; Choi, J.S.; Jang, B.C.; Byun, D.S.; et al. Isolation and identification of phlorotannins from Ecklonia stolonifera with antioxidant and anti-inflammatory properties. J. Agric. Food Chem. 2009, 57, 3483–3489. [Google Scholar] [CrossRef] [PubMed]
  18. Pereira, L. Therapeutic and Nutritional Uses of Algae; CRC Press: Boca Raton, FL, USA, 2018; ISBN 9781498755382. [Google Scholar]
  19. Maqsood, S.; Benjakul, S.; Shahidi, F. Emerging role of phenolic compounds as natural food additives in fish and fish products. Crit. Rev. Food Sci. Nutr. 2013, 53, 162–179. [Google Scholar] [CrossRef]
  20. Panzella, L.; Napolitano, A. Natural phenol polymers: Recent advances in food and health applications. Antioxidants 2017, 6, 30. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Leandro, A.; Pereira, L.; Gonçalves, A.M. Diverse applications of marine macroalgae. Mar. Drugs 2019, 18, 17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Pereira, L. Algae. Litoral of Viana do Castelo; Câmara Municipal de Viana do Castelo: Viana do Castelo, Portugal, 2010; pp. 7–8. ISBN 978-972-588-217-7. [Google Scholar]
  23. Pereira, L. Guia Ilustrado das Macroalgas—Conhecer e Reconhecer Algumas Espécies da Flora Portuguesa; Universityde Coimbra Press: Coimbra, Portugal, 2009; p. 91. ISBN 978-989-26-0002-4. [Google Scholar]
  24. Pereira, L. Chapter 4—Cytological and cytochemical aspects in selected carrageenophytes (Gigartinales, Rhodophyta). In Advances in Algal Cell Biology; Heimann, K., Katsaros, C., Eds.; De Gruyter: Berlin, Germany, 2012; pp. 81–104. ISBN 978-3-11-022960-8. [Google Scholar]
  25. González-Minero, F.J.; Bravo-Díaz, L. The use of plants in skin-care products, cosmetics and fragrances: Past and present. Cosmetics 2018, 5, 50. [Google Scholar] [CrossRef] [Green Version]
  26. Ibañez, E.; Herrero, M.; Mendiola, J.A.; Castro-Puyana, M. Extraction and characterization of bioactive compounds with health benefits from marine resources: Macro and micro algae, cyanobacteria, and invertebrates. In Marine Bioactive Compounds; Springer: Boston, MA, USA, 2012; pp. 55–98. [Google Scholar]
  27. Vo, T.S.; Kim, S.K. Fucoidans as a natural bioactive ingredient for functional foods. J. Funct. Foods 2013, 5, 16–27. [Google Scholar] [CrossRef]
  28. Venkatesan, J.; Kim, S.K. Osteoporosis treatment: Marine algal compounds. Adv. Food Nutr. Res. 2011, 64, 417–427. [Google Scholar]
  29. Berthon, J.Y.; Nachat-Kappes, R.; Bey, M.; Cadoret, J.P.; Renimel, I.; Filaire, E. Marine algae as attractive source to skin care. Free. Radic. Res. 2017, 51, 555–567. [Google Scholar] [CrossRef]
  30. Gam, D.H.; Park, J.H.; Hong, J.W.; Jeon, S.J.; Kim, J.H.; Kim, J.W. Effects of Sargassum thunbergii extract on skin whitening and anti-wrinkling through inhibition of TRP-1 and MMPs. Molecules 2021, 26, 7381. [Google Scholar] [CrossRef] [PubMed]
  31. Querellou, J.; Børresen, T.; Boyen, C.; Dobson, A.; Höfle, M.; Ianora, A.; Jaspars, M.; Kijjoa, A.; Olafsen, J.; Rigos, G. Marine biotechnology: Realising the full potential of Europe. VLIZ Spec. Publ. 2010, 47, 21. [Google Scholar]
  32. Acosta-Estrada, B.A.; Gutiérrez-Uribe, J.A.; Serna-Saldívar, S.O. Bound phenolics in foods, a review. Food Chem. 2014, 152, 46–55. [Google Scholar] [CrossRef]
  33. Komes, D.; Belščak-Cvitanović, A.; Horžić, D.; Rusak, G.; Likić, S.; Berendika, M. Phenolic composition and antioxidant properties of some traditionally used medicinal plants affected by the extraction time and hydrolysis. Phytochem. Anal. 2011, 22, 172–180. [Google Scholar] [CrossRef]
  34. Jesumani, V.; Du, H.; Aslam, M.; Pei, P.; Huang, N. Potential use of seaweed bioactive compounds in skincare—A review. Mar. Drugs 2019, 17, 688. [Google Scholar] [CrossRef] [Green Version]
  35. Chisté, R.C.; Godoy, H.T.; Prado, M.A. The phenolic compounds and the antioxidant potential of infusion of herbs from the Brazilian Amazonian region. Food Res. Int. 2013, 53, 875–881. [Google Scholar]
  36. Nurilmala, M.; Hidayat, T.; Sudirdjo, F. Characteristics of seaweed as raw materials for cosmetics. Aquat. Procedia 2016, 7, 177–180. [Google Scholar]
  37. Pereira, L. Seaweeds as source of bioactive substances and skin care therapy—Cosmeceuticals, algotheraphy, and thalassotherapy. Cosmetics 2018, 5, 68. [Google Scholar] [CrossRef] [Green Version]
  38. Gómez-Guzmán, M.; Rodríguez-Nogales, A.; Algieri, F.; Gálvez, J. Potential role of seaweed polyphenols in cardiovascular-associated disorders. Mar. Drugs 2018, 16, 250. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Morais, T.; Inácio, A.; Coutinho, T.; Ministro, M.; Cotas, J.; Pereira, L.; Bahcevandziev, K. Seaweed potential in the animal feed: A review. J. Mar. Sci. Eng. 2020, 8, 559. [Google Scholar] [CrossRef]
  40. Costa, P.; Gonçalves, S.; Valentão, P.; Andrade, P.B.; Almeida, C.; Nogueira, J.M.; Romano, A. Metabolic profile and biological activities of Lavandula pedunculata subsp. lusitanica (Chaytor) Franco: Studies on the Essential Oil and Polar Extracts. Food Chem. 2013, 141, 2501–2506. [Google Scholar] [PubMed]
  41. Wijesinghe, W.A.J.P.; Jeon, Y.J. Biological activities and potential cosmeceutical applications of bioactive components from brown seaweeds: A review. Phytochem. Rev. 2011, 10, 431–443. [Google Scholar] [CrossRef]
  42. Kizhakkekalam, V.K.; Chakraborty, K. Pharmacological properties of marine macroalgae-associated heterotrophic bacteria. Arch. Microbiol. 2019, 201, 505–518. [Google Scholar] [CrossRef]
  43. Holmquist, B.; Bunning, P.; Riordan, J.F. A continuous spectrophotometric assay for angiotensin converting enzyme. Anal. Biochem. 1979, 95, 540–548. [Google Scholar] [CrossRef] [PubMed]
  44. Brand-Williams, W.; Cuvelier, M.; Berset, C. Use of free radical method to evaluate antioxidant activity. LWT-Food Sci. Technol. 1995, 28, 25–30. [Google Scholar] [CrossRef]
  45. Re, R.; Pellegrini, N.; Proteggente, A.; Pannala, A.; Yang, M.; Rice Evans, C. Antioxidant activity applying an improved ABTS radical cation decolorization assay. Free Radic. Biol. Med. 1999, 26, 1231–1237. [Google Scholar] [CrossRef]
  46. Ademiluyi, A.O.; Oboh, G. Soybean phenolic-rich extracts inhibit key-enzymes linked to type-2 diabetes (α-amylase and α-glucosidase) and hypertension (angiotensin-I converting enzyme) in-vitro. Exp. Toxicol. Pathol. 2013, 65, 305–309. [Google Scholar] [CrossRef]
  47. Charlier, C.; Michaux, C. Dual inhibition of cyclooxygenase-2 (COX-2) and 5-lipoxygenase (5-LOX) as a new strategy to provide safer non-steroidal anti-inflammatory drugs. Eur. J. Med. Chem. 2003, 38, 645–659. [Google Scholar] [CrossRef] [PubMed]
  48. Kalasariya, H.S.; Yadav, V.K.; Yadav, K.K.; Tirth, V.; Algahtani, A.; Islam, S.; Gupta, N.; Jeon, B.-H. Seaweed-Based Molecules and Their Potential Biological Activities: An Eco-Sustainable Cosmetics. Molecules 2021, 26, 5313. [Google Scholar] [CrossRef] [PubMed]
  49. Nasab, S.B.; Homaei, A.; Pletschke, B.I.; Salinas-Salazar, C.; Castillo-Zacarias, C.; Parra-Saldívar, R. Marine resources effective in controlling and treating diabetes and its associated complications. Process Biochem. 2020, 92, 313–342. [Google Scholar] [CrossRef]
  50. Dias, R.; Oliveira, H.; Fernandes, I.; Simal-Gandara, J.; Perez-Gregorio, R. Recent advances in extracting phenolic compounds from food and their use in disease prevention and as cosmetics. Crit. Rev. Food Sci. Nutr. 2021, 61, 1130–1151. [Google Scholar] [CrossRef] [PubMed]
  51. Audibert, L.; Fauchon, M.; Blanc, N.; Hauchard, D.; Ar Gall, E. Phenolic compounds in the brown seaweed Ascophyllum nodosum: Distribution and radical-scavenging activities. Phytochem. Anal. 2010, 21, 399–405. [Google Scholar] [CrossRef]
  52. 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]
  53. Hwang, H.J. Skin elasticity and sea polyphenols. Seanol Sci. Centre Rev. 2010, 1, 17. [Google Scholar]
  54. Malakar, B.; Mohanty, K. The budding potential of algae in cosmetics. In Algae; Springer: Singapore, 2021; pp. 181–199. [Google Scholar]
  55. Sakthivel, R.; Devi, K.P. Antioxidant, anti-inflammatory and anticancer potential of natural bioactive compounds from seaweeds. Stud. Nat. Prod. Chem. 2019, 63, 113–160. [Google Scholar]
  56. Panzella, L. Natural phenolic compounds for health, food and cosmetic applications. Antioxidants 2020, 9, 427. [Google Scholar] [CrossRef]
  57. Tang, H.; Inoue, M.; Uzawa, Y.; Kawamura, Y. Anti-tumorigenic components of a sea weed, Enteromorpha clathrata. BioFactors 2004, 22, 107–110. [Google Scholar] [CrossRef]
  58. Khanavi, M.; Gheidarloo, R.; Sadati, N.; Ardekani, M.R.S.; Nabavi, S.M.B.; Tavajohi, S.; Ostad, S.N. Cytotoxicity of fucosterol containing fraction of marine algae against breast and colon carcinoma cell line. Pharmacogn. Mag. 2012, 8, 60. [Google Scholar] [PubMed] [Green Version]
  59. Sirbu, R.; Stanciu, G.; Tomescu, A.; Ionescu, A.M.; Cadar, E. Evaluation of antioxidant and antimicrobial activity in relation to total phenolic content of green algae from Black Sea. Rev. Chim. 2019, 70, 1197–1203. [Google Scholar] [CrossRef]
  60. Ko, S.C.; Ding, Y.; Kim, J.; Ye, B.R.; Kim, E.A.; Jung, W.K.; Heo, S.J.; Lee, S.H. Bromophenol (5-bromo-3, 4-dihydroxybenzaldehyde) isolated from red alga Polysiphonia morrowii inhibits adipogenesis by regulating expression of adipogenic transcription factors and AMP-activated protein kinase activation in 3T3-L1 adipocytes. Phytother. Res. 2019, 33, 737–744. [Google Scholar] [CrossRef] [PubMed]
  61. Tanna, B.; Yadav, S.; Mishra, A. Anti-proliferative and ROS-inhibitory activities reveal the anticancer potential of Caulerpa species. Mol. Biol. Rep. 2020, 47, 7403–7411. [Google Scholar] [CrossRef] [PubMed]
  62. Barreto, M.; Meyer, J.J.M. Isolation and antimicrobial activity of a lanosol derivative from Osmundaria serrata (Rhodophyta) and a visual exploration of its biofilm covering. S. Afr. J. Bot. 2006, 72, 521–528. [Google Scholar] [CrossRef] [Green Version]
  63. Lawrence, K.P.; Long, P.F.; Young, A.R. Mycosporine-like amino acids for skin photoprotection. Curr. Med. Chem. 2018, 25, 5512–5527. [Google Scholar] [CrossRef]
  64. Orfanoudaki, M.; Hartmann, A.; Alilou, M.; Gelbrich, T.; Planchenault, P.; Derbré, S.; Schinkovitz, A.; Richomme, P.; Hensel, A.; Ganzera, M. Absolute configuration of mycosporine-like amino acids, their wound healing properties and in vitro anti-aging effects. Mar. Drugs 2019, 18, 35. [Google Scholar] [CrossRef] [Green Version]
  65. Becker, K.; Hartmann, A.; Ganzera, M.; Fuchs, D.; Gostner, J.M. Immunomodulatory effects of the mycosporine-like amino acids shinorine and porphyra-334. Mar. Drugs 2016, 14, 119. [Google Scholar] [CrossRef]
  66. Shim, S.Y.; Choi, J.S.; Byun, D.S. Inhibitory effects of phloroglucinol derivatives isolated from Ecklonia stolonifera on FcεRI expression. Bioorg. Med. Chem. 2009, 17, 4734–4739. [Google Scholar] [CrossRef]
  67. Eom, S.H.; Lee, E.H.; Park, K.; Kwon, J.Y.; Kim, P.H.; Jung, W.K.; Kim, Y.M. Eckol from Eisenia bicyclis inhibits inflammation through the Akt/NF-κB signaling in Propionibacterium acnes-induced human keratinocyte Hacat cells. J. Food Biochem. 2017, 41, e12312. [Google Scholar] [CrossRef]
  68. Wijesekara, I.; Yoon, N.Y.; Kim, S.K. Phlorotannins from Ecklonia cava (Phaeophyceae): Biological activities and potential health benefits. Biofactors 2010, 36, 408–414. [Google Scholar] [CrossRef] [PubMed]
  69. Shrestha, S.; Zhang, W.; Smid, S.D. Phlorotannins: A review on biosynthesis, chemistry and bioactivity. Food Biosci. 2021, 39, 100832. [Google Scholar] [CrossRef]
  70. Park, D.C.; Ji, C.I.; Kim, S.H.; Jung, K.J.; Lee, T.G.; Kim, I.S.; Park, Y.H.; Kim, S.B. Characteristics of tyrosinase inhibitory extract from Ecklonia stolonifera. Fish. Aquat. Sci. 2000, 3, 195–199. [Google Scholar]
  71. Kilic, M.; Orhan, I.E.; Eren, G.; Okudan, E.S.; Estep, A.S.; Bencel, J.J.; Tabanca, N. Insecticidal activity of forty-seven marine algae species from the Mediterranean, Aegean, and Sea of Marmara in connection with their cholinesterase and tyrosinase inhibitory activity. S. Afr. J. Bot. 2021, 143, 435–442. [Google Scholar] [CrossRef]
  72. Kim, J.H.; Lee, S.; Park, S.; Park, J.S.; Kim, Y.H.; Yang, S.Y. Slow-binding inhibition of tyrosinase by Ecklonia cava phlorotannins. Mar. Drugs 2019, 17, 359. [Google Scholar] [CrossRef] [Green Version]
  73. Cha, S.H.; Ko, C.I.; Kim, D.; Jeon, Y.J. Protective effects of phlorotannins against ultraviolet B radiation in zebrafish (Danio rerio). Vet. Dermatol. 2012, 23, 51-e12. [Google Scholar] [CrossRef] [PubMed]
  74. Manandhar, B.; Wagle, A.; Seong, S.H.; Paudel, P.; Kim, H.R.; Jung, H.A.; Choi, J.S. Phlorotannins with potential anti-tyrosinase and antioxidant activity isolated from the marine seaweed Ecklonia stolonifera. Antioxidants 2019, 8, 240. [Google Scholar] [CrossRef] [Green Version]
  75. Santos, S.; Félix, R.; Pais, A.; Rocha, S.M.; Silvestre, A. The Quest for Phenolic Compounds from Macroalgae: A Review of Extraction and Identification Methodologies. Biomolecules 2019, 9, 847. [Google Scholar] [CrossRef] [Green Version]
  76. Jimenez-Lopez, C.; Pereira, A.G.; Lourenço-Lopes, C.; Garcia-Oliveira, P.; Cassani, L.; Fraga-Corral, M.; Prieto, M.A.; Simal-Gandara, J. Main bioactive phenolic compounds in marine algae and their mechanisms of action supporting potential health benefits. Food Chem. 2021, 341, 128262. [Google Scholar] [CrossRef]
  77. Parys, S.; Rosenbaum, A.; Kehraus, S.; Reher, G.; Glombitza, K.W.; König, G.M. Evaluation of quantitative methods for the determination of polyphenols in algal extracts. J. Nat. Prod. 2007, 70, 1865–1870. [Google Scholar] [CrossRef]
  78. Ktari, L.; Mdallel, C.; Aoun, B.; Chebil Ajjabi, L.; Sadok, S. Fucoxanthin and Phenolic Contents of Six Dictyotales From the Tunisian Coasts With an Emphasis for a Green Extraction Using a Supercritical CO2 Method. Front. Mar. Sci. 2021, 8, 647159. [Google Scholar] [CrossRef]
  79. Mateos, R.; Pérez-Correa, J.R.; Domínguez, H. Bioactive properties of marine phenolics. Mar. Drugs 2020, 18, 501. [Google Scholar] [CrossRef]
  80. Farasat, M.; Khavari-Nejad, R.A.; Nabavi, S.M.B.; Namjooyan, F. Antioxidant activity, total phenolics and flavonoid contents of some edible green seaweeds from northern coasts of the Persian Gulf. Iran. J. Pharm. Res. IJPR 2014, 13, 163. [Google Scholar] [PubMed]
  81. Cho, M.; Kang, I.J.; Won, M.H.; Lee, H.S.; You, S. The antioxidant properties of ethanol extracts and their solvent-partitioned fractions from various green seaweeds. J. Med. Food 2010, 13, 1232–1239. [Google Scholar] [CrossRef] [PubMed]
  82. Huang, P.; Hong, J.; Mi, J.; Sun, B.; Zhang, J.; Li, C.; Yang, W. Polyphenols extracted from Enteromorpha clathrata alleviates inflammation in lipopolysaccharide-induced RAW 264.7 cells by inhibiting the MAPKs/NF-κB signaling pathways. J. Ethnopharmacol. 2022, 286, 114897. [Google Scholar] [CrossRef]
  83. Abdul, Q.A.; Choi, R.J.; Jung, H.A.; Choi, J.S. Health benefit of fucosterol from marine algae: A review. J. Sci. Food Agric. 2016, 96, 1856–1866. [Google Scholar] [CrossRef]
  84. Lavoie, S.; Sweeney-Jones, A.M.; Mojib, N.; Dale, B.; Gagaring, K.; McNamara, C.W.; Quave, C.L.; Soapi, K.; Kubanek, J. Antibacterial oligomeric polyphenols from the green alga Cladophora socialis. J. Org. Chem. 2019, 84, 5035–5045. [Google Scholar] [CrossRef] [Green Version]
  85. Seedevi, P.; Moovendhan, M.; Viramani, S.; Shanmugam, A. Bioactive potential and structural chracterization of sulfated polysaccharide from seaweed (Gracilaria corticata). Carbohydr. Polym. 2017, 155, 516–524. [Google Scholar] [CrossRef] [PubMed]
  86. Santos-Sánchez, N.F.; Salas-Coronado, R.; Villanueva-Cañongo, C.; Hernández-Carlos, B. Antioxidant compounds and their antioxidant mechanism. Antioxidants 2019, 10, 1–29. [Google Scholar]
  87. Vermerris, W.; Nicholson, R. Families of phenolic compounds and means of classification. In Phenolic Compound Biochemistry; Springer: Dordrecht, The Netherlands, 2008; pp. 1–34. [Google Scholar]
  88. Tsimogiannis, D.; Oreopoulou, V. Classification of phenolic compounds in plants. In Polyphenols in Plants; Academic Press: Cambridge, MA, USA, 2019; pp. 263–284. [Google Scholar]
  89. Santoso, J.; Yoshie, Y.; Suzuki, T. Polyphenolic compounds from seaweeds: Distribution and their antioxidative effect. In Developments in Food Science; Elsevier: Amsterdam, The Netherlands, 2004; Volume 42, pp. 169–177. [Google Scholar]
  90. Generalić Mekinić, I.; Skroza, D.; Šimat, V.; Hamed, I.; Čagalj, M.; Popović Perković, Z. Phenolic content of brown algae (Pheophyceae) species: Extraction, identification, and quantification. Biomolecules 2019, 9, 244. [Google Scholar] [CrossRef]
  91. Ishii, T.; Okino, T.; Suzuki, M.; Machiguchi, Y. Tichocarpols A and B, Two Novel Phenylpropanoids with Feeding-Deterrent Activity from the Red Alga Tichocarpus c rinitus. J. Nat. Prod. 2004, 67, 1764–1766. [Google Scholar] [CrossRef] [PubMed]
  92. Hartmann, A.; Ganzera, M.; Karsten, U.; Skhirtladze, A.; Stuppner, H. Phytochemical and analytical characterization of novel sulfated coumarins in the marine green macroalga Dasycladus vermicularis (Scopoli) krasser. Molecules 2018, 23, 2735. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Gupta, S.; Abu-Ghannam, N. Recent developments in the application of seaweeds or seaweed extracts as a means for enhancing the safety and quality attributes of foods. Innov. Food Sci. Emerg. Technol. 2011, 12, 600–609. [Google Scholar] [CrossRef]
  94. Handique, J.G.; Baruah, J.B. Polyphenolic compounds: An overview. React. Funct. Polym. 2002, 52, 163–188. [Google Scholar] [CrossRef]
  95. Mouritsen, O.G. The science of seaweeds: Marine macroalgae benefit people culturally, industrially, nutritionally, and ecologically. Am. Sci. 2013, 101, 458–466. [Google Scholar] [CrossRef]
  96. Freile-Pelegrín, Y.; Robledo, D. Bioactive phenolic compounds from algae. In Bioactive Compounds from Marine Foods: Plant and Animal Sources; John Wiley & Sons Ltd.: Chichester, UK, 2013; pp. 113–129. [Google Scholar]
  97. Kim, M.M.; Kim, S.K. Effect of phloroglucinol on oxidative stress and inflammation. Food Chem. Toxicol. 2010, 48, 2925–2933. [Google Scholar] [CrossRef]
  98. Charoensiddhi, S.; Franco, C.; Su, P.; Zhang, W. Improved antioxidant activities of brown seaweed Ecklonia radiata extracts prepared by microwave-assisted enzymatic extraction. J. Appl. Phycol. 2015, 27, 2049–2058. [Google Scholar] [CrossRef]
  99. Chang, M.Y.; Byon, S.H.; Shin, H.C.; Han, S.E.; Kim, J.Y.; Byun, J.Y.; Lee, J.D.; Park, M.K. Protective effects of the seaweed phlorotannin polyphenolic compound dieckol on gentamicin-induced damage in auditory hair cells. Int. J. Pediatr. Otorhinolaryngol. 2016, 83, 31–36. [Google Scholar] [CrossRef]
  100. Piao, M.J.; Hewage, S.R.; Han, X.; Kang, K.A.; Kang, H.K.; Lee, N.H.; Hyun, J.W. Protective Effect of Diphlorethohydroxycarmalol against Ultraviolet B Radiation-Induced DNA Damage by Inducing the Nucleotide Excision Repair System in HaCaT Human Keratinocytes. Mar. Drugs 2015, 13, 5629–5641. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  101. Kang, S.M.; Heo, S.J.; Kim, K.N.; Lee, S.H.; Yang, H.M.; Kim, A.D.; Jeon, Y.J. Molecular docking studies of a phlorotannin, dieckol isolated from Ecklonia cava with tyrosinase inhibitory activity. Bioorg. Med. Chem. 2012, 20, 311–316. [Google Scholar] [CrossRef] [PubMed]
  102. Kirke, D.A.; Smyth, T.J.; Rai, D.K.; Kenny, O.; Stengel, D.B. The chemical and antioxidant stability of isolated low molecular weight phlorotannins. Food Chem. 2017, 221, 1104–1112. [Google Scholar] [CrossRef] [PubMed]
  103. de Lima Cherubim, D.J.; Buzanello Martins, C.V.; Oliveira Fariña, L.; da Silva de Lucca, R.A. Polyphenols as natural antioxidants in cosmetics applications. J. Cosmet. Dermatol. 2020, 19, 33–37. [Google Scholar] [CrossRef] [PubMed]
  104. 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] [PubMed]
  105. Kim, E.K.; Tang, Y.; Kim, Y.S.; Hwang, J.W.; Choi, E.J.; Lee, J.H.; Lee, S.H.; Jeon, Y.J.; Park, P.J. First evidence that Ecklonia cava-derived dieckol attenuates MCF-7 human breast carcinoma cell migration. Mar. Drugs 2015, 13, 1785–1797. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. 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]
  107. Lee, Y.J.; Park, J.H.; Park, S.A.; Joo, N.R.; Lee, B.H.; Lee, K.B.; Oh, S.M. Dieckol or phlorofucofuroeckol extracted from Ecklonia cava suppresses lipopolysaccharide-mediated human breast cancer cell migration and invasion. J. Appl. Phycol. 2020, 32, 631–640. [Google Scholar] [CrossRef]
  108. Li, Y.; Qian, Z.J.; Kim, M.M.; Kim, S.K. Cytotoxic activities of phlorethol and fucophlorethol derivatives isolated from Laminariaceae Ecklonia cava. J. Food Biochem. 2011, 35, 357–369. [Google Scholar] [CrossRef]
  109. Gager, L.; Lalegerie, F.; Connan, S.; Stiger-Pouvreau, V. Marine Algal Derived Phenolic Compounds and their Biological Activities for Medicinal and Cosmetic Applications. In Recent Advances in Micro and Macroalgal Processing: Food and Health Perspectives; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2021; pp. 278–334. [Google Scholar]
  110. Shibata, T.; Ishimaru, K.; Kawaguchi, S.; Yoshikawa, H.; Hama, Y. Antioxidant activities of phlorotannins isolated from Japanese Laminariaceae. In Nineteenth International Seaweed Symposium; Springer: Dordrecht, The Netherlands, 2007; pp. 255–261. [Google Scholar]
  111. Besednova, N.N.; Zvyagintseva, T.N.; Kuznetsova, T.A.; Makarenkova, I.D.; Smolina, T.P.; Fedyanina, L.N.; Kryzhanovsky, S.P.; Zaporozhets, T.S. Marine algae metabolites as promising therapeutics for the prevention and treatment of HIV/AIDS. Metabolites 2019, 9, 87. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Kim, A.R.; Lee, M.S.; Choi, J.W.; Utsuki, T.; Kim, J.I.; Jang, B.C.; Kim, H.R. Phlorofucofuroeckol A suppresses expression of inducible nitric oxide synthase, cyclooxygenase-2, and pro-inflammatory cytokines via inhibition of nuclear factor-κB, c-Jun NH2-terminal kinases, and Akt in microglial cells. Inflammation 2013, 36, 259–271. [Google Scholar] [CrossRef]
  113. 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]
  114. Kumar, L.R.; Paul, P.T.; Anas, K.K.; Tejpal, C.S.; Chatterjee, N.S.; Anupama, T.K.; Mathew, S.; Ravishankar, C.N. Phlorotannins–bioactivity and extraction perspectives. J. Appl. Phycol. 2022, 34, 2173–2185. [Google Scholar] [CrossRef]
  115. Shibata, T.; Fujimoto, K.; Nagayama, K.; Yamaguchi, K.; Nakamura, T. Inhibitory activity of brown algal phlorotannins against hyaluronidase. Int. J. Food Sci. Technol. 2002, 37, 703–709. [Google Scholar] [CrossRef]
  116. Kang, H.S.; Kim, H.R.; Byun, D.S.; Son, B.W.; Nam, T.J.; Choi, J.S. Tyrosinase inhibitors isolated from the edible brown alga Ecklonia stolonifera. Arch. Pharmacal Res. 2004, 27, 1226–1232. [Google Scholar] [CrossRef] [PubMed]
  117. Yoon, N.Y.; Eom, T.K.; Kim, M.M.; Kim, S.K. Inhibitory effect of phlorotannins isolated from Ecklonia cava on mushroom tyrosinase activity and melanin formation in mouse B16F10 melanoma cells. J. Agric. Food Chem. 2009, 57, 4124–4129. [Google Scholar] [CrossRef] [PubMed]
  118. Lee, S.H.; Kang, S.M.; Sok, C.H.; Hong, J.T.; Oh, J.Y.; Jeon, Y.J. Cellular activities and docking studies of eckol isolated from Ecklonia cava (Laminariales, Phaeophyceae) as potential tyrosinase inhibitor. Algae 2015, 30, 163–170. [Google Scholar]
  119. 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. In Vitro 2009, 23, 1123–1130. [Google Scholar] [CrossRef] [PubMed]
  120. Nurrochmad, A.; Wirasti, W.; Dirman, A.; Lukitaningsih, E.; Rahmawati, A.; Fakhrudin, N. Effects of Antioxidant, Anti-Collagenase, Anti-Elastase, Anti-Tyrosinase of The Extract and Fraction From Turbinaria decurrens Bory. Indones. J. Pharm. 2018, 29, 188. [Google Scholar] [CrossRef]
  121. Bak, S.S.; Sung, Y.K.; Kim, S.K. 7-Phloroeckol promotes hair growth on human follicles in vitro. Naunyn-Schmiedeberg’s Arch. Pharmacol. 2014, 387, 789–793. [Google Scholar] [CrossRef]
  122. Pérez, M.J.; Falqué, E.; Domínguez, H. Antimicrobial action of compounds from marine seaweed. Mar. Drugs 2016, 14, 52. [Google Scholar] [CrossRef] [Green Version]
  123. Cherry, P.; O’Hara, C.; Magee, P.J.; McSorley, E.M.; Allsopp, P.J. Risks and benefits of consuming edible seaweeds. Nutr. Rev. 2019, 77, 307–329. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Holdt, S.L.; Kraan, S. Bioactive compounds in seaweed: Functional food applications and legislation. J. Appl. Phycol. 2011, 23, 543–597. [Google Scholar] [CrossRef]
  125. De Almeida, C.L.F.; Falcão, H.D.S.; Lima, G.R.D.M.; Montenegro, C.D.A.; Lira, N.S.; de Athayde-Filho, P.F.; Rodrigues, L.C.; De Souza, M.D.F.V.; Barbosa-Filho, J.M.; Batista, L.M. Bioactivities from marine algae of the genus Gracilaria. Int. J. Mol. Sci. 2011, 12, 4550–4573. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Abu-Ghannam, N.; Rajauria, G. Antimicrobial activity of compounds isolated from algae. In Functional Ingredients from Algae for Foods and Nutraceuticals; Woodhead Publishing: Sawston, UK, 2013; pp. 287–306. [Google Scholar]
  127. Duan, X.J.; Li, X.M.; Wang, B.G. Highly brominated mono-and bis-phenols from the marine red alga Symphyocladia latiuscula with radical-scavenging activity. J. Nat. Prod. 2007, 70, 1210–1213. [Google Scholar] [CrossRef]
  128. Choi, J.S.; Park, H.J.; Jung, H.A.; Chung, H.Y.; Jung, J.H.; Choi, W.C. A cyclohexanonyl bromophenol from the red alga Symphyocladia latiuscula. J. Nat. Prod. 2000, 63, 1705–1706. [Google Scholar] [CrossRef]
  129. Colon, M.; Guevara, P.; Gerwick, W.H.; Ballantine, D. 5’-Hydroxyisoavrainvilleol, a new diphenylmethane derivative from the tropical green alga Avrainvillea nigricans. J. Nat. Prod. 1987, 50, 368–374. [Google Scholar] [CrossRef]
  130. Carte, B.K.; Troupe, N.; Chan, J.A.; Westley, J.W.; Faulkner, D.J. Rawsonol, an inhibitor of HMG-CoA reductase from the tropical green alga Avrainvillea rawsoni. Phytochemistry 1989, 28, 2917–2919. [Google Scholar] [CrossRef]
  131. Estrada, D.M.; Martin, J.D.; Perez, C. A new brominated monoterpenoid quinol from Cymopolia barbata. J. Nat. Prod. 1987, 50, 735–737. [Google Scholar] [CrossRef]
  132. Tanna, B.; Choudhary, B.; Mishra, A. Metabolite profiling, antioxidant, scavenging and anti-proliferative activities of selected tropical green seaweeds reveal the nutraceutical potential of Caulerpa spp. Algal Res. 2018, 36, 96–105. [Google Scholar] [CrossRef]
  133. Jeyaprakash, R.R.K. HPLC Analysis of flavonoids in Acanthophora specifera (red seaweed) collected from Gulf of Mannar, Tamilnadu, India. Int. J. Sci. Res. 2017, 6, 69–72. [Google Scholar]
  134. Cotas, J.; Leandro, A.; Monteiro, P.; Pacheco, D.; Figueirinha, A.; Gonçalves, A.M.; da Silva, G.J.; Pereira, L. Seaweed phenolics: From extraction to applications. Mar. Drugs 2020, 18, 384. [Google Scholar] [CrossRef] [PubMed]
  135. Arct, J.; Pytkowska, K. Flavonoids as components of biologically active cosmeceuticals. Clin. Dermatol. 2008, 26, 347–357. [Google Scholar] [CrossRef] [PubMed]
  136. Masaki, H. Role of antioxidants in the skin: Anti-aging effects. J. Dermatol. Sci. 2010, 58, 85–90. [Google Scholar] [CrossRef]
  137. Makkar, F.; Chakraborty, K. Highly oxygenated antioxidative 2 H-chromen derivative from the red seaweed Gracilaria opuntia with pro-inflammatory cyclooxygenase and lipoxygenase inhibitory properties. Nat. Prod. Res. 2018, 32, 2756–2765. [Google Scholar] [CrossRef] [PubMed]
  138. Pillai, S.; Oresajo, C.; Hayward, J. Ultraviolet radiation and skin aging: Roles of reactive oxygen species, inflammation and protease activation, and strategies for prevention of inflammation-induced matrix degradation—A review. Int. J. Cosmet. Sci. 2005, 27, 17–34. [Google Scholar] [CrossRef] [PubMed]
  139. Joe, C.; Robert, M.C. 1.16—Sesquiterpenes. In Comprehensive Natural Products II; Ben, H.-W., Mander, L.L., Eds.; Elsevier: Amsterdam, The Netherlands, 2010; Volume 1, pp. 609–641. ISBN 9780080453828. [Google Scholar] [CrossRef]
  140. Ruberto, G.; Baratta, M.T. Antioxidant activity of selected essential oil components in two lipid model systems. Food Chem. 2000, 69, 167–174. [Google Scholar] [CrossRef]
  141. Carreto, J.I.; Carignan, M.O. Mycosporine-like amino acids: Relevant secondary metabolites. Chemical and ecological aspects. Mar. Drugs 2011, 9, 387–446. [Google Scholar] [CrossRef] [Green Version]
  142. Rosic, N.N.; Braun, C.; Kvaskoff, D. Extraction and Analysis of Mycosporine-Like Amino Acids in Marine Algae. In Natural Products from Marine Algae: Methods and Protocols; Stengel, D.B., Connan, S., Eds.; Springer: New York, NY, USA, 2015. [Google Scholar]
  143. Guihéneuf, F.; Gietl, A.; Stengel, D.B. Temporal and spatial variability of mycosporine-like amino acids and pigments in three edible red seaweeds from western Ireland. J. Appl. Phycol. 2018, 30, 2573–2586. [Google Scholar] [CrossRef]
  144. Suh, S.S.; Oh, S.K.; Lee, S.G.; Kim, I.C.; Kim, S. Porphyra-334, a mycosporine-like amino acid, attenuates UV-induced apoptosis in HaCaT cells. Acta Pharm. 2017, 67, 257–264. [Google Scholar] [CrossRef] [Green Version]
  145. Pavia, H.; Brock, E. Extrinsic factors influencing phlorotannin production in the brown alga Ascophyllum nodosum. Mar. Ecol. Prog. Ser. 2000, 193, 285–294. [Google Scholar] [CrossRef]
  146. Jormalainen, V.; Honkanen, T. Variation in natural selection for growth and phlorotannins in the brown alga Fucus vesiculosus. J. Evol. Biol. 2004, 17, 807–820. [Google Scholar] [CrossRef] [PubMed]
  147. Sheih, I.C.; Wu, T.K.; Fang, T.J. Antioxidant properties of a new antioxidative peptide from algae protein waste hydrolysate in different oxidation systems. Bioresour. Technol. 2009, 100, 3419–3425. [Google Scholar] [CrossRef] [PubMed]
  148. Li, Y.; Qian, Z.J.; Ryu, B.; Lee, S.H.; Kim, M.M.; Kim, S.K. Chemical components and its antioxidant properties in vitro: An edible marine brown alga, Ecklonia cava. Bioorg. Med. Chem. 2009, 17, 1963–1973. [Google Scholar] [CrossRef] [PubMed]
  149. Arunkumar, K.; Raj, R.; Raja, R.; Carvalho, I.S. Brown seaweeds as a source of anti-hyaluronidase compounds. S. Afr. J. Bot. 2021, 139, 470–477. [Google Scholar] [CrossRef]
  150. Riani, M.K.L.; Anwar, E.; Nurhayati, T. Antioxidant and anti-collagenase activity of Sargassum plagyophyllum extract as an anti-wrinkle cosmetic ingredient. Pharmacogn. J. 2018, 10, 932–936. [Google Scholar]
  151. Busetti, A.; Maggs, C.A.; Gilmore, B.F. Marine macroalgae and their associated microbiomes as a source of antimicrobial chemical diversity. Eur. J. Phycol. 2017, 52, 452–465. [Google Scholar] [CrossRef]
  152. Handelman, G.J. The evolving role of carotenoids in human biochemistry. Nutrition 2001, 17, 818–822. [Google Scholar] [CrossRef]
  153. Wang, T.; Jonsdottir, R.; Ólafsdóttir, G. Total phenolic compounds, radical scavenging and metal chelation of extracts from Icelandic seaweeds. Food Chem. 2009, 116, 240–248. [Google Scholar] [CrossRef]
  154. Lee, J.H.; Kim, G.H. Evaluation of antioxidant activity of marine algae-extracts from Korea. J. Aquat. Food Prod. Technol. 2015, 24, 227–240. [Google Scholar] [CrossRef]
  155. Yoshie-Stark, Y.; Hsieh, Y.P.; Suzuki, T. Distribution of flavonoids and related compounds from seaweeds in Japan. J.-Tokyo Univ. Fish. 2003, 89, 1–6. [Google Scholar]
  156. Cho, S.H.; Kang, S.E.; Cho, J.Y.; Kim, A.R.; Park, S.M.; Hong, Y.K.; Ahn, D.H. The antioxidant properties of brown seaweed (Sargassum siliquastrum) extracts. J. Med. Food 2007, 10, 479–485. [Google Scholar] [CrossRef] [PubMed]
  157. Lola-Luz, T.; Hennequart, F.; Gaffney, M. Enhancement of phenolic and flavonoid compounds in cabbage (Brassica oleraceae) following application of commercial seaweed extracts of the brown seaweed, (Ascophyllum nodosum). Agric. Food Sci. 2013, 22, 288–295. [Google Scholar] [CrossRef] [Green Version]
  158. 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] [PubMed] [Green Version]
  159. Yan, X.; Yang, C.; Lin, G.; Chen, Y.; Miao, S.; Liu, B.; Zhao, C. Antidiabetic potential of green seaweed Enteromorpha prolifera flavonoids regulating insulin signaling pathway and gut microbiota in type 2 diabetic mice. J. Food Sci. 2019, 84, 165–173. [Google Scholar] [CrossRef] [PubMed]
  160. Farvin, K.S.; Jacobsen, C. Phenolic compounds and antioxidant activities of selected species of seaweeds from Danish coast. Food Chem. 2013, 138, 1670–1681. [Google Scholar] [CrossRef]
  161. Chakraborty, K.; Maneesh, A.; Makkar, F. Antioxidant activity of brown seaweeds. J. Aquat. Food Prod. Technol. 2017, 26, 406–419. [Google Scholar] [CrossRef]
  162. Vimaladevi, S.; Mahesh, A.; Dhayanithi, B.N.; Karthikeyan, N. Mosquito larvicidal efficacy of phenolic acids of seaweed Chaetomorpha antennina (Bory) Kuetz. against Aedes aegypti. Biologia 2012, 67, 212–216. [Google Scholar] [CrossRef]
  163. 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]
  164. Wang, H.M.D.; Li, X.C.; Lee, D.J.; Chang, J.S. Potential biomedical applications of marine algae. Bioresour. Technol. 2017, 244, 1407–1415. [Google Scholar] [CrossRef]
  165. Guillerme, J.B.; Couteau, C.; Coiffard, L. Applications for marine resources in cosmetics. Cosmetics 2017, 4, 35. [Google Scholar] [CrossRef] [Green Version]
  166. Anyanwu, R.C.; Rodriguez, C.; Durrant, A.; Olabi, A.G. Micro-Macroalgae Properties and Applications. In Reference Module in Materials Science and Materials Engineering; Elsevier: Amsterdam, The Netherlands, 2018; pp. 1–28. ISBN 9780128035818. [Google Scholar]
  167. Matanjun, P.; Mohamed, S.; Mustapha, N.M.; Muhammad, K.; Ming, C.H. Antioxidant activities and phenolics content of eight species of seaweeds from north Borneo. J. Appl. Phycol. 2008, 20, 367–373. [Google Scholar] [CrossRef]
  168. Cho, M.; Lee, H.S.; Kang, I.J.; Won, M.H.; You, S. Antioxidant properties of extract and fractions from Enteromorpha prolifera, a type of green seaweed. Food Chem. 2011, 127, 999–1006. [Google Scholar] [CrossRef] [PubMed]
  169. Wang, L.; Oh, J.Y.; Lee, W.; Jeon, Y.J. Fucoidan isolated from Hizikia fusiforme suppresses ultraviolet B-induced photodamage by down-regulating the expressions of matrix metalloproteinases and pro-inflammatory cytokines via inhibiting NF-κB, AP-1, and MAPK signaling pathways. Int. J. Biol. Macromol. 2021, 166, 751–759. [Google Scholar] [CrossRef] [PubMed]
  170. Pimentel, F.B.; Alves, R.C.; Rodrigues, F.; PP Oliveira, M.B. Macroalgae-derived ingredients for cosmetic industry—An update. Cosmetics 2017, 5, 2. [Google Scholar] [CrossRef]
  171. Bedoux, G.; Hardouin, K.; Burlot, A.S.; Bourgougnon, N. Bioactive components from seaweeds: Cosmetic applications and future development. In Advances in Botanical Research; Academic Press: Cambridge, MA, USA, 2014; Volume 71, pp. 345–378. [Google Scholar]
  172. Michalak, I.; Chojnacka, K. Algae as production systems of bioactive compounds. Eng. Life Sci. 2015, 15, 160–176. [Google Scholar] [CrossRef]
  173. Green Confertii Extract-NS—The Garden of Naturalsolution—Datasheet. Available online: https://cosmetics.specialchem.com/product/i-natural-solution-green-confertii-extract-ns (accessed on 10 April 2020).
  174. Lopes, G.; Sousa, C.; Silva, L.R.; Pinto, E.; Andrade, P.B.; Bernardo, J.; Mouga, T.; Valentão, P. Can phlorotannins purified extracts constitute a novel pharmacological alternative for microbial infections with associated inflammatory conditions? PLoS ONE 2012, 7, e31145. [Google Scholar] [CrossRef]
  175. Cha, S.H.; Kim, E.A.; Kim, K.N.; Heo, S.J.; Jun, H.S.; Jeon, Y.J. Prolonged exposure of marine algal phlorotannins with whitening effect did not cause inflammatory hyperpigmentation in zebrafish larva. Gen 2020, 5, 1–6. [Google Scholar] [CrossRef]
  176. Thomas, N.V.; Kim, S.K. Potential pharmacological applications of polyphenolic derivatives from marine brown algae. Environ. Toxicol. Pharmacol. 2011, 32, 325–335. [Google Scholar] [CrossRef]
  177. Cho, S.H.; Kim, H.S.; Lee, W.; Han, E.J.; Kim, S.Y.; Fernando, I.S.; Ahn, G.; Kim, K.N. Eckol from Ecklonia cava ameliorates TNF-α/IFN-γ-induced inflammatory responses via regulating MAPKs and NF-κB signaling pathway in HaCaT cells. Int. Immunopharmacol. 2020, 82, 106146. [Google Scholar] [CrossRef]
  178. Imbs, T.I.; Zvyagintseva, T.N. Phlorotannins are polyphenolic metabolites of brown algae. Russ. J. Mar. Biol. 2018, 44, 263–273. [Google Scholar] [CrossRef]
  179. Sugiura, Y.; Takeuchi, Y.; Kakinuma, M.; Amano, H. Inhibitory effects of seaweeds on histamine release from rat basophile leukemia cells (RBL-2H3). Fish. Sci. 2006, 72, 1286–1291. [Google Scholar] [CrossRef]
  180. Shibata, T.; Yamaguchi, K.; Nagayama, K.; Kawaguchi, S.; Nakamura, T. Inhibitory activity of brown algal phlorotannins against glycosidases from the viscera of the turban shell Turbo cornutus. Eur. J. Phycol. 2002, 37, 493–500. [Google Scholar] [CrossRef] [Green Version]
  181. Joe, M.J.; Kim, S.N.; Choi, H.Y.; Shin, W.S.; Park, G.M.; Kang, D.W.; Kim, Y.K. The inhibitory effects of eckol and dieckol from Ecklonia stolonifera on the expression of matrix metalloproteinase-1 in human dermal fibroblasts. Biol. Pharm. Bull. 2006, 29, 1735–1739. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  182. Kim, A.R.; Lee, M.S.; Shin, T.S.; Hua, H.; Jang, B.C.; Choi, J.S.; Byun, D.S.; Utsuki, T.; Ingram, D.; Kim, H.R. Phlorofucofuroeckol A inhibits the LPS-stimulated iNOS and COX-2 expressions in macrophages via inhibition of NF-κB, Akt, and p38 MAPK. Toxicol. In Vitro 2011, 25, 1789–1795. [Google Scholar] [CrossRef]
  183. Kong, C.S.; Kim, J.; Ahn, B.N.; Kim, S.K. Potential effect of phloroglucinol derivatives from Ecklonia cava on matrix metalloproteinase expression and the inflammatory profile in lipopolysaccharide-stimulated human THP-1 macrophages. Fish. Sci. 2011, 77, 867–873. [Google Scholar] [CrossRef]
  184. Thomas, N.V.; Kim, S.K. Beneficial effects of marine algal compounds in cosmeceuticals. Mar. Drugs 2013, 11, 146–164. [Google Scholar] [CrossRef] [Green Version]
  185. Stengel, D.B.; Connan, S.; Popper, Z.A. Algal chemodiversity and bioactivity: Sources of natural variability and implications for commercial application. Biotechnol. Adv. 2011, 29, 483–501. [Google Scholar] [CrossRef]
  186. Borowitzka, M.A. High-value products from microalgae—Their development and commercialization. J. Appl. Phycol. 2013, 25, 743–756. [Google Scholar] [CrossRef]
  187. Ryu, B.; Li, Y.; Qian, Z.J.; Kim, M.M.; Kim, S.K. Exhibitory effects of compounds from brown alga Ecklonia cava on the human osteoblasts. J. Biotech. 2008, 136, 577–588. [Google Scholar] [CrossRef]
  188. Sanjeewa, K.K.A.; Kim, E.A.; Son, K.T.; Jeon, Y.J. Bioactive properties and potentials cosmeceutical applications of phlorotannins isolated from brown seaweeds: A review. J. Photochem. Photobiol. B Biol. 2016, 162, 100–105. [Google Scholar] [CrossRef]
  189. Ding, Y.; Jiang, Y.; Im, S.T.; Myung, S.; Kim, H.S.; Lee, S.H. Diphlorethohydroxycarmalol inhibits melanogenesis via protein kinase A/cAMP response element-binding protein and extracellular signal-regulated kinase-mediated microphthalmia-associated transcription factor downregulation in α-melanocyte stimulating hormone-stimulated B16F10 melanoma cells and zebrafish. Cell Biochem. Funct. 2021, 39, 546–554. [Google Scholar] [PubMed]
  190. Song, T.Y.; Chen, C.H.; Yang, N.C.; Fu, C.S. The correlation of in vitro mushroom tyrosinase activity with cellular tyrosinase activity and melanin formation in melanoma cells A2058. J. Food Drug Anal. 2009, 17, 4. [Google Scholar] [CrossRef]
  191. Shanura Fernando, I.P.; Asanka Sanjeewa, K.K.; Samarakoon, K.W.; Kim, H.S.; Gunasekara, U.K.D.S.S.; Park, Y.J.; Abeytunga, D.T.U.; Lee, W.W.; Jeon, Y.J. The potential of fucoidans from Chnoospora minima and Sargassum polycystum in cosmetics: Antioxidant, anti-inflammatory, skin-whitening, and antiwrinkle activities. J. Appl. Phycol. 2018, 30, 3223–3232. [Google Scholar] [CrossRef]
  192. Liu, M.; Hansen, P.E.; Lin, X. Bromophenols in marine algae and their bioactivities. Mar. Drugs 2011, 9, 1273–1292. [Google Scholar] [CrossRef]
  193. Heo, S.J.; Ko, S.C.; Kang, S.M.; Cha, S.H.; Lee, S.H.; Kang, D.H.; Jung, W.K.; Affan, A.; Oh, C.; Jeon, Y.J. Inhibitory effect of diphlorethohydroxycarmalol on melanogenesis and its protective effect against UV-B radiation-induced cell damage. Food Chem. Toxicol. 2010, 48, 1355–1361. [Google Scholar] [CrossRef]
  194. Swanson, A.K.; Druehl, L.D. Induction, exudation and the UV protective role of kelp phlorotannins. Aquat. Bot. 2002, 73, 241–253. [Google Scholar] [CrossRef]
  195. Soleimani, S.; Babaei Mahani Nezhad, S.; Yousefzadi, M. Role of Phlorotannins Derived from Brown Alga Padina sp. as a Protective Agent against Ultraviolet Radiation and Oxidative Stress. J. Mar. Med. 2021, 3, 97–106. [Google Scholar]
  196. Ko, S.C.; Cha, S.H.; Heo, S.J.; Lee, S.H.; Kang, S.M.; Jeon, Y.J. Protective effect of Ecklonia cava on UVB-induced oxidative stress: In vitro and in vivo zebrafish model. J. Appl. Phycol. 2011, 23, 697–708. [Google Scholar] [CrossRef]
  197. Kim, K.C.; Piao, M.J.; Zheng, J.; Yao, C.W.; Cha, J.W.; Kumara, M.H.S.R.; Han, X.; Kang, H.K.; Lee, N.H.; Hyun, J.W. Fucodiphlorethol G purified from Ecklonia cava suppresses ultraviolet B radiation-induced oxidative stress and cellular damage. Biomol. Ther. 2014, 22, 301. [Google Scholar] [CrossRef] [Green Version]
  198. Huovinen, P.; Gómez, I. UV sensitivity of vegetative and reproductive tissues of two Antarctic Brown Algae is related to differential allocation of phenolic substances. Photochem. Photobiol. 2015, 91, 1382–1388. [Google Scholar] [CrossRef]
  199. Chan, Y.Y.; Kim, K.H.; Cheah, S.H. Inhibitory effects of Sargassum polycystum on tyrosinase activity and melanin formation in B16F10 murine melanoma cells. J. Ethnopharmacol. 2011, 137, 1183–1188. [Google Scholar] [CrossRef] [PubMed]
  200. Moon, H.E.; Islam, M.N.; Ahn, B.R.; Chowdhury, S.S.; Sohn, H.S.; Jung, H.A.; Choi, J.S. Protein tyrosine phosphatase 1B and α-glucosidase inhibitory phlorotannins from edible brown algae, Ecklonia stolonifera and Eisenia bicyclis. Biosci. Biotechnol. Biochem. 2011, 75, 1472–1480. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  201. Cha, S.H.; Ko, S.C.; Kim, D.; Jeon, Y.J. Screening of marine algae for potential tyrosinase inhibitor: Those inhibitors reduced tyrosinase activity and melanin synthesis in zebrafish. J. Dermatol. 2011, 38, 354–363. [Google Scholar] [CrossRef] [PubMed]
  202. Kang, K.A.; Lee, K.H.; Chae, S.; Koh, Y.S.; Yoo, B.S.; Kim, J.H.; Ham, Y.M.; Baik, J.S.; Lee, N.H.; Hyun, J.W. Triphlorethol-A from Ecklonia cava protects V79-4 lung fibroblast against hydrogen peroxide induced cell damage. Free Radic. Res. 2005, 39, 883–892. [Google Scholar] [CrossRef] [PubMed]
  203. Lee, M.S.; Kwon, M.S.; Choi, J.W.; Shin, T.; No, H.K.; Choi, J.S.; Byun, D.S.; Kim, J.I.; Kim, H.R. Anti-inflammatory activities of an ethanol extract of Ecklonia stolonifera in lipopolysaccharide-stimulated RAW 264.7 murine macrophage cells. J. Agric. Food Chem. 2012, 60, 9120–9129. [Google Scholar] [CrossRef] [PubMed]
  204. Hwang, E.; Park, S.Y.; Sun, Z.W.; Shin, H.S.; Lee, D.G.; Yi, T.H. The protective effects of fucosterol against skin damage in UVB-irradiated human dermal fibroblasts. Mar. Biotechnol. 2014, 16, 361–370. [Google Scholar] [CrossRef] [PubMed]
  205. Kim, K.N.; Yang, H.M.; Kang, S.M.; Kim, D.; Ahn, G.; Jeon, Y.J. Octaphlorethol A isolated from Ishige foliacea inhibits α-MSH-stimulated induced melanogenesis via ERK pathway in B16F10 melanoma cells. Food Chem. Toxicol. 2013, 59, 521–526. [Google Scholar] [CrossRef] [PubMed]
  206. Kim, K.N.; Yang, H.M.; Kang, S.M.; Ahn, G.; Roh, S.W.; Lee, W.; Kim, D.; Jeon, Y.J. Whitening effect of octaphlorethol A isolated from Ishige foliacea in an in vivo zebrafish model. J. Microbiol. Biotechnol. 2015, 25, 448–451. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. del Olmo, A.; Picon, A.; Nuñez, M. High pressure processing for the extension of Laminaria ochroleuca (kombu) shelf-life: A comparative study with seaweed salting and freezing. Innov. Food Sci. Emerg. Technol. 2019, 52, 420–428. [Google Scholar] [CrossRef]
  208. Ariede, M.B.; Candido, T.M.; Jacome, A.L.M.; Velasco, M.V.R.; de Carvalho, J.C.M.; Baby, A.R. Cosmetic attributes of algae—A review. Algal Res. 2017, 25, 483–487. [Google Scholar] [CrossRef]
  209. Sappati, P.K.; Nayak, B.; VanWalsum, G.P.; Mulrey, O.T. Combined effects of seasonal variation and drying methods on the physicochemical properties and antioxidant activity of sugar kelp (Saccharina latissima). J. Appl. Phycol. 2019, 31, 1311–1332. [Google Scholar] [CrossRef]
  210. Azam, M.S.; Choi, J.; Lee, M.S.; Kim, H.R. Hypopigmenting effects of brown algae-derived phytochemicals: A review on molecular mechanisms. Mar. Drugs 2017, 15, 297. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  211. Seo, Y.; Park, K.E.; Kim, Y.A.; Lee, H.J.; Yoo, J.S.; Ahn, J.W.; Lee, B.J. Isolation of tetraprenyltoluquinols from the brown alga Sargassum thunbergii. Chem. Pharm. Bull. 2006, 54, 1730–1733. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  212. Cian, R.E.; Bacchetta, C.; Rossi, A.; Cazenave, J.; Drago, S.R. Red seaweed Pyropia columbina as antioxidant supplement in feed for cultured juvenile Pacú (Piaractus mesopotamicus). J. Appl. Phycol. 2019, 31, 1455–1465. [Google Scholar] [CrossRef]
  213. Li, K.; Li, X.M.; Gloer, J.B.; Wang, B.G. New nitrogen-containing bromophenols from the marine red alga Rhodomela confervoides and their radical scavenging activity. Food Chem. 2012, 135, 868–872. [Google Scholar] [CrossRef] [PubMed]
  214. Chew, Y.L.; Lim, Y.Y.; Omar, M.; Khoo, K.S. Antioxidant activity of three edible seaweeds from two areas in South East Asia. LWT-Food Sci. Technol. 2008, 41, 1067–1072. [Google Scholar] [CrossRef]
  215. Fernandes, H.; Salgado, J.M.; Martins, N.; Peres, H.; Oliva-Teles, A.; Belo, I. Sequential bioprocessing of Ulva rigida to produce lignocellulolytic enzymes and to improve its nutritional value as aquaculture feed. Bioresour. Technol. 2019, 281, 277–285. [Google Scholar] [CrossRef] [Green Version]
  216. Li, Y.; Xia, C.; Yao, G.; Zhang, X.; Zhao, J.; Gao, X.; Yong, J.; Wang, H. Protective effects of liquiritin on UVB-induced skin damage in SD rats. Int. Immunopharmacol. 2021, 97, 107614. [Google Scholar] [CrossRef]
  217. Manandhar, B.; Paudel, P.; Seong, S.H.; Jung, H.A.; Choi, J.S. Characterizing eckol as a therapeutic aid: A systematic review. Mar. Drugs 2019, 17, 361. [Google Scholar] [CrossRef] [Green Version]
  218. Kim, J.A.; Lee, J.M.; Shin, D.B.; Lee, N.H. The antioxidant activity and tyrosinase inhibitory activity of phloro-tannins in Ecklonia cava. Food Sci. Biotechnol. 2004, 13, 476–480. [Google Scholar]
  219. Okeke, E.S.; Nweze, E.J.; Chibuogwu, C.C.; Anaduaka, E.G.; Chukwudozie, K.I.; Ezeorba, T.P.C. Aquatic Phlorotannins and Human Health: Bioavailability, Toxicity, and Future Prospects. Nat. Prod. Commun. 2021, 16, 1934578X211056144. [Google Scholar] [CrossRef]
  220. Vo, T.S.; Kim, S.K.; Ryu, B.; Ngo, D.H.; Yoon, N.Y.; Bach, L.G.; Hang, N.T.N.; Ngo, D.N. The suppressive activity of fucofuroeckol-A derived from brown algal Ecklonia stolonifera Okamura on UVB-induced mast cell degranulation. Mar. Drugs 2018, 16, 1. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  221. 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]
  222. Hermund, D.B.; Plaza, M.; Turner, C.; Jónsdóttir, R.; Kristinsson, H.G.; Jacobsen, C.; Nielsen, K.F. Structure dependent antioxidant capacity of phlorotannins from Icelandic Fucus vesiculosus by UHPLC-DAD-ECD-QTOFMS. Food Chem. 2018, 240, 904–909. [Google Scholar] [CrossRef] [Green Version]
  223. Jang, J.; Ye, B.R.; Heo, S.J.; Oh, C.; Kang, D.H.; Kim, J.H.; Affan, A.; Yoon, K.T.; Choi, Y.U.; Park, S.C.; et al. Photo-oxidative stress by ultraviolet-B radiation and antioxidative defense of eckstolonol in human keratinocytes. Environ. Toxicol. Pharmacol. 2012, 34, 926–934. [Google Scholar] [CrossRef]
  224. 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]
  225. Kim, S.M.; Kang, K.; Jeon, J.S.; Jho, E.H.; Kim, C.Y.; Nho, C.W.; Um, B.H. Isolation of phlorotannins from Eisenia bicyclis and their hepatoprotective effect against oxidative stress induced by tert-butyl hyperoxide. Appl. Biochem. Biotechnol. 2011, 165, 1296–1307. [Google Scholar] [CrossRef] [PubMed]
  226. Li, Y.X.; Wijesekara, I.; Li, Y.; Kim, S.K. Phlorotannins as bioactive agents from brown algae. Process Biochem. 2011, 46, 2219–2224. [Google Scholar] [CrossRef]
  227. 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]
  228. Ferreres, F.; Lopes, G.; Gil-Izquierdo, A.; Andrade, P.B.; Sousa, C.; Mouga, T.; Valentão, P. Phlorotannin extracts from fucales characterized by HPLC-DAD-ESI-MSn: Approaches to hyaluronidase inhibitory capacity and antioxidant properties. Mar. Drugs 2012, 10, 2766–2781. [Google Scholar] [CrossRef] [Green Version]
  229. Le Lann, K.; Surget, G.; Couteau, C.; Coiffard, L.; Cérantola, S.; Gaillard, F.; Larnicol, M.; Zubia, M.; Guérard, F.; Poupart, N.; et al. Sunscreen, antioxidant, and bactericide capacities of phlorotannins from the brown macroalga Halidrys siliquosa. J. Appl. Phycol. 2016, 28, 3547–3559. [Google Scholar] [CrossRef] [Green Version]
  230. Gager, L.; Connan, S.; Molla, M.; Couteau, C.; Arbona, J.F.; Coiffard, L.; Cérantola, S.; Stiger-Pouvreau, V. Active phlorotannins from seven brown seaweeds commercially harvested in Brittany (France) detected by 1H NMR and in vitro assays: Temporal variation and potential valorization in cosmetic applications. J. Appl. Phycol. 2020, 32, 2375–2386. [Google Scholar] [CrossRef]
  231. Lee, M.S.; Yoon, H.D.; Kim, J.I.; Choi, J.S.; Byun, D.S.; Kim, H.R. Dioxinodehydroeckol inhibits melanin synthesis through PI3K/Akt signalling pathway in α-melanocyte-stimulating hormone-treated B16F10 cells. Exp. Dermatol. 2012, 21, 471–473. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  232. Santana-Gálvez, J.; Jacobo-Velázquez, D.A. Classification of phenolic compounds. In Phenolic Compounds in Food; CRC Press: Boca Raton, FL, USA, 2018; pp. 3–20. [Google Scholar]
  233. Padilla, M.; Palma, M.; Barroso, C.G. Determination of phenolics in cosmetic creams and similar emulsions. J. Chromatogr. A 2005, 1091, 83–88. [Google Scholar] [CrossRef] [PubMed]
  234. Kim, S.M.; Kang, S.W.; Jeon, J.S.; Jung, Y.J.; Kim, W.R.; Kim, C.Y.; Um, B.H. Determination of major phlorotannins in Eisenia bicyclis using hydrophilic interaction chromatography: Seasonal variation and extraction characteristics. Food Chem. 2013, 138, 2399–2406. [Google Scholar] [CrossRef] [PubMed]
  235. Stengel, D.B.; Connan, S. Natural products from marine algae: Methods and protocols. Nat. Prod. Mar. Algae Methods Protoc. 2015, 1308, 1–439. [Google Scholar]
  236. Santos-Buelga, C.; Gonzalez-Manzano, S.; Dueñas, M.; Gonzalez-Paramas, A.M. Extraction and isolation of phenolic compounds. In Natural Products Isolation; Springer: Berlin/Heidelberg, Germany, 2012; pp. 427–464. [Google Scholar]
  237. Vieira, V.; Prieto, M.A.; Barros, L.; Coutinho, J.A.; Ferreira, I.C.; Ferreira, O. Enhanced extraction of phenolic compounds using choline chloride based deep eutectic solvents from Juglans regia L. Ind. Crops Prod. 2018, 115, 261–271. [Google Scholar] [CrossRef]
  238. Ospina, M.; Castro-Vargas, H.I.; Parada-Alfonso, F. Antioxidant capacity of Colombian seaweeds: 1. extracts obtained from Gracilaria mammillaris by means of supercritical fluid extraction. J. Supercrit. Fluids 2017, 128, 314–322. [Google Scholar] [CrossRef]
  239. Matulja, D.; Vranješević, F.; Kolympadi Markovic, M.; Pavelić, S.K.; Marković, D. Anticancer Activities of Marine-Derived Phenolic Compounds and Their Derivatives. Molecules 2022, 27, 1449. [Google Scholar] [CrossRef] [PubMed]
  240. Dai, J.; Mumper, R.J. Plant phenolics: Extraction, analysis and their antioxidant and anticancer properties. Molecules 2010, 15, 7313–7352. [Google Scholar] [CrossRef] [PubMed]
  241. Otero, P.; López-Martínez, M.I.; García-Risco, M.R. Application of pressurized liquid extraction (PLE) to obtain bioactive fatty acids and phenols from Laminaria ochroleuca collected in Galicia (NW Spain). J. Pharm. Biomed. Anal. 2019, 164, 86–92. [Google Scholar] [CrossRef] [PubMed]
  242. Kalil, S.J.; Moraes, C.C.; Sala, L.; Burkert, C.A. Bioproduct extraction from microbial cells by conventional and nonconventional techniques. In Food Bioconversion; Academic Press: Cambridge, MA, USA, 2017; pp. 179–206. [Google Scholar]
  243. Cikoš, A.M.; Jokić, S.; Šubarić, D.; Jerković, I. Overview on the application of modern methods for the extraction of bioactive compounds from marine macroalgae. Mar. Drugs 2018, 16, 348. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  244. Pinela, J.; Prieto, M.A.; Carvalho, A.M.; Barreiro, M.F.; Oliveira, M.B.P.; Barros, L.; Ferreira, I.C. Microwave-assisted extraction of phenolic acids and flavonoids and production of antioxidant ingredients from tomato: A nutraceutical-oriented optimization study. Sep. Purif. Technol. 2016, 164, 114–124. [Google Scholar] [CrossRef] [Green Version]
  245. Heleno, S.A.; Diz, P.; Prieto, M.A.; Barros, L.; Rodrigues, A.; Barreiro, M.F.; Ferreira, I.C. Optimization of ultrasound-assisted extraction to obtain mycosterols from Agaricus bisporus L. by response surface methodology and comparison with conventional Soxhlet extraction. Food Chem. 2016, 197, 1054–1063. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  246. Herrero, M.; Cifuentes, A.; Ibañez, E. Sub-and supercritical fluid extraction of functional ingredients from different natural sources: Plants, food-by-products, algae and microalgae: A review. Food Chem. 2006, 98, 136–148. [Google Scholar] [CrossRef] [Green Version]
  247. Zakaria, S.M.; Kamal, S.M.M. Subcritical water extraction of bioactive compounds from plants and algae: Applications in pharmaceutical and food ingredients. Food Eng. Rev. 2016, 8, 23–34. [Google Scholar] [CrossRef]
  248. Lourenço-Lopes, C.; Fraga-Corral, M.; Jimenez-Lopez, C.; Pereira, A.G.; Garcia-Oliveira, P.; Carpena, M.; Prieto, M.A.; SimalGandara, J. Metabolites from Macroalgae and Its Applications in the Cosmetic Industry: A Circular Economy Approach. Resources 2020, 9, 101. [Google Scholar] [CrossRef]
  249. Torres, P.; Santos, J.P.; Chow, F.; Dos Santos, D.Y. A comprehensive review of traditional uses, bioactivity potential, and chemical diversity of the genus Gracilaria (Gracilariales, Rhodophyta). Algal Res. 2019, 37, 288–306. [Google Scholar] [CrossRef]
Figure 1. (au) Examples of marine macroalgae. (a) Chaetomorpha antennina (green); (b) Corallina sp. (red); (c) Sargassum linearifolium (brown); (d) Palisada perforata (red); (e) Sargassum cinereum; (f) Palisada perforata (as Laurencia papillosa); (g) Laurencia glandulifera (red); (h) Sargassum tenerrimum (brown); (i) Sargassum tenerrimum (brown); (j) Ulva lactuca (green); (k) Laurencia sp. (red); (l) Gracilaria sp. (red); (m) Gracilaria debilis (red); (n) Udotea indica (green); (o) Champia compressa (red); (p) Tricleocarpa fragilis (red); (q) Caulerpa taxifolia (green); (r) Botryocladia leptopoda (red); (s) Centroceras clavulatum (red); (t) Scinaia moniliformis (red); (u) Padina tetrastromatica (brown).
Figure 1. (au) Examples of marine macroalgae. (a) Chaetomorpha antennina (green); (b) Corallina sp. (red); (c) Sargassum linearifolium (brown); (d) Palisada perforata (red); (e) Sargassum cinereum; (f) Palisada perforata (as Laurencia papillosa); (g) Laurencia glandulifera (red); (h) Sargassum tenerrimum (brown); (i) Sargassum tenerrimum (brown); (j) Ulva lactuca (green); (k) Laurencia sp. (red); (l) Gracilaria sp. (red); (m) Gracilaria debilis (red); (n) Udotea indica (green); (o) Champia compressa (red); (p) Tricleocarpa fragilis (red); (q) Caulerpa taxifolia (green); (r) Botryocladia leptopoda (red); (s) Centroceras clavulatum (red); (t) Scinaia moniliformis (red); (u) Padina tetrastromatica (brown).
Applsci 12 11954 g001aApplsci 12 11954 g001b
Figure 2. (ac). Chemical structures of Cladophora socialis (green)-derived phenolic compounds (a): 2,3,8,9-tetrahydroxybenzo[c]chromen-6-one; (b): 3,4,3′,4′-Tetrahydroxy-1,1′-biphenyl; (c): New phenolic natural compound. Source (permission obtained from the authors): Lavoie et al. [84].
Figure 2. (ac). Chemical structures of Cladophora socialis (green)-derived phenolic compounds (a): 2,3,8,9-tetrahydroxybenzo[c]chromen-6-one; (b): 3,4,3′,4′-Tetrahydroxy-1,1′-biphenyl; (c): New phenolic natural compound. Source (permission obtained from the authors): Lavoie et al. [84].
Applsci 12 11954 g002aApplsci 12 11954 g002b
Figure 3. Classification of phenolic compounds listed according to their chemical structures. Source: Cotas et al. [36].
Figure 3. Classification of phenolic compounds listed according to their chemical structures. Source: Cotas et al. [36].
Applsci 12 11954 g003
Figure 4. Chemical structures of some phenolic compounds. (A): Catechol, (B): Gallic acid, (C): Gentisic acid, (D): Phloroglucinol, (E): Hydroquinone (HQ), (F): Chlorogenic acid, (G): Coumaric acid, (H): 4-Hydroxybenzoic acid, (I): (+)-Catechin, (J): Caffeic acid, (K): Ferulic acid, (L): Sinapic acid, (M): 7-Hydroxy coumarin sulfate, (N): Eckol, (O): Dieckol, (P): Eckstolonol, (Q): Triphloroethol-A, (R): Epigallocatechin, (S): Catechin 3-O-gallate.
Figure 4. Chemical structures of some phenolic compounds. (A): Catechol, (B): Gallic acid, (C): Gentisic acid, (D): Phloroglucinol, (E): Hydroquinone (HQ), (F): Chlorogenic acid, (G): Coumaric acid, (H): 4-Hydroxybenzoic acid, (I): (+)-Catechin, (J): Caffeic acid, (K): Ferulic acid, (L): Sinapic acid, (M): 7-Hydroxy coumarin sulfate, (N): Eckol, (O): Dieckol, (P): Eckstolonol, (Q): Triphloroethol-A, (R): Epigallocatechin, (S): Catechin 3-O-gallate.
Applsci 12 11954 g004
Figure 5. Techniques of extraction and characterization for phenolic compounds. Source: Santana-Gálvez and Jacobo-Velázquez [232].
Figure 5. Techniques of extraction and characterization for phenolic compounds. Source: Santana-Gálvez and Jacobo-Velázquez [232].
Applsci 12 11954 g005
Table 1. Applications of marine macroalgae-derived phenolic compounds in skin benefits.
Table 1. Applications of marine macroalgae-derived phenolic compounds in skin benefits.
No.Name of Marine AlgaeSeaweed-Based Bioactive CompoundsCosmetic Properties/Benefits References
1Macroalgal speciesCatechins, Flavanols, Flavanol glycosides, Gallic acid, Epicatechin, Phloroglucinol, Pyro catechol, Gallate, Flavonoids, Anthocyanins, Stilbenes, LignansMatrix Metalloproteinase (MMP) inhibitors, Reduce collagen degradation[145,146,147]
2Corallina pilulifera (R)-Inhibit the expression of MMP2 and MMP-9[148]
3Sargassum horneri (B)Sargachromanol EAntiaging[149]
4Phycocalidia vietnamensis (R)Mycosporine-like amino acids (MAAs)UV absorber[150]
5Ecklonia cava (B)PhlorotanninsSkin whitener, Tyrosinase inhibition[151]
6Macroalgal species-Antioxidant activity[152,153]
7Macroalgal speciesPhlorotanninsAnti-wrinkle, Antiaging[154,155]
8Sargassum fusiforme
(as Hizikia fusiformis) (B)
PhlorotanninsTyrosinase inhibition, Skin
whitener
[156]
9Corallina pilulifera (R)Phlorotannins, Eckol, Fucols, Fucophorethols, Fuhalols, PhloretholsAntiaging, Antiphotoaging, Antioxidant, Tyrosinase inhibition[157,158,159,160]
10Macroalgal speciesPhlorotanninsInhibit melanin synthesis, Anti UVB photodamage[161]
11Ecklonia cava (B)PhlorotanninsMelanin synthesis, UV protector[162,163]
12Ecklonia cava (B)Phlorotannins such as eckstolonol, dieckolAntioxidant, photoprotective, UV protector[164]
13Brown algae speciesPhlorotannins such as Phloroeckol, Tetrameric phloroglucinolAnti-skin aging, Antioxidant[165]
14Corallina pilulifera (R)PhlorotanninsMetalloproteinase inhibitors, and UV protectors, Prevent collagen degradation, Wrinkle formation[166]
15Ulva clathrata, Ulva compressa (as Enteromorpha compressa), Ulva intestinalis, Ulva linza, Ulva flexuosa, Ulva australis, Capsosiphon fulvescens, Chaetomorpha moniligera (G)Bromophenols, FlavonoidsHighly radical scavenger[167,168]
16E. cava (B)PhlorotanninsUVB protector[169]
17Saccharina japonica
(as Laminaria japonica), Ecklonia cava, (B)
PhlorotanninsUtilized in facial masks, UV protectors, Anti-acne[170,171,172]
18Ulva compressa (as Enteromorpha compressa) (G)Flavonoids, Tannins, PhlorotanninsAntioxidant effect, Anti-aging[173]
19Fucus vesiculosus (B)Flavonoids, Phenols, HQ, SaponinTyrosinase inhibitor, Melanin Inhibition[174]
20Ecklonia cava (B)Phlorotannins; Eckol, Dieckol, Dioxinodehydroeckol, 7-
phloroeckol, Phloroglucinol
Tyrosinase inhibition (Skin whitener)[175,176,177]
21Ericaria selaginoides (as Cystoseira tamariscifolia), Gongolaria usneoides (as Cystoseira usneoides), Fucus spiralis, Gongolaria nodicaulis (as Cystoseira nodicaulis (B)Phlorotannins, Fucophloroethol, Bieckol, Phlorofucofuroeckol, 7-phloroeckolAntioxidant, Anti-aging, anti-wrinkling, Hyaluronidase inhibition, Lipid peroxidase inhibition[178]
22Ecklonia bicyclis
(as Eisenia bicyclis) (B)
Phlorotannins (Phlorofucofuroeckol-
A, Dieckol, Eckol, Phloroglucinol, 8,8′ bieckol
Hyaluronidase inhibitor, Anti-wrinkle[179]
23Ecklonia kurome (B)Phlorofucofuroeckol A, 8-8 bieckol, Dieckol, Eckol, PhloroglucinolHyaluronidase inhibition, Anti-wrinkle[180]
24Ecklonia stolonifera (B)Phlorotannins: Eckol, Phlorofucofuroeckol A, Dieckol, EckstolonolTyrosinase inhibitor, Skin whitener
Metalloproteinase inhibitors, Anti-wrinkle
[181]
25Ecklonia stolonifera (B)Phlorotannins: phlorofucofuroeckol AAnti-inflammatory[182]
26Ecklonia cava (B)PhlorotanninsUVB protector[183]
27Ecklonia cava (B)Phlorotannins, 6,6′-Bieckol, dioxinodehydroeckolMetalloproteinase inhibitors, Anti-wrinkle[183]
28Fucus vesiculosus, Ecklonia cava, Corallina pilulifera (R)Eckols, Fucols, Fuhalols, Phlorethols, FucolphloretholsAntiphotoaging, Antiaging, Antioxidants, UV protector,
Tyrosinase inhibition, Hyaluronidase inhibition
[184,185,186,187]
29Ishige foliacea (B)Octaphlorethol ATyrosinase inhibitor (whitening effect)[188]
30Ishige okamurae (B)DiphlorethohydroxycarmalolAntioxidant, UV protector[189]
31Sargassum horneri (B)Sargachromanol EAntiaging, Metalloproteinase inhibitors [189]
32Gracilaria gracilis (R)PhenolAntioxidant, ROS scavenger[190]
33Sargassum polycystum (B)Flavonoids, Tannins, Terpenoids, Phenols, SaponinsAnti-melanogenesis (skin whitener)[190,191]
34Laurencia sp. ®BromophenolsAntioxidant[192]
35Halidrys siliquosa, Ecklonia cava, Ascoseira mirabilis, Cystosphaera jacquinotii, Ishige okamurae, (B)Phlorotannins: diphlorethol, triphloroethol, trifuhalol and
tetrafuhalol, phloroglucinol, eckol, eckstolonol
Antioxidant, UV protector[193,194,195,196,197,198]
36Fucus vesiculosus (B)high polyphenol contentIncreased brightness and skin age spot reduction, UV protector, and soothing benefit[198]
37Sargassum polycystum, Ecklonia cava subsp. stolonifera (as Ecklonia stolonifera), Ecklonia cava, Sargassum siliquastrum (B)Unspecified flavonoids,
Tannins,
Phlorotannins
Tyrosinase inhibition, Anti melanogenesis[199,200,201]
38Eisenia bicyclis, Ecklonia cava subsp. kurome (as Ecklonia kurome), Ecklonia cava (B)Phlorofucofuroeckol-A, PhlorotanninsHyaluronidase inhibition Anti-inflammatory
Inhibit melanin synthesis, Antioxidant
[202]
39Ecklonia stolonifera (B)Phlorofucofuroeckol A and BAnti-inflammation, Antiaging (Metalloproteinase inhibitors)[203]
40Sargassum fusiforme
(as Hizikia fusiformis) (B)
FucosterolAntiaging, Metalloproteinase inhibitors[204]
41Ecklonia cava (B)Eckol, dieckolSkin whitener[204]
42Ishige foliacea (B)PhlorotanninDownregulation of tyrosinase and melanin
synthesis
[205,206]
43Laminaria ochroleuca (B)PolyphenolAntioxidant[207]
44Macrocystis pyrifera (B)PhlorotanninAntioxidant, ROS scavenger[208]
45Saccharina latissima (B)PhenolAntioxidant[209]
46Sargassum serratifolium (B)SargachromenolAnti-melanogenic[210]
47Schizymenia dubyi (R)PhenolAnti-melanogenic, tyrosinase inhibition[210]
48Sargassum thunbergia (B)ThunbergolAntioxidant[211]
49Pyropia columbina (R)PhenolAntioxidant[212]
50Rhodomela confervoides (R)BromophenolAntioxidant[213]
51Ulva prolifera (G)Phenol, flavonoidAntioxidant[214]
52Ulva rigida (G)PhenolAntioxidant[215]
53Ecklonia cava (B)DioxinodehydroeckolUVB protector[216]
54Eisenia bicyclis,
Ecklonia cava subsp. stolonifera (as E. stolonifera) (B)
EcokolAnti-inflammatory, Tyrosinase inhibition[217,218,219]
55Ecklonia cava subsp. stolonifera (as E. stolonifera) (B)Fucofuroeckol-AUVB protector[220]
56Cystoseira compressa (B)FuhalolAntioxidant[221]
57Fucus vesiculosus (B)FucophloroetholAntioxidant[222]
58Ecklonia cava (B)EckstolonolAntioxidant[223]
59Ishige foliacea (B)Octaphlorethol-AAntioxidant effects[224]
60Eisenia bicyclis,
Ecklonia cava subsp. stolonifera (as E. stolonifera) (B)
Phlorofucofuroeckol-AHepatoprotective against oxidative stress, Tyrosinase inhibition[225,226]
61Ecklonia cava (B)2-phloroeckol,
2-O-(2,4,6-Trihydroxyphenyl)-6,6′-bieckol
Tyrosinase inhibition[227]
62Ascophyllum nodosum, Fucus serratus, Himanthalia elongata, Halidrys siliquosa, (B)PhlorotanninsAntioxidant, Photoprotective[228,229,230]
63Ecklonia cava subsp. stolonifera (as E. stolonifera) (B)DioxinodehydroeckolDownregulation of melanogenic enzymes that are namely TYR, TRP1, and TRP2[231]
(B: Brown algae; G: Green algae; R: Red algae).
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kalasariya, H.S.; Pereira, L. Dermo-Cosmetic Benefits of Marine Macroalgae-Derived Phenolic Compounds. Appl. Sci. 2022, 12, 11954. https://doi.org/10.3390/app122311954

AMA Style

Kalasariya HS, Pereira L. Dermo-Cosmetic Benefits of Marine Macroalgae-Derived Phenolic Compounds. Applied Sciences. 2022; 12(23):11954. https://doi.org/10.3390/app122311954

Chicago/Turabian Style

Kalasariya, Haresh S., and Leonel Pereira. 2022. "Dermo-Cosmetic Benefits of Marine Macroalgae-Derived Phenolic Compounds" Applied Sciences 12, no. 23: 11954. https://doi.org/10.3390/app122311954

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