Next Article in Journal
Parametric Study of the Effect of Increased Magnetic Field Exposure on Microalgae Chlorella vulgaris Growth and Bioactive Compound Production
Previous Article in Journal
Morphological and Molecular Characters Differentiate Common Morphotypes of Atlantic Holopelagic Sargassum
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Cosmeceutical Significance of Seaweed: A Focus on Carbohydrates and Peptides in Skin Applications

by
Haresh S. Kalasariya
1,2,
Carlos Eliel Maya-Ramírez
3,
João Cotas
4 and
Leonel Pereira
4,*
1
Centre for Natural Products Discovery, School of Pharmacy and Biomolecular Sciences, Liverpool John Moores University, Byrom Street, Liverpool L3 3AF, UK
2
Faculty of Health and Life Sciences, INTI International University, Persiaran Perdana BBN, Putra Nilai, Nilai 71800, Negeri Sembilan, Malaysia
3
Laboratorio de Diagnóstico Molecular, Departamento de Bioquímica, Escuela Nacional de Ciencias Biológicas, Instituto Politécnico Nacional, Carpio y Plan de Ayala S/N, Colonia, Casco de Santo Tomás, Ciudad de México C.P. 11340, Mexico
4
Marine Resources, Conservation and Technology, Marine Algae Lab, CFE—Centre for Functional Ecology: Science for People & Planet, Department of Life Sciences, University of Coimbra, 3000-456 Coimbra, Portugal
*
Author to whom correspondence should be addressed.
Phycology 2024, 4(2), 276-313; https://doi.org/10.3390/phycology4020015
Submission received: 17 April 2024 / Revised: 20 May 2024 / Accepted: 21 May 2024 / Published: 27 May 2024

Abstract

:
The term ‘cosmeceutical’ refers to cosmetic products that offer medicinal or drug-like benefits. Marine algae are rich sources of bioactive compounds, particularly carbohydrates and peptides, which have gained attention for their potential in cosmeceuticals. These compounds are abundant, safe, and have minimal cytotoxicity effects. They offer various benefits to the skin, including addressing rashes, pigmentation, aging, and cancer. Additionally, they exhibit properties such as antimicrobial, skin-whitening, anti-aging, antioxidant, and anti-melanogenic effects. This review surveys the literature on the cosmeceutical potentials of algae-derived compounds, focusing on their roles in skin whitening, anti-aging, anticancer, antioxidant, anti-inflammatory, and antimicrobial applications. The discussion also includes current challenges and future opportunities for using algae for cosmeceutical purposes.

1. Introduction

Cosmeceutical ingredients are active compounds utilized to enhance the appearance and health of the human body, representing a hybrid category positioned between cosmetics and pharmaceuticals. Cosmeceutical formulations aim to improve skin health and beauty [1,2,3]. They combine cosmetic products with bioactive molecules, offering medicinal or drug-like applications to improve skin health and texture [4,5]. With the growing emphasis on skincare and modernization, cosmetic companies are expanding globally each year. To meet customer demands, these companies are increasingly relying on synthetic cosmetics and constituents. However, the ineffectiveness of synthetic components poses risks as they can accumulate in the skin, potentially causing toxic effects and harm to the skin’s structure. For instance, hydroxybenzoic acid esters (parabens), widely used in cosmetic formulations, have been linked to adverse effects on the skin and an increased incidence of malignant melanoma and breast cancer [6]. Globally, the cosmeceutical sector experiences annual growth driven by evolving beauty trends.
In response to consumer preferences, industries are increasingly relying on synthetic cosmetic ingredients such as hydroquinone (HQ), phthalates, para-aminobenzoic acid (PABA), benzophenones, butylated hydroxyanisole (BHA), butylated hydroxytoluene (BHT), and dibenzoylmethane (DBM) in formulations. The Scientific Committee on Consumer Safety warns that the excessive use of synthetic ingredients in cosmeceutical formulations may result in various toxicities, including acute toxicity, corrosion, irritation, dermal absorption, repeated dose toxicity, reproductive toxicity, mutagenicity/genotoxicity, carcinogenicity, and photoinduced toxicity on the skin and human health. For example, hydroxyanisole, commonly found in skin-whitening creams, has been associated with harmful effects such as ochronosis and potential mutagenicity [7,8,9]. Phthalates, another common substance in cosmetic formulations, have been linked to DNA mutations and damage in human male gametes [10,11]. Moreover, synthetic chemical compounds can have detrimental effects on animals, including reduced sperm counts, altered pregnancy outcomes, and congenital disabilities in male genitalia [12].
Consequently, consumers are shifting towards natural cosmetic products [13]. This growing demand for skincare formulations and the search for alternative natural ingredients have led to the production of various types of cosmeceutical skin products [14]. Throughout history, cosmetics have played a significant role in society, traditionally utilizing natural compounds extracted from milk, flowers, fruits, seeds, vegetables, herbs, marine algae, and minerals [15]. Some regions still utilize marine macroalgae as a sustainable natural remedy, incorporating it into skincare products [16]. Cosmetic products serve as both cleaning and beautifying agents, aiming to enhance the user’s aesthetics without causing harmful side effects. Many products, including creams, lotions, and ointments, contain bioactive molecules such as vitamins, minerals, and antioxidants that promote skin, nail, and hair health [17,18,19].
This review aims to analyze the potential of two major seaweed components as natural ingredients in cosmetics that substitute synthetic compounds to promote more eco-friendly solutions.

2. Natural Cosmetic Products

The cosmetic products can come in various forms, ranging from simple creams or lotions to edible options like pills or functional foods with cosmeceutical activity [20,21]. However, while they cannot claim a genuine therapeutic function, it’s crucial to differentiate them from cosmetic preparations that target skin disease prevention. For instance, sunscreen is considered a drug for preventing skin diseases caused by solar exposure [20,21]. To accommodate cosmetic products claiming biological action, the term “cosmeceutical” was coined and is now widely used across countless products [22,23]. As discussed earlier, the cosmetic industry has long relied on synthetic ingredients. However, various factors, including modern lifestyle changes and the shortcomings of synthetic cosmetics, such as low absorption rates or harmful side effects, have spurred a shift towards natural bioactive compounds [24]. Consider parabens (hydroxybenzoic acid esters), extensively used in synthetic cosmetic products, which mimic the female hormone estrogen. However, this mimicry could potentially increase the risk of developing malignant melanoma or breast cancer [25,26]. Another example is phthalate, a compound commonly used in plastic production and found in several cosmetic products like nail polish, which has been shown to cause DNA modification and damage in human sperm cells [27].
In essence, understanding skin growth and damage mechanisms, coupled with the successful utilization of natural products as solutions to the aforementioned issues, has underscored the benefits of adopting a cosmeceutical approach to cosmetics rather than purely aesthetic ones [28]. Cosmetics have held a central role in human society since ancient times, primarily for religious and ornamental purposes. In antiquity, these products were derived from natural compounds like milk, flowers, fruits, seeds, and vegetables, as well as minerals like clay and ash [29]. In certain regions, seaweed is utilized as an alternative remedy for skin-related ailments, making it an incredible natural raw material for cosmetics [29,30]. The bioactive compounds found in seaweed possess multiple activities, making them valuable as active ingredients in cosmetic formulations [31,32,33].

Synthetic Cautions: A Road to the Natural Ingredients

According to Fernández-Álvarez et al. [34] and Knowland et al. [35], benzophenones, DBM, and PABA have exhibited allergic phototoxicities, dermatitis, and skin irritations. Similarly, BHA and BHT, commonly found in moisturizers and lipsticks, have been associated with allergic reactions, irritation, and skin corrosivity. Parabens, another ingredient, have been linked to carcinogenic and neurotoxic effects, among other harmful health consequences. Despite their widespread use in cosmetic formulations, parabens pose a high risk of breast cancer and the development of malignant melanoma in women [36]. However, the ACDS Contact Allergen Management Program (CAMP) report suggests that while around 19% of products contain different types of parabens, mainly methylparaben, ethylparaben, propylparaben, and butylparaben, these components exhibit little allergenicity compared to other preservatives, with minimal adverse reactions and low toxicity, ensuring safety and cost-effectiveness [37].
Hafeez and Maibach [38] reported fewer sensitizing effects of parabens in commercial applications, with limited reports often attributed to their use on damaged skin. Polyethylene glycol (PEG) is identified as a genotoxic compound that can cause skin irritation, systemic toxicities, and skin damage. In skin cosmetics, PEGs serve as emollients, emulsifiers, and vehicles, aiding in softening and lubricating the skin, proper mixing of water-based and oil-based ingredients, and delivering ingredients deeper into the skin. Additionally, the Agency for Toxic Substances and Disease Registry (ATSDR) and the Centers for Disease Control and Prevention (CDC) highlighted the toxicity of dibutyl phthalate (DBP), which can lead to DNA damage in male reproductive cells [39].
Previous studies by Khan and Alam [40] and Warbanski [41] have reported harmful effects of cosmetic ingredients in animal studies, including male genitalia disabilities, altered pregnancy outcomes, and reduced sperm counts. Moreover, the EC 1223/2009 regulation advocates for the prohibition of testing finished cosmetic products on animals and their marketing. To address the toxicities associated with these formulations, consumers have increasingly favored natural skincare products in recent years. Consequently, industries are shifting towards natural bioactive ingredients sourced from various natural resources, which are eco-friendly and less toxic [42,43]. Additionally, studies by Kim et al. [44], Pereira [45], Maqsood, Benjakul, and Shahidi [46], Panzella and Napolitano [47], and de Jesus Raposo et al. [48] suggest that a wide range of natural resources, including terrestrial plants, fungi, marine algae, bacteria, and animals, can be utilized in skin cosmetic products.

3. Introduction to Seaweeds

Marine macroalgae, commonly referred to as seaweed, have garnered significant attention for their skincare benefits. These organisms, which are eukaryotic and photosynthetic, thrive in aquatic environments along coastlines and in seawater. Classified into three primary taxonomic groups—red algae, brown algae, and green algae—marine macroalgae belong to the Rhodophyta, Ochrophyta, and Chlorophyta phyla, respectively [49,50,51,52]. Seaweeds are macroscopic, multicellular organisms with the ability to perform photosynthesis, thanks to chlorophyll and other pigments. They are widespread in coastal regions, including intertidal and sub-tidal zones, as well as in brackish water [53]. Brown algae, red algae, and green algae are categorized based on their pigment composition, with brown algae belonging to the Chromista kingdom and green and red algae to the Plantae kingdom [54,55]. Seaweeds boast a rich diversity of bioactive compounds, surpassing those found in terrestrial organisms [56,57,58].
The cosmetic and cosmeceutical industries are increasingly turning to macroalgae-derived compounds due to their promising biological activities [59,60,61]. Thus, seaweed compounds, such as lipids, fatty acids, polysaccharides, vitamins, minerals, amino acids, phenolic compounds, proteins, and pigments, have attracted attention for their potential skincare benefits [62,63,64]. Utilization of seaweed biomass in skincare formulations depends on their specific constituents, such as polysaccharides, carbohydrates, proteins, peptides, amino acids, phenolic compounds, vitamins, minerals, fatty acids, and pigments [31,65]. These bioactive constituents offer a range of benefits for the skin, including anti-tumor, anti-allergic, antimicrobial, antioxidant, anti-inflammatory, antilipidemic, antiwrinkle, anti-aging, moisturizing, and photoprotective properties [66,67,68,69,70]. Driven by consumer demand for natural skincare solutions, the cosmetic industry is increasingly incorporating marine macroalgae into its products. With their diverse array of bioactive compounds and proven skincare benefits, seaweeds are becoming valued ingredients in the pursuit of healthier, more radiant skin.

Seaweeds: Interesting Metabolites

Seaweed-derived metabolites exhibit significant potential for application in cosmeceutical products owing to their diverse bioactive constituents and associated biological activities [71]. Polysaccharides, in particular, play pivotal roles in cosmetics, serving functions such as moisturization, emulsification, wound healing, and thickening [72]. In a study by Fernando et al. [73], fucoidan extracted from Chnoospora minima (Phaeophyceae) demonstrated anti-inflammatory properties. This compound inhibited nitric oxide production induced by lipopolysaccharides and reduced levels of inducible nitric oxide, cyclooxygenase-2, and prostaglandin E2 in RAW macrophages. Similarly, Ariede et al. [74], Wang et al. [75], and Teas and Irhimeh [76] reported diverse beneficial effects of polysaccharides derived from Fucus vesiculosus (Phaeophyceae), including anti-aging, anti-melanogenic, anti-cancer, and antioxidant activities. These effects were attributed to the stimulation of collagen production, inhibition of tyrosinase, reduction of melanoma growth, and prevention of oxidation, respectively. Moreover, Ghorbanzadeh et al. [77] demonstrated the anti-inflammatory activity of sulfated polysaccharides from Padina tetrastromatica (Phaeophyceae) through inhibition of COX-2 and iNOS in a rat model of Paw edema. Khan et al. [78] reported the anti-inflammatory effects of polyunsaturated fatty acids isolated from Undaria pinnatifida (Phaeophyceae) on mouse ear edema and erythema. Additionally, Vasconcelos et al. [79] and Santos et al. [80] explored the antioxidant potential of methanolic extracts from Osmundaria obtusilo and Palisada flagellifera (Rhodophyta) using various assays, including DPPH, ABTS, metal chelating, Folin ciocalteau, and β-carotene bleaching assays.
Sargachromanol E, a phenolic compound extracted from Sargassum horneri (Phaeophyceae), demonstrated anti-aging effects by inhibiting matrix metalloprotein expression in UVA-irradiated dermal fibroblasts [81]. Similarly, Lee et al. [82] investigated the anti-melanogenic activity of Sargachromanol E from Sargassum serratifolium (Phaeophyceae), observing the downregulation of microphthalmia-associated transcription factors in B16F10 melanoma cells. Yoon et al. [83] also reported anti-melanogenic activity, showing inhibition of tyrosinase and melanin expression in the ethanolic extract of the brown macroalgae Petalonia binghamiae.
Furthermore, Wang et al. [84] found collagenase inhibition and control of matrix metalloproteinase (MMP) expression in vitro with mycosporine amino acids (MAAs) derived from Porphyra sp. (Rhodophyta). Pyropia yezoensis (formerly Porphyra yezoensis) (Rhodophyta)-derived mycosporine amino acids exhibited antioxidant activity through ROS scavenging potential and MMP expression modulation in human skin fibroblasts [85]. Hartmann et al. [86] identified anti-aging properties through collagenase inhibition with MAAs derived from Palmaria palmata (Rhodophyta). Moreover, certain macroalgae species, such as Laurencia pacifica, Palisada rigida (formerly Laurencia rigida), Wilsonosiphonia howei (formerly Polysiphonia howei), Rhodomela confervoides, and Schizymenia dubyi (Rhodophyta), were identified as rich sources of phenolic compounds [87,88].

4. Seaweed-Derived Polysaccharides for Skin Benefits

Polysaccharides derived from marine macroalgae are renowned for their myriad biological benefits. Various research studies have highlighted the presence of polysaccharides such as ulvan, fucoidan, alginate, laminarin, carrageenan, sulfated polysaccharides, agar, and agarose in macroalgae, emphasizing their potential cosmeceutical advantages. Berthon et al. [19] underscored the diverse bioactivities of fucoidan, including antioxidant, antimicrobial, anti-inflammatory, anticancer, and antihyperlipidemic properties. Venkatesan et al. [89] proposed the utilization of polysaccharides as bioactive constituents in skincare formulations, where they play pivotal roles as moisturizers, emulsifiers, wound healers, and thickeners. Previous studies, such as those by Yu and Sun [90], have demonstrated the inhibitory effects of fucoidan obtained from brown algae like Fucus sp. and Sargassum sp., as well as brown alga Laminaria sp. (Phaeophyceae), on tyrosinase activity. Polysaccharides are regarded as the most significant and beneficial compounds present in macroalgae, prized for their skin beneficial activities. Seaweeds boast a plethora of polysaccharides, including chitin, fucoidans, agar, carrageenan, alginates, ulvans, terpenoids, and tocopherol. In skincare cosmeceuticals, marine algae have shown promising activities such as anti-melanogenesis, antioxidant, anti-skin-aging, anti-inflammation, anti-atopic dermatitis, anti-skin-cancer, and repair of UV-induced damage. However, their potential depends on their chemical and biochemical properties [89,91,92,93,94,95,96,97,98,99,100,101,102,103,104,105,106,107,108,109,110,111,112,113,114]. The chemical structures of seaweed-derived polysaccharides depicted below in Figure 1.
Numerous macroalgal species, including Kappaphycus alvarezii (formerly Eucheuma cottonii) (Rhodophyta), Sargassum polycystum (Phaeophyceae), Padina boryana (formerly Padina tenuis) (Phaeophyceae), Fucus vesiculosus (Phaeophyceae), and Porphyra umbilicalis (Rhodophyta), are rich sources of carbohydrates. Polysaccharides find wide-ranging applications in skincare products, including photoprotection, moisturization, wound healing, thickening, emulsification, and preservation. Presently, skincare product manufacturers focus on compounds capable of regulating tyrosinase inhibition, collagenase and elastase inhibition, reduction of matrix metalloproteinase (MMP) activity, reactive oxygen species (ROS) reduction, and antioxidant activity. Holtkamp et al. [115] highlighted the multifaceted benefits of fucoidan, particularly in skin protection, antioxidants, antiaging, antiviral, anti-inflammatory, antitumor, and anticoagulant properties, supported by epidemiological and experimental studies. Fujimura et al. [116] demonstrated a significant reduction in skin thickness and improved elasticity with the application of a gel formulation containing 1% fucus extract. Additionally, Kakita and Kamishima [117] explored the use of fucoidan in topical anti-inflammatory formulations for cosmetic after-sun damage, allergic condition soothing products, and post-surgical formulations. Polysaccharides have also been widely acknowledged for their antioxidant, antiviral, anticoagulant, and antitumor properties in commercial skincare products. Park and Choi [118] demonstrated enhanced inhibition of melanoma cells with low-molecular-weight fucoidan. Polysaccharides play a crucial role in inhibiting collagenase and elastase. Sulfated polysaccharides from the brown alga Sargassum fusiforme (formerly Hizikia fusiformis) (Phaeophyceae) exhibited potential inhibition of collagenase and elastase pathways in UVB-exposed HDF cells [119]. Fucoidan derived from Chnoospora minima and Sargassum polycystum showed dose-dependent inhibition of elastase and collagenase activities [120]. Additionally, Moon et al. [121] and Yu et al. [122] investigated the inhibitory effect of fucoidan on the MMP1 promoter and its ability to increase Type-1 procollagen synthesis. Polysaccharides possess excellent water-retention properties, making them effective humectants and moisturizers in cosmetic formulations. Research by Huang et al. [123] highlighted the superior hydration and moisturizing effects of polysaccharides from Saccharina japonica (formerly Laminaria japonica) (Phaeophyceae) compared to hyaluronic acid. Sulfated polysaccharides isolated from the green macroalgae Ulva linza (formerly Ulva fasciata) exhibited superior moisture absorption and retention over 96 h compared to glycerol [124]. Polysaccharides from Saccharina japonica, Pyropia haitanensis (formerly Porphyra haitanensis) (Rhodophyta), Codium fragile, Ulva linza (formerly Enteromorpha linza), and Bryopsis plumosa (Chlorophyta) also displayed significant moisture absorption and retention capabilities [125]. The moisture-holding capacity was influenced by factors such as sulfate content and molecular weight. An oligosaccharide zinc complex from Laminaria digitata (Phaeophyceae) exhibited anti-acne activity by reducing sebum production [126]. Sebaaly et al. [127] reported the bactericidal activity of sulfated galactan derived from Corallina sp. (Rhodophyta) against Enterococcus faecalis and Streptococcus epidermidis. Alginates have been utilized in cosmetics for face masks and body wash ingredients due to their beneficial effects on skin structure and function, particularly in solidifying and stabilizing emulsions at low pH levels [128,129]. Balboa et al. [74] proposed agar’s versatile use in creams, serving as an emulsifier, stabilizer, moisturizer, and component in various cosmetic products like lotions, deodorants, anti-aging treatments, exfoliants, and acne treatments. This thematic is also discussed largely in the literature [130,131,132,133,134]. Alginic acid, derived from various brown algal species such as Fucales and Laminariales, Ascophyllum sp., Durvillaea sp., Ecklonia sp., Laminaria sp., Macrocystis sp., Saccharina sp., Sargassum sp., and Turbinaria sp., has been highlighted for its potential applications by Hempel, Colepicolo, and Zambotti-Villela et al. [135], particularly in skin-protective or barrier creams for dermatitis treatment, as well as in beauty masks or facial packs.
Furthermore, Kappa-, Iota-, and Lambda-carrageenan, extracted from different carrageenophytes like Betaphycus gelatinus, Chondrus crispus, Eucheuma denticulatum, Gigartina sp., Kappaphycus alvarezii, Hypnea musciformis, Mastocarpus sp., and Mazzaella sp., from the Rhodophyta, find usage in cosmetology for various purposes such as lotions, sunscreens, medicines, deodorant sticks, sprays, and foams [136,137,138]. Porphyrin, a sulfated polysaccharide obtained from the aqueous extract of red algae Porphyra sp. and Bangia sp. [49,139], has demonstrated potential cosmeceutical applications including skin-whitening, antiulcer, analgesic, and anti-inflammatory properties. Various brown seaweed species like Laminaria sp., Saccharina sp., Ascophyllum sp., Fucus sp., Sargassum sp., and Undaria sp. are recognized for their laminaran properties, associated with antitumor, anti-inflammatory, antiviral, antioxidant, anticoagulant, and anti-cellulite properties [24,140,141]. Among these, sulfated polysaccharides have garnered significant attention in cosmeceutical applications, exhibiting properties such as UV protection, anti-inflammatory, anticoagulant, antithrombotic, tyrosinase inhibition, antitumoral, antibacterial, antidiabetic, and antioxidative effects [142,143,144]. Na et al. [145] elucidated the role of fucoidan in inhibiting matrix metalloproteinase induced by UVB radiation.
According to Senni et al. [146], fucoidan shows promise in preventing photoaging of the skin. Additionally, it acts as a tyrosinase inhibitor, reducing skin pigmentation when utilized in skin-whitening formulations [147,148]. Ulvan, as suggested by Carvalho and Pereira [149], is considered a desirable raw material for cosmeceuticals due to its beneficial moisturizing, protective, antitumor, and antioxidative properties, particularly in gel formulations [150,151]. Polysaccharides constitute approximately 60% of all active metabolites found in seaweed. Composed of various monosaccharides linked by glycosidic bonds, these compounds form long-chained carbohydrates with hydrophilic and water-soluble properties, possessing a regular structure [152]. Serving a structural role in seaweed cell walls and acting as an energetic reservoir, polysaccharides exhibit proven biological activities and can be utilized in cosmeceutical products as moisturizers and antioxidants [153]. Among polysaccharides, macroalgal hydrocolloids, known as phycocolloids, hold significant industrial value. These structural polysaccharides, found in seaweed, typically form colloidal solutions—an intermediate phase between a solution and a suspension. As such, polysaccharides find applications across various industries, particularly in cosmetics, serving as thickeners, gelling agents, and stabilizers for suspensions and emulsions [152]. Table 1 below tabulates seaweed-derived compounds and their skin-beneficial actions.

4.1. Agar

Agar, also known as agar-agar or agarose, is a powerful gelatinous hydrocolloid extracted from various species of red seaweed. It comprises a heterogeneous mixture of two carbohydrates, agarose and agaropectin, which are distributed within the cell wall and intercellular spaces, serving as structural carbohydrates [182]. Agarose, constituting 50–90% of agar, primarily imparts its gelling properties [183,184]. This polysaccharide, characterized by its high molecular weight, consists of multiple (1–3)-β-D-galactopyranosyl-(1–4)-3,6-anhydro-α-L-galactopyranose units. However, modifications to this molecule’s structure are observed due to factors such as seaweed species and variations in biotic and abiotic parameters [185]. Agaropectin, present in smaller quantities, is a heterogeneous mixture of α-1,3-linked D-galactose containing sulfate and pyruvate moieties. The gelling properties of agar depend on the degree of sulfation and the concentration of 3,6-anhydrogalactose [186,187]. Beyond its gelling properties, agar is utilized as an emulsifier and stabilizer in creams and regulates moisture content in cosmetic products, including hand lotions, liquid soap, deodorants, foundation, exfoliants, cleansers, shaving creams, and facial moisturizers/lotions, as well as in treatments for acne and aging [188]. Agar exhibits high-temperature tolerance (up to 250 °C) and retains its characteristics even at near-boiling temperatures, making it suitable for applications in jellied confections, where ingredients can be treated at high temperatures and subsequently cooled [189]. Additionally, there may be other bioactivities associated with the sulfated-galactan present in agar [190,191].
Agar applications in pharmaceuticals and industrial cosmetics include use as a thickener and as an ingredient for tablets or capsules for drug delivery [192,193]. Carrageenans, approved for food applications and generally recognized as safe (GRAS), are high-molecular-weight sulfated linear polysaccharides with a backbone of alternating 3-D-galactopyranose and 4-D-galactopyranose with anhydrogalactose residues [137,194,195,196,197,198]. Porphyran, a complex sulfated galactan found in Porphyra sp., exhibits therapeutic properties such as anti-allergic effects, tyrosinase inhibition, protection against ultraviolet B radiation, anti-inflammatory and antitumoral activity, and promotion of beneficial bacteria growth in the intestinal microbiota without toxicity in mouse models [199,200,201,202,203,204,205].

4.2. Alginic Acid

Alginic acid is the primary polysaccharide found in many brown seaweeds, including species such as Ascophyllum, Durvillaea, Ecklonia, Laminaria, Lessonia, Macrocystis, Saccharina, Sargassum, and Turbinaria. It is a linear copolymer of β-D-mannuronic acid and α-L-guluronic acid units linked by 1,4-glycosidic bonds. The extraction process involves converting an insoluble mixture of alginic acid salts into soluble salts, known as alginate or algin, which can be extracted in aqueous solutions [206,207]. In the presence of divalent cations such as Mg2+, Ca2+, Sr2+, and Ba2+, alginates readily form hydrogels. Extraction involves converting insoluble alginic acid and its salts into water-soluble potassium and sodium alginate. The stiffness or viscosity of these gels depends on the structural properties of the polymers (the amount of α-L-guluronate residues) and is influenced by the amount of external salt present [208,209]. Furthermore, under acidic conditions, alginic acid becomes insoluble, forming gels with its sodium and potassium salts, while still retaining water-soluble properties [210]. Alginates find widespread use as gelling agents, thickeners, protective colloids, or emulsion stabilizers in various cosmetic formulations, including hand jellies, lotions, ointment bases, pomades, hair preparations, greaseless creams, dentifrices, and other cosmetic products, owing to their chelating properties. This polysaccharide can also be employed in the formulation of skin protectants to prevent industrial dermatitis. Creams containing alginates form versatile films with enhanced skin adhesion, making them suitable for a variety of cosmetic applications [211].
Alginate exhibits various properties relevant to cosmetics and wellness products, particularly anti-allergic properties [212,213]. This action is observed in hydrogel formulations with alginate [214] and may contribute to preventing obesity [215,216,217].

4.3. Carrageenan

Certain species of red seaweed belonging to various families of the order Gigartinales, known as carrageenophytes, produce a pure polysaccharide known as “carrageenin”. However, carrageenin is unstable and difficult to extract; thus, it binds to one or more cations to form different carrageenan salts (carrageenans), which constitute approximately 30% to 75% of the seaweed’s dry weight. These linear sulfated polygalactans are composed of alternating residues of galactose linked by (1–3) and (1–4) bonds [218]. From a commercial perspective and according to several regulatory authorities (e.g., FDA, EFSA), carrageenan has been classified as safe based on minimal or no adverse physiological impacts observed in toxicological evaluations. The most important types of carrageenans are kappa (κ), iota (ι), and lambda (λ). Gelling properties are characteristic of iota and kappa carrageenan, while lambda carrageenan exhibits thickening properties [219,220,221]. Carrageenan can be extracted from various red seaweeds, such as Betaphycus gelatinus, Chondrus crispus, Eucheuma denticulatum, Sarcopeltis skottsbergii (formerly Gigartina skottsbergii), Kappaphycus alvarezii, Hypnea musciformis, Mastocarpus stellatus, Mazzaella laminarioides, and Sarcothalia crispata, which are valuable for several industries. More than 20% of carrageenan production is used in pharmacy and cosmetics. Many everyday cosmetic products contain carrageenan in their formulations, including toothpastes, hair wash products, lotions, medicines, sunscreens, shaving creams, deodorant sticks, sprays, and foams [222,223,224]. Research has demonstrated the diverse bioactive properties of carrageenan. Its ability to form hydrogels allows for applications in various areas due to its antiviral, antibacterial, and management effects on pathophysiological processes such as hyperlipidemia. These properties make carrageenan compelling for use in various applications, given its reported high safety, effectiveness, biocompatibility, biodegradability, and non-toxicity. However, further studies are necessary to fully assess the potential effectiveness of carrageenan [225,226].

4.4. Porphyran

Porphyran, a well-researched group of sulfated polysaccharides extracted aqueously, is derived from red seaweeds such as Porphyra/Pyropia and Bangia species. Structurally, porphyran is a linear polysaccharide composed of glycosidic linkages between repeated and alternated units of substituted -D-galactopyranose at carbon 3 and -L-galactopyranose units substituted at carbon 4, arranged in a disaccharide pattern represented as [!3)--D-galactopyranose-(1!4)--L-galactopyranose-(1!] n. Research indicates porphyran’s potential applications in skin whitening, anti-inflammatory, pain relief, and antiulcer activities, making it appealing for cosmeceutical use [227,228,229].

4.5. Laminaran

Laminaran, also known as laminarin or leucosin, is a glucan that can exist in either soluble or insoluble forms. The soluble form dissolves completely in cold water, while the insoluble form is soluble only in hot water [230]. Structurally, laminaran consists of _-(1,3)-linked D-glucose with intra-chain branching of _-(1,6). This polysaccharide is predominantly found in seaweeds belonging to the class Phaeophyceae (brown algae), including Laminaria and Saccharina, and in smaller amounts in Ascophyllum, Fucus, Sargassum, and Undaria. Laminarans have been studied for their anti-tumoral, anti-inflammatory, anticoagulant, antiviral, and antioxidant properties. In cosmetics, laminarins are commonly used in products targeting cellulite reduction [231,232,233,234].
Laminarin sulfate, known for its wound healing properties, has led to the development of novel hydrogel systems [235,236,237,238]. Additionally, promising outcomes have been demonstrated in numerous biomedical applications, including tissue engineering, cancer therapies, and antioxidant and anti-inflammatory properties [239]. Degradation by irradiation can enhance radical scavenging capacity and inhibitory activity against melanin synthesis in melanoma cells [240,241].

4.6. Fucoidan

Fucoidan, a hydrocolloid mainly composed of α-L-fucopyranose residues, unveils a heterogeneous and dynamic structure and composition. Besides α-L-fucose and sulfates, this molecule can also contain other monosaccharides, such as galactose, xylose, mannose, rhamnose, glucose, and/or uronic acids, or even acetylated groups. Moreover, there can also be differences among the molecular weight, branching, and substitutions according to the targeted species, as well as the selected extraction and purification methods [242]. These sulfated polysaccharides are interesting due to the wide range of bioactivities they present, such as the prevention of obesity, diabetes, and tumor development, as well as their anti-coagulant, anti-thrombotic, anti-inflammatory, UV blocker, tyrosinase inhibitor, antibacterial, antioxidative, and antihyperlipidemic activities [243,244,245,246]. Fucoidans’ anticoagulant bioactivity is highlighted by several authors, who claim that this polysaccharide extracted from the seaweeds Eklonia cava and Fucus vesiculosus has the ability to inhibit thrombin, which is mediated by anti-thrombin III plasma, mimicking the heparin activity [247,248,249]. In general, studies on the anticoagulant/anti-thrombin activity of fucoidans reported that they are directly dependent on the molecular weight, concentration, and/or placement of polysaccharide sulfate groups. Furthermore, binding and branching forms and their monomeric composition can also play an essential role in modulating the biological properties of these sulfated polysaccharides [250,251]. In fact, Moon et al. [252] found that, when the skin is exposed to UVB (ultraviolet B) radiation, fucoidan can promote the synthesis of procollagen type I and the inhibition of the expression of the metalloproteinase matrix. So, it is theorized that this polysaccharide can be employed as a therapeutic agent in order to prevent skin photoaging. In addition, studies have shown that fucoidans application in human leukocytes can decrease elastase activity, protecting the elastic fibers of the skin. Fucoidans also act as tyrosinase inhibitors and can, as well, minimize skin pigmentation [252,253]. Fucoidan can also protect the hair and skin by eliminating free radicals, reducing inflammation, wrinkles, allergies, and sensitive skin reactions. This polysaccharide can also promote skin elasticity, firmness, and brightness, as well as hair safety, growth, rigidity, cleanliness, and gloss [222]. The effects of fucoidan rely on a variety of cellular and molecular mechanisms, such as radical scavenging, down-regulation of COX-1/2, MAPK p38, inhibition of hyaluronidase, DPP-IV, and extension of APTT and TT. Literature shows that molecular weight, sulfate and fucose content, and polyphenols may have a role in these activities [254,255,256].
Fucoidans, heteropolysaccharides containing fucose and other monosaccharides like xylose, galactose, mannose, and glucuronic acid, along with sulfate, uronic acids, and acetyl groups, offer potential as cosmetic ingredients [257,258,259]. They are non-toxic, biodegradable, and biocompatible [260,261], with diverse biological properties [170,262,263,264,265,266,267], including antioxidant and antiradical effects [268,269,270,271]. These properties vary depending on molecular weight and sulfate content [272,273]. Fucoidans have demonstrated benefits in preventing and treating skin photoaging, inhibiting UVB-induced collagenase and gelatinase activities, and ex vivo inhibition of elastase activity in human skin [122,274,275,276]. They also inhibit wrinkle-related enzymes, enhance collagen synthesis in human dermal fibroblasts, and possess anti-inflammatory action against extracellular matrix degradation by matrix metalloproteinases [277,278,279].

4.7. Ulvan

Ulvan is a hydrocolloid extracted from green seaweeds, constituting approximately 8% to 29% of the algae’s dry weight. Its composition includes rhamnose, xylose, glucose, mannose, galactose, and uronic acids, forming two main repetitive disaccharides: ulvanobiuronic acids type A [(!4)-α-D-GlcA-(1!4)-α-L-Rha 3S-(1!)] and type B [(!4)-α-L-IdoA(1!4)-α-L-Rha 3S-(1!)]. However, variations in polysaccharide structure may occur due to taxonomic and/or eco-physiological differences [175,280]. Ulvan exhibits gelling properties in the presence of divalent cations like Ca2+, Cu2+, and Zn2+, within a pH range of 7.5 to 8.0, and it can withstand temperatures up to 180 °C. These rheological and bio-functional properties make ulvans appealing as raw materials for cosmeceuticals [281]. The gel formation mechanism of ulvans is complex, involving the creation of spherical structures in the presence of boric acid and calcium ions. Ulvans possess moisturizing, protective, antitumor, and antioxidative properties [282,283,284,285,286]. Ulvan finds applications as clouding and flavoring agents in beverages and as stabilizers in cosmetics, owing to its desirable characteristics [287,288]. Moreover, the antioxidant activity of ulvan is of interest to the cosmetic industry due to its demonstrated ability to protect against hydrogen peroxide-induced oxidative stress in vitro. Additionally, the presence of glucuronic acid, known for its moisturizing properties, and rhamnosyl residues, studied for their roles in cell proliferation and collagen synthesis, further enhance ulvan’s appeal as a raw material for the cosmetic industry [289,290].
Ulvans, complex and variable sulfated polysaccharides from ulvales, are primarily composed of rhamnose, xylose, glucose, glucuronic acid, iduronic acid, and sulfate [291,292]. They exhibit a range of activities, including gelling, anti-aging, anti-hyperlipidemic, and antiherpetic properties [293,294,295,296]. Additionally, seaweed and seaweed-derived compounds’ skin benefits/activities are noted in Table 2 below.

4.8. Remarks

Seaweeds possess a significant carbohydrate fraction forming their cell walls, with specific polysaccharides characteristic of each type of algae: brown algae contain alginate, laminaran, and fucoidan; green algae feature ulvan; and red algae are rich in agar and carrageenan. Polysaccharides are garnering increasing attention due to their bio-functional and physicochemical properties [327]. Sulfated polysaccharides, in particular, are highly regarded for their health benefits and diverse biological activities [170,328,329,330,331,332,333,334]. A critical aspect of these polysaccharides is the close relationship between their activity and composition, particularly their molecular weight. Depolymerization is often proposed to enhance activity [170], although other structural modifications are also feasible. Simple hydrophobization reactions, such as esterification, acylation, alkylation, amidation, or cross-linking reactions on native hydroxyl, amine, or carboxylic acid groups, can also enhance bioactivity [335,336].
Sulfated polysaccharides from green algae, such as rhamnans, arabinogalactans, galactans, and mannans, have variable compositions and structures, with properties highly influenced by molecular weight, particularly in terms of antiradical and chelating properties [337,338,339,340].
Agarooligosaccharides (AOS) and carrageenan-oligosaccharides (COS) demonstrate enhanced biological properties compared to native polysaccharides, particularly in terms of prebiotic, antitumoral, and antioxidant actions, attributed to their chemical structure, molecular weight, degree of polymerization, and glycosidic linkage flexibility [341]. The chemical structures of different red-algae-derived oligosaccharides are shown in Figure 2 below.

5. Seaweed-Derived Proteins, Peptides, and Amino Acids for Skin Benefits

Proteins, comprising one or more amino acid chains, are vital biological macromolecules found in all living organisms, playing essential roles in numerous cellular processes, including DNA replication, signal transduction, and molecular transport. Many proteins serve as enzymes that catalyze biochemical reactions crucial for various metabolic pathways [343]. Seaweed contains proteins in various forms, either as single or conjugate entities, often accompanied by protein derivatives and free amino acids like enzymes or peptides [344]. The bioactivity of proteins depends on their chemical structure and cellular localization, offering diverse cosmetic applications such as anti-tumor and anti-inflammatory properties, antioxidant effects, anti-aging benefits, and skin protection. Hence, seaweed proteins are utilized as moisturizers for hair and body and effectively serve as cosmeceuticals [345]. Amino acids, abundant in seaweed, play roles as hydrating agents in cosmetic products, many of them being integral components of the natural moisturizing factor (NMF) in human skin [346]. Seaweed is a rich source of both non-essential amino acids like alanine, serine, and proline, as well as essential amino acids such as histidine, tyrosine, and tryptophan [347]. For instance, Ulva australis exhibits antioxidant and antihypertensive effects due to its content of histidine and taurine [348]. Additionally, red seaweeds like Palmaria palmata (Dulse) and brown seaweeds like Himanthalia elongata (Sea Spaghetti) are noted for their abundance in serine, alanine, and glutamic acid [349]. Mycosporine-like amino acids (MAAs) are secondary metabolites found in seaweed, crucial for absorbing sunlight and protecting marine organisms from UV radiation [350]. MAAs are particularly abundant in red seaweeds such as Chondrus crispus, Palmaria palmata, Gelidium sp., Porphyra/Pyropia/Neopyropia sp., Gracilaria cornea, Asparagopsis armata, Solieria chordalis, Grateloupia lanceola, and Curdiea racovitzae. These compounds serve as effective UV protectors and stimulate cell proliferation in cosmetics and personal care products [351]. Different types of bioactive peptides, phycobiliproteins, and MAAs are shown in Figure 3 below.
Seaweeds serve as a rich protein source, with cultivation offering higher protein yields per unit area compared to terrestrial crops (2.5–7.5 tons/ha/year), although successful extraction is influenced by the presence of polysaccharides like alginates in brown seaweed or carrageenans in red seaweed [353]. Seasonal variations and habitat conditions affect the protein, peptide, and amino acid contents in seaweed; red algae generally exhibit higher contents (up to 47%) than green (between 9–26%), while brown algae have a lower concentration (3–15%) [354,355,356,357]. Proteins from all three macroalgae groups contain essential and non-essential amino acids, with bioactive peptides demonstrating numerous health benefits and high antioxidant properties, particularly those with low molecular weights, which are considered safer than synthetic molecules with reduced side effects [358,359,360,361,362,363,364,365]. Also, different dermatological benefits are tabulated in Table 3.

Peptides Bioactivities

Bioactive peptides typically comprise 3–20 amino acid residues, with their activities, including antioxidant and antimicrobial effects, influenced by both amino acid composition and sequence [379,380,381,382]. Examples of such peptides include carnosine, glutathione, and taurine, which exhibit antioxidant and chelating properties [383]. Taurine, although lacking a carboxyl group, possesses health-promoting properties and accumulates in various red algae species, such as Ahnfeltia plicata, Euthora cristata, and Ceramium virgatum [384]. PPY1, a peptide derived from Pyropia yezoensis through enzymatic hydrolysis, demonstrates anti-inflammatory effects by suppressing inflammatory cytokines [385]. Peptides like PYP1-5 and Porphyra 334, extracted from Porphyra yezoensis f. coreana, enhance elastin and collagen production while inhibiting the expression of matrix metalloproteinases (MMP) [386]. Ultrasound-assisted enzymatic hydrolysis has been proposed for extracting iodinated amino acids from red seaweeds like Palmaria palmata and Porphyra umbilicalis [387]. Moreover, seaweed-derived enzymes and peptide-related compounds with their bioactivity are represented in Table 4.
Mycosporine-like amino acids (MAAs) are secondary metabolites synthesized by marine organisms for protection against solar radiation [403,404,405]. Comprising cyclohexenone or cyclohexenimine chromophores with various amino acids, mainly glycine or iminoalcohol groups, as substituents, MAAs exhibit antioxidant and photoprotective properties [406,407,408,409,410,411,412]. Rhodophyceae species contain abundant MAAs such as shinorine, porphyra-334, palythine, asterina-330, mycosporine-glycine, palythinol, and palythene, with their contents varying based on geographic, seasonal, and bathymetric conditions, peaking during summer, and decreasing with water depth [413,414].
A multifunctional cosmetic liposome formulation incorporating UV filters, vitamins (A, C, and E), Ginkgo biloba extract (rich in quercetin), and Porphyra umbilicalis extract (abundant in proteins, vitamins, minerals, and MAAs such as porphyra-334 and shinorine) effectively combats signs of aging by improving hydration and reducing wrinkles and skin roughness [375,415]. Extracts like ASPAR’AGE™ from Asparagopsis armata and Aosaine® from Ulva lactuca, rich in amino acids, enhance skin elasticity, reduce wrinkles, and stimulate collagen production [416]. Similarly, extracts from Gelidium corneum, containing minerals, trace elements, and amino acids, improve skin softness and elasticity. MAAs offer diverse properties, serving as natural sunscreens, antioxidants, anti-inflammatories, and anti-aging agents, and stimulating skin renewal and cell proliferation, making them promising options for pharmaceutical and cosmetic applications [417].
Given the toxicity of synthetic dyes and the demand for natural colors in various industries, there is growing interest in phycobiliproteins like C-phycocyanin (used in food) and R-phycoerythrin (used in cosmetics). Phycobiliproteins are water-soluble compounds comprising proteins covalently bound to phycobilins, acting as reddish colorants [277,418,419,420,421,422]. B-phycoerythrin, resistant to pH changes, exhibits antioxidant properties and serves as a pink or purple dye in cosmetics [423,424]. Extracted from Gracilaria gracilis, phycobiliproteins show high antioxidant and radical-scavenging activities, particularly during winter harvests [425]. Extraction methods using aqueous solutions of ionic liquids, such as cholinium chloride, can yield up to 46.5% of R-phycoerythrin from fresh algal biomass [426]. Studies on Furcellaria lumbricalis and Coccotylus truncatus demonstrate an exponential correlation between R-phycoerythrin and allophycocyanin concentrations and collection depth [427]. Additionally, phycoerythrin and phycocyanin contents in dried Porphyra sp. extracts are slightly higher and lower, respectively, compared to Spirulina sp. [428].
Proteins, as macromolecules composed of amino acids, play a crucial role in skincare, acting as natural moisturizing factors that help prevent water loss from the skin [429]. Marine macroalgae are rich sources of various amino acids, including glycine, alanine, valine, leucine, proline, arginine, serine, histidine, tyrosine, and mycosporine amino acids (MAAs), offering numerous benefits for skin health and cosmetics [430]. Essential amino acids like histidine, tyrosine, and tryptophan, along with non-essential amino acids such as alanine, serine, and proline, are abundant in macroalgae, contributing to their antioxidant properties [431]. For instance, Ulva australis-derived amino acids like histidine and taurine exhibit antioxidant activity [432], while red macroalgae like Palmaria palmata and brown algae like Himanthalia elongata are rich in glutamic acid, alanine, and serine [433].
MAAs found in seaweeds play essential roles in protecting against UV radiation damage, scavenging radicals, and aiding DNA repair systems, making them valuable for UV protection and antioxidant activity in cosmetics [434]. Peptides derived from Pyropia yezoensis have been shown to increase elastin and collagen production and reduce the expression of matrix metalloprotein (MMP), contributing to anti-aging effects [374,435]. MAAs from Gracilaria genus species exhibit photoprotection activity [436,437], while Porphyra-334 obtained from Porphyra umbilicalis acts as an effective UV filter, reducing intracellular radical oxygen species and MMP expression [438,439]. MAAs also demonstrate antioxidant, anti-aging, and anti-inflammatory activities, scavenging reactive oxygen species, increasing UV-suppressed gene expression, and reducing inflammatory markers like COX-2 and involucrin expression [440]. Various seaweeds, including Dixoniella grisea, Ecklonia cava, and Undaria pinnatifida, have been reported for their antioxidant activity [441,442,443,444]. Peptides derived from Pyropia yezoensis stimulate collagen synthesis and elastin synthesis and suppress MMP-1 protein expression, contributing to anti-aging effects [445]. Additionally, proteins and amino acids from marine sources offer numerous skin benefits, including anti-inflammatory, antioxidant, antitumor, anti-aging, protective, and moisturizing effects in cosmetics [440,446,447]. Ulva australis is noted for its richness in essential amino acids like histidine and taurine [447], while red algae like Palmaria palmata and Himanthalia elongata are abundant in serine, alanine, and glutamic acid [448,449,450]. MAAs, such as those found in various red macroalgae, serve as UV protectors and cell proliferation activators in cosmetics [440,451,452]. MAAs like Porphyra-334 and Shinorine from Pyropia elongata (formerly Porphyra rosengurttii) demonstrate photostability and photoprotection against UV radiation, preventing sunburn cell formation and protecting against UV-induced skin thickening [453,454]. These compounds hold promise for preventing skin damage and promoting skin health in cosmetic applications.

6. Future Roads toward Seaweed-Based Cosmetics

Seaweeds, abundant and diverse marine organisms, have emerged as a promising reservoir of bioactive compounds with significant implications for skincare and cosmetic formulations. Within this intricate matrix of marine-derived ingredients, polysaccharides, proteins, peptides, and amino acids play pivotal roles, offering a diverse array of benefits for enhancing skin health and beauty. Polysaccharides extracted from seaweeds encompass a rich diversity of compounds, each possessing unique physicochemical characteristics that render them indispensable assets in the realm of cosmetic science. These polysaccharides, including agar, alginic acid, carrageenan, porphyran, laminaran, fucoidan, and ulvan, offer a broad spectrum of functionalities that extend far beyond their roles as mere thickeners, stabilizers, and emulsifiers. Delving deeper into their multifaceted nature reveals a treasure trove of bioactive properties, serving as the cornerstone for their integration into cosmetic formulations. Beyond their conventional uses, they demonstrate remarkable antioxidant capabilities, effectively combating oxidative stress and free radical damage, which are primary contributors to premature skin aging. Additionally, their anti-inflammatory effects provide soothing relief for irritated or inflamed skin, promoting a calmer and more balanced complexion. Moreover, these polysaccharides exhibit exceptional moisturizing properties, imparting hydration to the skin by attracting and retaining moisture, thus enhancing skin suppleness and resilience. Their innate ability to form protective barriers on the skin surface shields against environmental pollutants and UV radiation, contributing to overall skin health and vitality.
The versatility of these polysaccharides transcends traditional cosmetic boundaries, permeating a wide spectrum of skincare products. From opulent creams and serums that lavish the skin with indulgent nourishment to everyday essentials like cleansers and sunscreens that provide essential protection, their inclusion elevates formulations to new heights. Moreover, their sensorial qualities enhance the user experience, imbuing products with luxurious textures and pleasurable application rituals. As the cosmetic industry continues to innovate, these seaweed-derived polysaccharides stand poised as indispensable ingredients, offering a harmonious fusion of science and nature. Their multifunctional nature and diverse benefits underscore their relevance in meeting the evolving demands of consumers for effective, safe, and sustainable skincare solutions. In essence, the polysaccharides extracted from seaweeds represent a cornerstone of modern cosmetic science, embodying a symphony of functional and sensorial attributes that enrich formulations and elevate skincare experiences. Their boundless potential continues to inspire exploration and innovation, promising a bright future for seaweed-based cosmetics in the pursuit of radiant and healthy skin.
Proteins, peptides, and amino acids derived from seaweeds serve as fundamental pillars within marine-inspired skincare formulations, constituting a rich repertoire of bioactive compounds with multifaceted benefits for skin health and rejuvenation. Their multifunctional properties extend across various aspects of skincare, ranging from hydration and antioxidant protection to anti-inflammatory action, making them indispensable ingredients in the pursuit of radiant and youthful-looking skin. At the forefront of this marine skincare revolution are mycosporine-like amino acids (MAAs), abundantly present in seaweeds. These remarkable compounds serve as potent photoprotectors, effectively shielding the skin from the harmful effects of UV radiation. By absorbing and dissipating UV rays, MAAs mitigate the risk of sunburn, premature aging, and DNA damage, thereby bolstering the skin’s resilience against environmental stressors. Their innate ability to combat solar-induced skin damage underscores their significance in skincare formulations, offering a natural and sustainable approach to sun protection.
Additionally, proteins, peptides, and amino acids from seaweeds function as natural moisturizing factors, replenishing the skin’s hydration levels and maintaining its moisture barrier integrity. By attracting and retaining water molecules, these compounds help to prevent dehydration, dryness, and roughness, promoting a plump, smooth, and supple complexion. Moreover, their antioxidant properties neutralize free radicals, reducing oxidative stress and minimizing the appearance of fine lines, wrinkles, and other signs of aging. This dual action of hydration and antioxidant protection ensures optimal skin health and vitality. Furthermore, phycobiliproteins, serving as both natural colorants and antioxidants, enhance the allure of seaweed-based cosmetic formulations. Beyond their aesthetic appeal, these vibrant pigments offer additional skincare benefits, including antioxidant protection and brightening effects. By neutralizing harmful free radicals and promoting skin radiance, phycobiliproteins contribute to a luminous and healthy complexion, further enhancing the efficacy of marine-inspired skincare products.
In summary, proteins, peptides, amino acids, MAAs, and phycobiliproteins sourced from seaweeds represent a goldmine of bioactive compounds with unparalleled potential in skincare formulations. Their multifunctional properties, encompassing hydration, antioxidant protection, photoprotection, and aesthetic enhancement, make them invaluable assets in the quest for youthful and radiant skin. As the beauty industry continues to embrace natural and sustainable ingredients, seaweed-derived compounds stand at the forefront of innovation, offering safe, effective, and eco-friendly solutions for skincare enthusiasts worldwide. The integration of seaweed-derived ingredients into cosmetic formulations not only satisfies the growing demand for natural, sustainable skincare solutions but also embodies a comprehensive approach to skincare that addresses a wide array of concerns, spanning from hydration and elasticity to shielding against external aggressors. This holistic perspective underscores the multifaceted benefits of seaweed extracts, which offer not only cosmetic enhancement but also therapeutic properties that promote skin health and resilience. Furthermore, the renewable nature of seaweed resources highlights their potential to drive innovation within the cosmetic industry while simultaneously minimizing environmental impact. As consumer preferences increasingly shift towards eco-conscious beauty products, the utilization of seaweed-derived compounds aligns with this trend, offering a sustainable alternative to traditional skincare ingredients sourced from non-renewable or environmentally taxing sources.

Future Prospects towards Seaweed-Based Cosmetics for Safe Applications

Cosmetic raw ingredient supply is becoming increasingly important in terms of sustainability and environmental responsibility. Future advancements may include the production of targeted seaweed species in controlled cultivation in specific areas, which reduces environmental impact, enhances compound stability and purity rate, and ensures a reliable supply chain. Seaweed extracts might be used in formulations tailored to certain skin types and issues, resulting in a more focused and successful approach. Biotechnological procedures, including genetic manipulation and bioengineering, might be used to increase the production of beneficial substances in seaweed. This might lead to the creation of seaweed types with improved characteristics for cosmetic use.
Continued research into novel green extraction technologies will promote more effective extraction of specific compounds from cultivated seaweed. This increases the production, purity, and safety of targeted components, ensure their efficacy in cosmetic compositions and reduce the problems of contaminants or impurities. Formulators need to look for the chemical and biochemical interactions between seaweed extracts, compounds, and other natural substances renowned for their skincare advantages. Combining seaweed with plant extracts, vitamins, and antioxidants can result in potent, multifunctional cosmetic products.
Beyond topical uses, the cosmetic industry and RD teams can start to study the use of seaweed extracts in functional meals and nutricosmetics. Seaweed-based food products can help improve skin health and appearance by changing the human metabolism. However, for this to happen, there is a need to create and develop rigorous clinical trials to establish their efficacy and safety. In conclusion, scientific research will be vital in determining the advantages of seaweed for skin health in the future. Also, this green future trend needs to include more attempts to educate the public about the benefits of seaweed-based cosmetics and their problems. And regulatory organizations need to develop more specific criteria and requirements for the incorporation of seaweed extracts in skincare products because, today, there are few requirements to assure product safety and quality. Ultimately, the future of seaweed-based cosmetics is tending to be defined by a complex mix of scientific developments, industrial research, pharmaceutical safety approaches, environmental concerns, customer tastes, and a dedication to safe, responsible, and ethical sourcing.

7. Conclusions

This review demonstrated the potential of the exploitation of seaweed polysaccharides and proteins into natural cosmetics. Furthermore, continued exploration and research into the mechanisms of action, formulation optimization, and sustainable sourcing of seaweed-derived compounds are crucial for unlocking their full potential in skincare. By delving deeper into the bioactive properties of seaweeds and refining extraction techniques, scientists and skincare experts can further enhance the efficacy and versatility of these natural ingredients, paving the way for groundbreaking advancements in cosmetic science. Moreover, the synergistic combination of nature-inspired ingredients and cutting-edge technology holds immense promise for driving innovation in skincare. By harnessing the inherent bioactivity and adaptability of seaweeds, the cosmetic industry can develop novel formulations that deliver tangible results while also adhering to strict sustainability standards. This forward-thinking approach not only benefits consumers by providing them with efficacious skincare solutions but also contributes to the preservation of marine ecosystems and biodiversity. In essence, seaweeds, with their vast reservoir of bioactive compounds, represent a veritable treasure trove for the skincare industry. Their unique blend of natural goodness and scientific potential holds the key to revolutionizing skincare, offering a harmonious fusion of nature, science, and sustainability in the pursuit of radiant and resilient skin. As research continues to uncover the myriad benefits of seaweed-derived ingredients, the future of skincare looks brighter than ever, promising a healthier, more sustainable approach to beauty for generations to come.

Author Contributions

Writing—original draft preparation, H.S.K. and C.E.M.-R.; writing—review and editing, H.S.K., L.P. and J.C.; visualization, H.S.K.; supervision, L.P. and J.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflicts 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. Goyal, A.; Sharma, A.; Kaur, J.; Kumari, S.; Garg, M.; Sindhu, R.K.; Rahman, H.; Akhtar, M.F.; Tagde, P.; Najda, A.; et al. Bioactive-based cosmeceuticals: An update on emerging trends. Molecules 2022, 27, 828. [Google Scholar] [CrossRef] [PubMed]
  3. Draelos, Z.D. The cosmeceutical realm. Clin. Dermatol. 2008, 26, 627–632. [Google Scholar] [CrossRef] [PubMed]
  4. Kligman, D. Cosmeceuticals. Dermatol. Clin. 2000, 18, 609–615. [Google Scholar] [CrossRef] [PubMed]
  5. Dureja, H.; Kaushik, D.; Gupta, M.; Kumar, V.; Lather, V. Cosmeceuticals: An emerging concept. Indian J. Pharmacol. 2005, 37, 155. [Google Scholar] [CrossRef]
  6. Mogoşanu, G.D.; Grumezescu, A.M.; Bejenaru, C.; Bejenaru, L.E. Natural products used for food preservation. In Food Preservation; Academic Press: Cambridge, MA, USA, 2017; pp. 365–411. [Google Scholar]
  7. 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]
  8. Briganti, S.; Camera, E.; Picardo, M. Chemical and instrumental approaches to treat hyperpigmentation. Pigment. Cell Res. 2003, 16, 101–110. [Google Scholar] [CrossRef] [PubMed]
  9. 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] [PubMed]
  10. Zota, A.R.; Shamasunder, B. The environmental injustice of beauty: Framing chemical exposures from beauty products as a health disparities concern. Am. J. Obstet. Gynecol. 2017, 217, 418.e1–418.e6. [Google Scholar] [CrossRef] [PubMed]
  11. Zheng, L.X.; Liu, Y.; Tang, S.; Zhang, W.; Cheong, K.L. Preparation methods, biological activities, and potential applications of marine algae oligosaccharides: A review. Food Sci. Hum. Wellness 2023, 12, 359–370. [Google Scholar] [CrossRef]
  12. Pereira, J.X.; Pereira, T.C. Cosmetics and its health risks. Glob. J. Med. Res. 2018, 18, 63–70. [Google Scholar] [CrossRef]
  13. Ridder, M. Market Value for Natural and Organic BeautyWorldwide 2018–2027. Available online: https://www.statista.com/statistics/673641/global-market-value-for-natural-cosmetics/ (accessed on 7 February 2024).
  14. Mukherjee, P.K.; Maity, N.; Nema, N.K.; Sarkar, B.K. Bioactive compounds from natural resources against skin aging. Phytomedicine 2011, 19, 64–73. [Google Scholar] [CrossRef] [PubMed]
  15. de Jesus Raposo, M.F.; De Morais, A.M.B.; De Morais, R.M.S.C. Marine polysaccharides from algae with potential biomedical applications. Mar. Drugs 2015, 13, 2967–3028. [Google Scholar] [CrossRef] [PubMed]
  16. Lee, M.C.; Yeh, H.Y.; Shih, W.L. Extraction procedure, characteristics, and feasibility of Caulerpa microphysa (Chlorophyta) polysaccharide extract as a cosmetic ingredient. Marine Drugs. 2021, 19, 524. [Google Scholar] [CrossRef] [PubMed]
  17. Kim, S.K.; Pangestuti, R. 15 Biological Properties of Cosmeceuticals Derived from Marine Algae. In Marine Cosmeceuticals: Trends and Prospects; CRC Press: Boca Raton, FL, USA, 2011; p. 191. [Google Scholar]
  18. Polat, S.; Trif, M.; Rusu, A.; Šimat, V.; Čagalj, M.; Alak, G.; Meral, R.; Özogul, Y.; Polat, A.; Özogul, F. Recent advances in industrial applications of seaweeds. Crit. Rev. Food Sci. Nutr. 2023, 63, 4979–5008. [Google Scholar] [CrossRef] [PubMed]
  19. Resende, D.I.; Ferreira, M.; Magalhães, C.; Lobo, J.S.; Sousa, E.; Almeida, I.F. Trends in the use of marine ingredients in anti-aging cosmetics. Algal Res. 2021, 55, 102273. [Google Scholar] [CrossRef]
  20. Couteau, C.; Coiffard, L. Seaweed Application in Cosmetics. In Seaweed in Health and Disease Prevention; Fleurence, J., Levine, I., Eds.; Academic Press: Boston, MA, USA, 2016; pp. 423–441. [Google Scholar]
  21. Martins, A.; Vieira, H.; Gaspar, H.; Santos, S. Marketed marine natural products in the pharmaceutical and cosmeceutical industries: Tips for success. Mar. Drugs 2014, 12, 1066–1101. [Google Scholar] [CrossRef] [PubMed]
  22. Aranaz, I.; Acosta, N.; Civera-Tejuca, C.; Elorza, B.; Mingo, J.; Castro, C.; Civera-Tejuca, C.; Heras, A. Cosmetics and Cosmeceutical Applications of Chitin, Chitosan and Their Derivatives. Polymers 2018, 10, 213. [Google Scholar] [CrossRef] [PubMed]
  23. Imchen, T.; Singh, K.S. Marine algae colorants: Antioxidant, anti-diabetic properties and applications in food industry. Algal Research. 2023, 69, 102898. [Google Scholar] [CrossRef]
  24. Joshi, S.; Kumari, R.; Upasani, V.N. Applications of algae in cosmetics: An overview. Int. J. Innov. Res. Sci. Eng. Technol. 2018, 7, 1269. [Google Scholar]
  25. Fonseca, S.; Amaral, M.N.; Reis, C.P.; Custódio, L. Marine natural products as innovative cosmetic ingredients. Marine Drugs 2023, 21, 170. [Google Scholar] [CrossRef] [PubMed]
  26. Swetman, A.A.; Nicolaides, L.; Wareing, P.W.; New, J.H.; Wood, J.F.; Hammond, L. Food processing and preservation. Crop Post-Harvest. Sci. Technol. Princ. Pract. 2002, 1, 360–422. [Google Scholar]
  27. Bilal, M.; Iqbal, H.M. An insight into toxicity and human-health-related adverse consequences of cosmeceuticals—A review. Sci. Total Environ. 2019, 670, 555–568. [Google Scholar] [CrossRef] [PubMed]
  28. Wassie, T.; Niu, K.; Xie, C.; Wang, H.; Xin, W. Extraction techniques, biological activities and health benefits of marine algae Enteromorpha prolifera polysaccharide. Front. Nutr. 2021, 8, 747928. [Google Scholar] [CrossRef] [PubMed]
  29. Peñalver, R.; Lorenzo, J.M.; Ros, G.; Amarowicz, R.; Pateiro, M.; Nieto, G. Seaweeds as a functional ingredient for a healthy diet. Mar. Drugs 2020, 18, 301. [Google Scholar] [CrossRef] [PubMed]
  30. 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]
  31. Quitério, E.; Soares, C.; Ferraz, R.; Delerue-Matos, C.; Grosso, C. Marine health-promoting compounds: Recent trends for their characterization and human applications. Foods 2021, 10, 3100. [Google Scholar] [CrossRef] [PubMed]
  32. Bot, F.; Cossuta, D.; O’Mahony, J.A. Inter-relationships between composition, physicochemical properties and functionality of lecithin ingredients. Trends Food Sci. Technol. 2021, 111, 261–270. [Google Scholar] [CrossRef]
  33. Guo, Z.; Wei, Y.; Zhang, Y.; Xu, Y.; Zheng, L.; Zhu, B.; Yao, Z. Carrageenan oligosaccharides: A comprehensive review of preparation, isolation, purification, structure, biological activities and applications. Algal Res. 2022, 61, 102593. [Google Scholar] [CrossRef]
  34. 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]
  35. 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]
  36. 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]
  37. 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] [PubMed]
  38. Hafeez, F.; Maibach, H. An overview of parabens and allergic contact dermatitis. Skin Ther. Lett. 2013, 18, 5–7. [Google Scholar]
  39. Barrett, J. Chemical Exposures: The Ugly Side of Beauty Products. Environ. Health Perspect. 2005, 113, A24. [Google Scholar] [CrossRef]
  40. 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]
  41. Warbanski, M. The ugly side of the beauty industry. Herizons 2007, 21, 24–28. [Google Scholar]
  42. 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]
  43. Agatonovic-Kustrin, S.; Morton, D. Cosmeceuticals derived from bioactive substances found in marine algae. Oceanography 2013, 1, 106. [Google Scholar]
  44. 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]
  45. Pereira, L. Therapeutic and Nutritional Uses of Algae; CRC Press: Boca Raton, FL, USA, 2018; ISBN 9781498755382. [Google Scholar]
  46. 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] [PubMed]
  47. Panzella, L.; Napolitano, A. Natural phenol polymers: Recent advances in food and health applications. Antioxidants 2017, 6, 30. [Google Scholar] [CrossRef] [PubMed]
  48. de Jesus Raposo, M.F.; de Morais, R.M.S.C.; de Morais, A.M.M.B. Health applications of bioactive compounds from marine microalgae. Life Sci. 2013, 93, 479–486. [Google Scholar] [CrossRef] [PubMed]
  49. 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]
  50. 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]
  51. 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]
  52. 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]
  53. Fleurence, J.; Morançais, M.; Dumay, J.; Decottignies, P.; Turpin, V.; Munier, M.; Garcia-Bueno, N.; Jaouen, P. What are the prospects for using seaweed in human nutrition and for marine animals raised through aquaculture? Trends Food Sci. Technol. 2012, 27, 57–61. [Google Scholar] [CrossRef]
  54. García-Poza, S.; Leandro, A.; Cotas, C.; Cotas, J.; Marques, J.C.; Pereira, L.; Gonçalves, A.M. The evolution road of seaweed aquaculture: Cultivation technologies and the industry 4.0. Int. J. Environ. Res. Public Health 2020, 17, 6528. [Google Scholar] [CrossRef] [PubMed]
  55. Veluchamy, C.; Palaniswamy, R. A review on marine algae and its applications. Asian J. Pharm. Clin. Res. 2020, 13, 21–27. [Google Scholar] [CrossRef]
  56. Fu, W.; Nelson, D.R.; Yi, Z.; Xu, M.; Khraiwesh, B.; Jijakli, K.; Chaiboonchoe, A.; Alzahmi, A.; Al-Khairy, D.; Brynjolfsson, S.; et al. Bioactive compounds from microalgae: Current development and prospects. Stud. Nat. Prod. Chem. 2017, 54, 199–225. [Google Scholar]
  57. Couteau, C.; Coiffard, L. Phycocosmetics and other marine cosmetics, specific cosmetics formulated using marine resources. Mar. Drugs 2020, 18, 322. [Google Scholar] [CrossRef] [PubMed]
  58. Admassu, H.; Gasmalla, M.A.A.; Yang, R.; Zhao, W. Bioactive peptides derived from seaweed protein and their health benefits: Antihypertensive, antioxidant, and antidiabetic properties. J. Food Sci. 2018, 83, 6–16. [Google Scholar] [CrossRef] [PubMed]
  59. 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]
  60. 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]
  61. Venkatesan, J.; Kim, S.K. Osteoporosis treatment: Marine algal compounds. Adv. Food Nutr. Res. 2011, 64, 417–427. [Google Scholar] [PubMed]
  62. Miguel, S.P.; Ribeiro, M.P.; Otero, A.; Coutinho, P. Application of microalgae and microalgal bioactive compounds in skin regeneration. Algal Res. 2021, 58, 102395. [Google Scholar] [CrossRef]
  63. 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]
  64. 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]
  65. Freitas, R.; Martins, A.; Silva, J.; Alves, C.; Pinteus, S.; Alves, J.; Teodoro, F.; Ribeiro, H.M.; Gonçalves, L.; Petrovski, Ž.; et al. Highlighting the biological potential of the brown seaweed Fucus spiralis for skin applications. Antioxidants 2020, 9, 611. [Google Scholar] [CrossRef] [PubMed]
  66. Ning, L.; Yao, Z.; Zhu, B. Ulva (Enteromorpha) polysaccharides and oligosaccharides: A potential functional food source from green-tide-forming macroalgae. Mar. Drugs 2022, 20, 202. [Google Scholar] [CrossRef] [PubMed]
  67. Pallela, R.; Na-Young, Y.; Kim, S.K. Anti-photoaging and photoprotective compounds derived from marine organisms. Mar. Drugs 2010, 8, 1189–1202. [Google Scholar] [CrossRef] [PubMed]
  68. Fernando, I.S.; Kim, M.; Son, K.T.; Jeong, Y.; Jeon, Y.J. Antioxidant activity of marine algal polyphenolic compounds: Amechanistic approach. J. Med. Food 2016, 19, 615–628. [Google Scholar] [CrossRef] [PubMed]
  69. Indira, K.; Balakrishnan, S.; Srinivasan, M.; Bragadeeswaran, S.; Balasubramanian, T. Evaluation of in vitro antimicrobial property of seaweed (Halimeda tuna) from Tuticorin coast, Tamil Nadu, Southeast coast of India. Afr. J. Biotechnol. 2013, 12, 284–289. [Google Scholar]
  70. Liu, N.; Fu, X.; Duan, D.; Xu, J.; Gao, X.; Zhao, L. Evaluation of bioactivity of phenolic compounds from the brown seaweed of Sargassum fusiforme and development of their stable emulsion. J. Appl. Phycol. 2018, 30, 1955–1970. [Google Scholar] [CrossRef]
  71. Brunt, E.G.; Burgess, J.G. The promise of marine molecules as cosmetic active ingredients. Int. J. Cosmet. Sci. 2018, 40, 1–15. [Google Scholar] [CrossRef] [PubMed]
  72. Percival, E. The polysaccharides of green, red and brown seaweeds: Their basic structure, biosynthesis and function. Br. Phycol. J. 1979, 14, 103–117. [Google Scholar] [CrossRef]
  73. Fernando, I.S.; Sanjeewa, K.A.; Samarakoon, K.W.; Lee, W.W.; Kim, H.S.; Kang, N.; Ranasinghe, P.; Lee, H.S.; Jeon, Y.J. A fucoidan fraction purified from Chnoospora minima; a potential inhibitor of LPS-induced inflammatory responses. Int. J. Biol. Macromol. 2017, 104, 1185–1193. [Google Scholar] [CrossRef] [PubMed]
  74. Balboa, E.M.; Conde, E.; Soto, M.L.; Pérez-Armada, L.; Domínguez, H. Cosmetics from marine sources. In Springer Handbook of Marine Biotechnology; Springer: Berlin/Heidelberg, Germany, 2015; pp. 1015–1042. [Google Scholar]
  75. Wang, Z.J.; Xu, W.; Liang, J.W.; Wang, C.S.; Kang, Y. Effect of fucoidan on B16 murine melanoma cell melanin formation and apoptosis. Afr. J. Tradit. Complement. Altern. Med. 2017, 14, 149–155. [Google Scholar] [CrossRef] [PubMed]
  76. Teas, J.; Irhimeh, M.R. Melanoma and brown seaweed: An integrative hypothesis. J. Appl. Phycol. 2017, 29, 941–948. [Google Scholar] [CrossRef]
  77. Ghorbanzadeh, B.; Mansouri, M.T.; Hemmati, A.A.; Naghizadeh, B.; Mard, S.A.; Rezaie, A. Mechanism underlying the anti-inflammatory effect of sulphated polysaccharide from Padina tetrastromatica against carrageenan induced paw edema in rats. Indian J. Pharmacol. 2015, 47, 292–298. [Google Scholar] [CrossRef] [PubMed]
  78. Khan, M.N.; Yoon, S.J.; Choi, J.S.; Park, N.G.; Lee, H.H.; Cho, J.Y.; Hong, Y.K. Anti-edema effects of brown seaweed (Undaria pinnatifida) extract on phorbol 12-myristate 13-acetate-induced mouse ear inflammation. Am. J. Chin. Med. 2009, 37, 373–381. [Google Scholar] [CrossRef]
  79. Vasconcelos, J.B.; de Vasconcelos, E.R.; Urrea-Victoria, V.; Bezerra, P.S.; Reis, T.N.; Cocentino, A.L.; Navarro, D.M.; Chow, F.; Areces, A.J.; Fujii, M.T. Antioxidant activity of three seaweeds from tropical reefs of Brazil: Potential sources for bioprospecting. J. Appl. Phycol. 2019, 31, 835–846. [Google Scholar] [CrossRef]
  80. Santos, J.P.; Torres, P.B.; dos Santos, D.Y.; Motta, L.B.; Chow, F. Seasonal effects on antioxidant and anti-HIV activities of Brazilian seaweeds. J. Appl. Phycol. 2019, 31, 1333–1341. [Google Scholar] [CrossRef]
  81. Kim, J.A.; Ahn, B.N.; Kong, C.S.; Kim, S.K. The chromene sargachromanol E inhibits ultraviolet A-induced ageing of skin in human dermal fibroblasts. Br. J. Dermatol. 2013, 168, 968–976. [Google Scholar] [CrossRef]
  82. Lee, H.Y.; Jang, E.J.; Bae, S.Y.; Jeon, J.E.; Park, H.J.; Shin, J.; Lee, S.K. Anti-melanogenic activity of gagunin D, a highly oxygenated diterpenoid from the marine sponge Phorbas sp., via modulating tyrosinase expression and degradation. Mar. Drugs 2016, 14, 212. [Google Scholar] [CrossRef] [PubMed]
  83. Yoon, H.S.; Koh, W.B.; Oh, Y.S.; Kim, I.J. The Anti-melanogenic effects of Petalonia binghamiae extracts in α-melanocyte stimulating hormone-induced B16/F10 murine melanoma cells. J. Korean Soc. Appl. Biol. Chem. 2009, 52, 564–567. [Google Scholar] [CrossRef]
  84. Wang, H.M.; Li, X.C.; Lee, D.J.; Chang, J.S. Potential biomedical applications of marine algae. Bioresour. Technol. 2017, 244, 1407–1415. [Google Scholar] [CrossRef] [PubMed]
  85. Sakai, S.; Komura, Y.; Nishimura, Y.; Sugawara, T.; Hirata, T. Inhibition of mast cell degranulation by phycoerythrin and its pigment moiety phycoerythrobilin, prepared from Porphyra yezoensis. Food Sci. Technol. Res. 2011, 17, 171–177. [Google Scholar] [CrossRef]
  86. Hartmann, A.; Gostner, J.; Fuchs, J.E.; Chaita, E.; Aligiannis, N.; Skaltsounis, L.; Ganzera, M. Inhibition of collagenase by mycosporine-like amino acids from marine sources. Planta Med. 2015, 81, 813–820. [Google Scholar] [CrossRef] [PubMed]
  87. 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]
  88. 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]
  89. Venkatesan, J.; Anil, S.; Kim, S.K. Introduction to Seaweed Polysaccharides. In Seaweed Polysaccharides—Isolation, Biological and Biomedical Applications, 1st ed.; Venkatesan, J., Anil, S., Kim, S.-K., Eds.; Elsevier: Amsterdam, The Netherlands, 2017. [Google Scholar]
  90. Yu, P.; Sun, H. Purification of a fucoidan from kelp polysaccharide and its inhibitory kinetics for tyrosinase. Carbohydr. Polym. 2014, 99, 278–283. [Google Scholar] [CrossRef] [PubMed]
  91. Fitton, J.H.; Dell’Acqua, G.; Gardiner, V.A.; Karpiniec, S.S.; Stringer, D.N.; Davis, E. Topical benefits of two fucoidan-rich extracts from marine macroalgae. Cosmetics 2015, 2, 66–81. [Google Scholar] [CrossRef]
  92. Moreira, J.B.; Vaz, B.D.S.; Cardias, B.B.; Cruz, C.G.; Almeida, A.C.A.D.; Costa, J.A.V.; Morais, M.G.D. Microalgae polysaccharides: An alternative source for food production and sustainable agriculture. Polysaccharides 2022, 3, 441–457. [Google Scholar] [CrossRef]
  93. Ahmed, A.; Taha, R. Marine Phytochemical Compounds and Their Cosmeceutical Applications. In Marine Cosmeceuticals: Trends and Prospects; Kim, S., Ed.; CRC Press: Boca Raton, FL, USA, 2011; pp. 51–61. [Google Scholar]
  94. Godlewska, K.; Michalak, I.; Tuhy, Ł.; Chojnacka, K. Plant Growth Biostimulants Based on Different Methods of Seaweed Extraction with Water. BioMed. Res. Int. 2016, 2016, 5973760. [Google Scholar] [CrossRef] [PubMed]
  95. Paduch, R.; Kandefer-Szerszeń, M.; Trytek, M.; Fiedurek, J. Terpenes: Substances useful in human healthcare. Arch. Immunol. Ther. Exp. 2007, 55, 315. [Google Scholar] [CrossRef] [PubMed]
  96. 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]
  97. Murugan, K.; Iyer, V.V. Differential growth inhibition of cancer cell lines and antioxidant activity of extracts of red, brown, and green marine algae. Vitr. Cell. Dev. Biol.-Anim. 2013, 49, 324–334. [Google Scholar] [CrossRef] [PubMed]
  98. Griffiths, M.; Harrison, S.T.; Smit, M.; Maharajh, D. Major commercial products from micro-and macroalgae. In Algae biotechnology: Products and Processes; Springer: Berlin/Heidelberg, Germany, 2016; pp. 269–300. [Google Scholar]
  99. Buono, S.; Langellotti, A.L.; Martello, A.; Bimonte, M.; Tito, A.; Carola, A.; Apone, F.; Colucci, G.; Fogliano, V. Biological activities of dermatological interest by the water extract of the microalga Botryococcus braunii. Arch. Dermatol. Res. 2012, 304, 755–764. [Google Scholar] [CrossRef] [PubMed]
  100. Kang, H.; Lee, C.H.; Kim, J.R.; Kwon, J.Y.; Seo, S.G.; Han, J.G.; Kim, B.G.; Kim, J.E.; Lee, K.W. Chlorella vulgaris attenuates Dermatophagoides farinae-induced atopic dermatitis-like symptoms in NC/Nga mice. Int. J. Mol. Sci. 2015, 16, 21021–21034. [Google Scholar] [CrossRef] [PubMed]
  101. Pimentel, F.B.; Alves, R.C.; Rodrigues, F.; Oliveira, M.B.P.P. Macroalgae-Derived Ingredients for Cosmetic Industry—An Update. Cosmetics 2017, 5, 2. [Google Scholar] [CrossRef]
  102. Kim, S.; You, D.H.; Han, T.; Choi, E.-M. Modulation of viability and apoptosis of UVB-exposed human keratinocyte HaCaT cells by aqueous methanol extract of laver (Porphyra yezoensis). J. Photochem. Photobiol. B Biol. 2014, 141, 301–307. [Google Scholar] [CrossRef] [PubMed]
  103. Luthfiyana, N.; Hidayat, T.; Nurilmala, M.; Anwar, E. Utilization of seaweed porridge Sargassum sp. and Eucheuma cottonii as cosmetic in protecting the skin. IOP Conf. Ser. Earth Environ. Sci. 2019, 278, 012055. [Google Scholar]
  104. Heo, S.-J.; Ko, S.-C.; Cha, S.-H.; Kang, D.-H.; Park, H.-S.; Choi, Y.-U.; Kim, D.; Jung, W.-K.; Jeon, Y.-J. Effect of phlorotannins isolated from Ecklonia cava on melanogenesis and their protective effect against photo-oxidative stress induced by UV-B radiation. Toxicol. Vitr. 2009, 23, 1123–1130. [Google Scholar] [CrossRef]
  105. Quah, C.C.; Kim, K.H.; Lau, M.S.; Kim, W.R.; Cheah, S.H.; Gundamaraju, R. Pigmentation and Dermal Conservative Effects of the Astonishing Algae Sargassum polycystum and Padina tenuis on Guinea Pigs, Human Epidermal Melanocytes (HEM) and Chang Cells. Afr. J. Tradit. Complement. Altern. Med. 2014, 11, 77–83. [Google Scholar] [CrossRef] [PubMed]
  106. Maia Campos, P.M.; de Melo, M.O.; de Camargo, F.B., Jr. Effects of polysaccharide-based formulations on human skin. In Polysaccharides; Ramawat, K.G., Mérillon, J.-M., Eds.; Springer: Berlin/Heidelberg, Germany, 2014; pp. 1–8. [Google Scholar]
  107. Sotelo, C.G.; Blanco, M.; Ramos, P.; Vázquez, J.A.; Perez-Martin, R.I. Sustainable sources from aquatic organisms for cosmeceuticals ingredients. Cosmetics 2021, 8, 48. [Google Scholar] [CrossRef]
  108. Jutur, P.P.; Nesamma, A.A.; Shaikh, K.M. Algae-Derived Marine Oligosaccharides and Their Biological Applications. Front. Mar. Sci. 2016, 3, 83. [Google Scholar] [CrossRef]
  109. Li, J.; Chi, Z.; Yu, L.; Jiang, F.; Liu, C. Sulfated modification, characterization, and antioxidant and moisture absorption/retention activities of a soluble neutral polysaccharide from Enteromorpha prolifera. Int. J. Biol. Macromol. 2017, 105, 1544–1553. [Google Scholar] [CrossRef] [PubMed]
  110. Fujimura, T.; Tsukahara, K.; Moriwaki, S.; Kitahara, T.; Takema, Y. Effects of natural product extracts on contraction and mechanical properties of fibroblast populated collagen gel. Biol. Pharm. Bull. 2000, 23, 291–297. [Google Scholar] [CrossRef] [PubMed]
  111. Silchenko, A.S.; Rasin, A.B.; Kusaykin, M.I.; Malyarenko, O.S.; Shevchenko, N.M.; Zueva, A.O.; Kalinovsky, A.I.; Zvyagintseva, T.N.; Ermakova, S.P. Modification of native fucoidan from Fucus evanescens by recombinant fucoidanase from marine bacteria Formosa algae. Carbohydr. Polym. 2018, 193, 189–195. [Google Scholar] [CrossRef] [PubMed]
  112. Miyazaki, K.; Hanamizu, T.; Sone, T.; Chiba, K.; Kinoshita, T.; Yoshikawa, S. Topical application of Bifidobacterium-fermented soy milk extract containing genistein and daidzein improves rheological and physiological properties of skin. J. Cosmet. Sci. 2004, 55, 473–480. [Google Scholar] [CrossRef] [PubMed]
  113. Teixeira, M.M.; Hellewell, P.G. The effect of the selectin binding polysaccharide fucoidin on eosinophil recruitment in vivo. Br. J. Pharmacol. 1997, 120, 1059–1066. [Google Scholar] [CrossRef] [PubMed]
  114. Chen, Y.H.; Tu, C.J.; Wu, H.T. Growth-inhibitory effects of the red alga Gelidium amansii on cultured cells. Biol. Pharm. Bull. 2004, 27, 180–184. [Google Scholar] [CrossRef] [PubMed]
  115. Silchenko, A.S.; Kusaykin, M.I.; Kurilenko, V.V.; Zakharenko, A.M.; Isakov, V.V.; Zaporozhets, T.S.; Gazha, A.K.; Zvyagintseva, T.N. Hydrolysis of fucoidan by fucoidanase isolated from the marine bacterium, Formosa algae. Mar. Drugs 2013, 11, 2413–2430. [Google Scholar] [CrossRef]
  116. Fujimura, T.; Tsukahara, K.; Moriwaki, S.; Kitahara, T.; Sano, T.; Takema, Y. Treatment of human skin with an extract of Fucus vesiculosus changes its thickness and mechanical properties. J. Cosmet. Sci. 2002, 53, 1–9. [Google Scholar] [PubMed]
  117. Kakita, H.; Kamishima, H. Some properties of alginate gels derived from algal sodium alginate. In Nineteenth International Seaweed Symposium: Proceedings of the 19th International Seaweed Symposium, Held in Kobe, Japan, 26–31 March 2007; Springer: Dordrecht, The Netherlands, 2009; pp. 93–99. [Google Scholar]
  118. Park, E.J.; Choi, J.I. Melanogenesis inhibitory effect of low molecular weight fucoidan from Undaria pinnatifida. J. Appl. Phycol. 2017, 29, 2213–2217. [Google Scholar] [CrossRef]
  119. Wang, L.; Lee, W.; Oh, J.; Cui, Y.; Ryu, B.; Jeon, Y.J. Protective effect of sulfated polysaccharides from celluclast-assisted extract of Hizikia fusiforme against ultraviolet B-induced skin damage by regulating NF-κB, AP-1, and MAPKs signalling pathways in vitro in human dermal fibroblasts. Mar. Drugs 2018, 16, 239. [Google Scholar] [CrossRef]
  120. Lee, K.Y.; You, H.J.; Jeong, H.G.; Kang, J.S.; Kim, H.M.; Dal Rhee, S.; Jeon, Y.J. Polysaccharide isolated from Poria cocos sclerotium induces NF-κB/Rel activation and iNOS expression through the activation of p38 kinase in murine macrophages. Int. Immunopharmacol. 2004, 4, 1029–1038. [Google Scholar] [CrossRef] [PubMed]
  121. Staniforth, V.; Huang, W.C.; Aravindaram, K.; Yang, N.S. Ferulic acid, a phenolic phytochemical, inhibits UVB-induced matrix metalloproteinases in mouse skin via posttranslational mechanisms. J. Nutr. Biochem. 2012, 23, 443–451. [Google Scholar] [CrossRef] [PubMed]
  122. Yu, B.C.; Lee, D.-S.; Bae, S.M.; Jung, W.-K.; Chun, J.H.; Urm, S.H.; Lee, D.-Y.; Heo, S.-J.; Park, S.-G.; Seo, S.-K.; et al. The effect of cilostazol on the expression of matrix metalloproteinase-1 and type I procollagen in ultraviolet-irradiated human dermal fibroblasts. Life Sci. 2013, 92, 282–288. [Google Scholar] [CrossRef] [PubMed]
  123. Huang, T.H.; Wang, P.W.; Yang, S.C.; Chou, W.L.; Fang, J.Y. Cosmetic and therapeutic applications of fish oil’s fatty acids on the skin. Mar. Drugs 2018, 16, 256. [Google Scholar] [CrossRef] [PubMed]
  124. Shao, P.; Shao, J.; Han, L.; Lv, R.; Sun, P. Separation, preliminary characterization, and moisture-preserving activity of polysaccharides from Ulva fasciata. Int. J. Biol. Macromol. 2015, 72, 924–930. [Google Scholar] [CrossRef] [PubMed]
  125. Wang, J.; Jin, W.; Hou, Y.; Niu, X.; Zhang, H.; Zhang, Q. Chemical composition and moisture-absorption/retention ability of polysaccharides extracted from five algae. Int. J. Biol. Macromol. 2013, 57, 26–29. [Google Scholar] [CrossRef] [PubMed]
  126. Ruxton, C.H.; Jenkins, G. A novel topical ingredient derived from seaweed significantly reduces symptoms of Acne vulgaris: A general literature review. J. Cosmet. Sci. 2013, 64, 219–226. [Google Scholar] [PubMed]
  127. Sebaaly, C.; Kassem, S.; Grishina, E.; Kanaan, H.; Sweidan, A.; Chmit, M.S.; Kanaan, H.M. Anticoagulant and antibacterial activities of polysaccharides of red algae Corallina collected from Lebanese coast. J. Appl. Pharm. Sci. 2014, 4, 30. [Google Scholar]
  128. Podkorytova, A.V.; Vafina, L.H.; Kovaleva, E.A.; Mikhailov, V.I. Production of algal gels from the brown alga, Laminaria japonica Aresch., and their biotechnological applications. J. Appl. Phycol. 2007, 19, 827–830. [Google Scholar] [CrossRef]
  129. Paudel, P.; Wagle, A.; Seong, S.H.; Park, H.J.; Jung, H.A.; Choi, J.S. A New Tyrosinase Inhibitor from the Red Alga Symphyocladia latiuscula (Harvey) Yamada (Rhodomelaceae). Mar. Drugs 2019, 17, 295. [Google Scholar] [CrossRef] [PubMed]
  130. Synytsya, A.; Čopíková, J.; Kim, W.J.; Park, Y.I. Cell wall polysaccharides of marine algae. In Springer Handbook of Marine Biotechnology; Springer: Berlin/Heidelberg, Germany, 2015; pp. 543–590. [Google Scholar]
  131. Sreekumar, K.; Bindhu, B. Alginic acid: A potential biopolymer from brown algae. Mater. Int. 2020, 2, 433–438. [Google Scholar]
  132. Michalak, I.; Dmytryk, A.; Chojnacka, K. Algae cosmetics. Encycl. Mar. Biotechnol. 2020, 1, 65–85. [Google Scholar]
  133. Feng, D.; Aldrich, C. Adsorption of heavy metals by biomaterials derived from the marine alga Ecklonia maxima. Hydrometallurgy 2004, 73, 1–10. [Google Scholar] [CrossRef]
  134. Pereira, L.; Amado, A.M.; Critchley, A.T.; van de Velde, F.; Ribeiro-Claro, P.J.A. Identification of selected seaweed polysaccharides (Phycocolloids) by vibrational spectroscopy (FTIR-ATR and FT-Raman). Food Hydrocoll. 2009, 23, 1903–1909. [Google Scholar] [CrossRef]
  135. Hempel, M.D.S.S.; Colepicolo, P.; Zambotti-Villela, L. Macroalgae Biorefinery for the Cosmetic Industry: Basic Concept, Green Technology, and Safety Guidelines. Phycology 2023, 3, 211–241. [Google Scholar] [CrossRef]
  136. Kleinübing, S.J.; Gai, F.; Bertagnolli, C.; Silva, M.G.C.D. Extraction of alginate biopolymer present in marine alga Sargassum filipendula and bioadsorption of metallic ions. Mater. Res. 2013, 16, 481–488. [Google Scholar] [CrossRef]
  137. Barros, F.C.; da Silva, D.C.; Sombra, V.G.; Maciel, J.S.; Feitosa, J.P.; Freitas, A.L.; de Paula, R.C. Structural characterization of polysaccharide obtained from red seaweed Gracilaria caudata (J Agardh). Carbohydr. Polym. 2013, 92, 598–603. [Google Scholar] [CrossRef] [PubMed]
  138. Peris-Ortiz, M.; Álvarez-García, J. Health and Wellness Tourism: Emergency of New Market Segment; Springer: Madrid, Spain, 2014. [Google Scholar]
  139. Gutiérrez, G. Compositions of Padina Algae or Their Extracts, and Their Pharmaceutical, Food Compositions, or Use for the Culture of Molluscs or Arthropods. European Patent EP 0655250 Al, 31 May 1995. Available online: https://patents.google.com/patent/EP0655250A1/en (accessed on 1 October 2018).
  140. Morrice, L.M.; McLean, M.W.; Long, W.F.; Williamson, F.B. Porphyran primary structure. In Eleventh International Seaweed Symposium: Proceedings of the Eleventh International Seaweed Symposium, Held in Qingdao, People’s Republic of China, 19–25 June 1983; Springer: Dordrecht, The Netherlands, 1983; pp. 572–575. [Google Scholar]
  141. Lourenço-Lopes, C.; Fraga-Corral, M.; Jimenez-Lopez, C.; Pereira, A.G.; Garcia-Oliveira, P.; Carpena, M.; Prieto, M.A.; Simal- Gandara, J. Metabolites from Macroalgae and Its Applications in the Cosmetic Industry: A Circular Economy Approach. Resources 2020, 9, 101. [Google Scholar] [CrossRef]
  142. Kim, M.-S.; Oh, G.-H.; Kim, M.-J.; Hwang, J.-K. Fucosterol Inhibits Matrix Metalloproteinase Expression and Promotes Type-1 Procollagen Production in UVB-induced HaCaT Cells. Photochem. Photobiol. 2013, 89, 911–918. [Google Scholar] [CrossRef] [PubMed]
  143. Lorbeer, A.J.; Tham, R.; Zhang, W. Potential products from the highly diverse and endemic macroalgae of Southern Australia and pathways for their sustainable production. Environ. Boil. Fishes 2013, 25, 717–732. [Google Scholar] [CrossRef]
  144. Drozd, N.N.; Tolstenkov, A.S.; Makarov, V.A.; Kuznetsova, T.A.; Besednova, N.N.; Shevchenko, N.M.; Zvyagintseva, T.N. Pharmacodynamic parameters of anticoagulants based on sulfated polysaccharides from marine algae. Bull. Exp. Biol. Med. 2006, 142, 591–593. [Google Scholar] [CrossRef] [PubMed]
  145. Na, M.; Jang, J.; Njamen, D.; Mbafor, J.T.; Fomum, Z.T.; Kim, B.Y.; Oh, W.K.; Ahn, J.S. Protein tyrosine phosphatase-1B inhibitory activity of isoprenylated flavonoids isolated from Erythrina mildbraedii. J. Nat. Prod. 2006, 69, 1572–1576. [Google Scholar] [CrossRef] [PubMed]
  146. Blondin, C.; de Agostiniz, A. Biological activities of polysaccharides from marine algae. Drugs Fut. 1995, 20, 1237–1249. [Google Scholar]
  147. Kim, K.B.; Jeong, S.M.; Kim, M.J.; Ahn, D.H. Tyrosinase Inhibitory Effects of Sargachromanol G, Sargachromanol I and Mojabanchromanol b isolated from Myagropsis myagroides. Indian J. Pharm. Sci. 2020, 82, 170–173. [Google Scholar] [CrossRef]
  148. Mensah, E.O.; Kanwugu, O.N.; Panda, P.K.; Adadi, P. Marine fucoidans: Structural, extraction, biological activities and their applications in the food industry. In Food Hydrocolloids; Elsevier: Amsterdam, The Netherlands, 2023; p. 108784.
  149. Carvalho, L.G.; Pereira, L. Review of marine algae as source of bioactive metabolites. In Marine Algae—Biodiversity, Taxonomy, Environmental Assessment and Biotechnology, 1st ed.; Pereira, L., Neto, J.M., Eds.; CRC Press: Boca Raton, FL, USA, 2015; ISBN 9781466581678. [Google Scholar]
  150. Ukai, K.; Mizutani, Y.; Hisada, K.; Yokoyama, M.; Futaki, S.; Toya, H. Fuel Electrode Material, a Fuel Electrode, and a Solid Oxide Fuel Cell. U.S. Patent 20060110633 A1, 25 May 2006. [Google Scholar]
  151. Thanh, T.T.T.; Quach, T.M.T.; Nguyen, T.N.; Luong, D.V.; Bui, M.L.; Van Tran, T.T. Structure and cytotoxic activity of ulvan extracted from green seaweed Ulva lactuca. Int. J. Biol. Macromol. 2016, 93, 695–702. [Google Scholar] [CrossRef]
  152. Mourelle, M.L.; Gómez, C.P.; Legido, J.L. The potential use of marine microalgae and cyanobacteria in cosmetics and thalassotherapy. Cosmetics 2017, 4, 46. [Google Scholar]
  153. Bodin, J.; Adrien, A.; Bodet, P.E.; Dufour, D.; Baudouin, S.; Maugard, T.; Bridiau, N. Ulva intestinalis protein extracts promote in vitro collagen and hyaluronic acid production by human dermal fibroblasts. Molecules 2020, 25, 2091. [Google Scholar] [CrossRef] [PubMed]
  154. Jiang, N.; Li, B.; Wang, X.; Xu, X.; Liu, X.; Li, W.; Chang, X.; Li, H.; Qi, H. The antioxidant and antihyperlipidemic activities of phosphorylated polysaccharide from Ulva pertusa. Int. J. Biol. Macromol. 2020, 145, 1059–1065. [Google Scholar] [CrossRef] [PubMed]
  155. De Luca, M.; Pappalardo, I.; Limongi, A.R.; Viviano, E.; Radice, R.P.; Todisco, S.; Martelli, G.; Infantino, V.; Vassallo, A. Lipids from microalgae for cosmetic applications. Cosmetics 2021, 8, 52. [Google Scholar] [CrossRef]
  156. Ayoub, A.; Pereira, J.M.; Rioux, L.E.; Turgeon, S.L.; Beaulieu, M.; Moulin, V.J. Role of seaweed laminaran from Saccharina longicruris on matrix deposition during dermal tissue-engineered production. Int. J. Biol. Macromol. 2015, 75, 13–20. [Google Scholar] [CrossRef] [PubMed]
  157. Ozanne, H.; Toumi, H.; Roubinet, B.; Landemarre, L.; Lespessailles, E.; Daniellou, R.; Cesaro, A. Laminarin effects, a β-(1, 3)-glucan, on skin cell inflammation and oxidation. Cosmetics 2020, 7, 66. [Google Scholar] [CrossRef]
  158. Susano, P.; Silva, J.; Alves, C.; Martins, A.; Gaspar, H.; Pinteus, S.; Mouga, T.; Goettert, M.I.; Petrovski, Ž.; Branco, L.B.; et al. Unravelling the dermatological potential of the brown seaweed Carpomitra costata. Marine Drugs 2021, 19, 135. [Google Scholar] [CrossRef] [PubMed]
  159. El-Beltagi, H.S.; Mohamed, A.A.; Mohamed, H.I.; Ramadan, K.M.; Barqawi, A.A.; Mansour, A.T. Phytochemical and potential properties of seaweeds and their recent applications: A review. Mar. Drugs 2022, 20, 342. [Google Scholar] [CrossRef] [PubMed]
  160. Ganesan, A.R.; Tiwari, U.; Rajauria, G. Seaweed nutraceuticals and their therapeutic role in disease prevention. Food Sci. Hum. Wellness 2019, 8, 252–263. [Google Scholar] [CrossRef]
  161. Morya, V.K.; Kim, J.; Kim, E.K. Algal fucoidan: Structural and size-dependent bioactivities and their perspectives. Appl. Microbiol. Biotechnol. 2012, 93, 71–82. [Google Scholar] [CrossRef] [PubMed]
  162. Senni, K.; Pereira, J.; Gueniche, F.; Delbarre-Ladrat, C.; Sinquin, C.; Ratiskol, J.; Godeau, G.; Fischer, A.M.; Helley, D.; Colliec-Jouault, S. Marine polysaccharides: A source of bioactive molecules for cell therapy and tissue engineering. Mar. Drugs 2011, 9, 1664–1681. [Google Scholar] [CrossRef] [PubMed]
  163. Sanjeewa, K.A.; Kang, N.; Ahn, G.; Jee, Y.; Kim, Y.T.; Jeon, Y.J. Bioactive potentials of sulfated polysaccharides isolated from brown seaweed Sargassum spp. in related to human health applications: A review. Food Hydrocoll. 2018, 81, 200–208. [Google Scholar] [CrossRef]
  164. Nagaoka, M.; Shibata, H.; Kimura-Takagi, I.; Hashimoto, S.; Aiyama, R.; Ueyama, S.; Yokokura, T. Anti-ulcer effects and biological activities of polysaccharides from marine algae. Biofactors 2000, 12, 267. [Google Scholar] [CrossRef] [PubMed]
  165. Moon, H.J.; Lee, S.H.; Ku, M.J.; Yu, B.C.; Jeon, M.J.; Jeong, S.H.; Stonik, V.A.; Zvyagintseva, T.N.; Ermakova, S.P.; Lee, Y.H. Fucoidan inhibits UVB-induced MMP-1 promoter expression and down regulation of type I procollagen synthesis in human skin fibroblasts. Eur. J. Dermatol. 2009, 19, 129–134. [Google Scholar] [CrossRef] [PubMed]
  166. 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]
  167. Su, W.; Wang, L.; Fu, X.; Ni, L.; Duan, D.; Xu, J.; Gao, X. Protective effect of a fucose-rich fucoidan isolated from Saccharina japonica against ultraviolet B-induced photodamage in vitro in human keratinocytes and in vivo in zebrafish. Mar. Drugs 2020, 18, 316. [Google Scholar] [CrossRef] [PubMed]
  168. Wang, L.; Jayawardena, T.U.; Yang, H.W.; Lee, H.G.; Jeon, Y.J. The potential of sulfated polysaccharides isolated from the brown seaweed Ecklonia maxima in cosmetics: Antioxidant, anti-melanogenesis, and photoprotective activities. Antioxidants 2020, 9, 724. [Google Scholar] [CrossRef] [PubMed]
  169. Pereira, L. Seaweeds as source of bioactive substances and skin care therapy—Cosmeceuticals, algotheraphy, and thalassotherapy. Cosmetics 2018, 5, 68. [Google Scholar] [CrossRef]
  170. Fernando, P.S.; Kim, K.N.; Kim, D.; Jeon, Y.J. Algal polysaccharides: Potential bioactive substances for cosmeceutical applications. Crit. Rev. Biotechnol. 2018, 1–15. [Google Scholar]
  171. Xue, C.; Yu, G.; Hirata, T.; Terao, J.; Lin, H. Antioxidative activities of several marine polysaccharides evaluated in a phosphatidylcholine-liposomal suspension and organic solvents. Biosci. Biotechnol. Biochem. 1998, 62, 206–209. [Google Scholar] [CrossRef] [PubMed]
  172. Perde-Schrepler, M.; Chereches, G.; Brie, I.; Tatomir, C.; Postescu, I.D.; Soran, L.; Filip, A. Grape seed extract as photochemopreventive agent against UVB-induced skin cancer. J. Photochem. Photobiol. B Biol. 2013, 118, 16–21. [Google Scholar] [CrossRef] [PubMed]
  173. Sun, Z.; Mohamed, M.A.A.; Park, S.Y.; Yi, T.H. Fucosterol protects cobalt chloride induced inflammation by the inhibition of hypoxia-inducible factor through PI3K/Akt pathway. Int. Immunopharmacol. 2015, 29, 642–647. [Google Scholar] [CrossRef]
  174. Verdin, A.; Cazier, F.; Fitoussi, R.; Blanchet, N.; Vié, K.; Courcot, D.; Momas, I.; Seta, N.; Achard, S. An in vitro model to evaluate the impact of environmental fine particles (PM 0.3–2.5) on skin damage. Toxicol. Lett. 2019, 305, 94–102. [Google Scholar] [CrossRef] [PubMed]
  175. Hentati, F.; Tounsi, L.; Djomdi, D.; Pierre, G.; Delattre, C.; Ursu, A.V.; Fendri, I.; Abdelkafi, S.; Michaud, P. Bioactive polysaccharides from seaweeds. Molecules 2020, 25, 3152. [Google Scholar] [CrossRef] [PubMed]
  176. Rengasamy, K.R.; Mahomoodally, M.F.; Aumeeruddy, M.Z.; Zengin, G.; Xiao, J.; Kim, D.H. Bioactive compounds in seaweeds: An overview of their biological properties and safety. Food Chem. Toxicol. 2020, 135, 111013. [Google Scholar] [CrossRef] [PubMed]
  177. Sachan, N.K.; Pushkar, S.; Jha, A.; Bhattcharya, A.J.J.P.R. Sodium alginate: The wonder polymer for controlled drug delivery. J. Pharm. Res. 2009, 2, 1191–1199. [Google Scholar]
  178. Priyadarshani, I.; Rath, B. Commercial and industrial applications of micro algae–A review. J. Algal Biomass Util. 2012, 3, 89–100. [Google Scholar]
  179. Xu, S.Y.; Kan, J.; Hu, Z.; Liu, Y.; Du, H.; Pang, G.C.; Cheong, K.L. Quantification of neoagaro-oligosaccharide production through enzymatic hydrolysis and its anti-oxidant activities. Molecules 2018, 23, 1354. [Google Scholar] [CrossRef] [PubMed]
  180. Zhang, W.; Jin, W.; Duan, D.; Zhang, Q. Structural analysis and anti-complement activity of polysaccharides extracted from Grateloupia livida (Harv.) Yamada. J. Oceanol. Limnol. 2019, 37, 806–814. [Google Scholar] [CrossRef]
  181. Zhang, Y.H.; Song, X.N.; Lin, Y.; Xiao, Q.; Du, X.P.; Chen, Y.H.; Xiao, A.F. Antioxidant capacity and prebiotic effects of Gracilaria neoagaro oligosaccharides prepared by agarase hydrolysis. Int. J. Biol. Macromol. 2019, 137, 177–186. [Google Scholar] [CrossRef] [PubMed]
  182. Leandro, A.; Pereira, L.; Gonçalves, A.M. Diverse applications of marine macroalgae. Mar. Drugs 2019, 18, 17. [Google Scholar] [CrossRef] [PubMed]
  183. Ktari, L.; Chebil Ajjabi, L.; De Clerck, O.; Gómez Pinchetti, J.L.; Rebours, C. Seaweeds as a promising resource for blue economy development in Tunisia: Current state, opportunities, and challenges. J. Appl. Phycol. 2022, 34, 489–505. [Google Scholar] [CrossRef]
  184. Nussinovitch, A. Hydrocolloid Applications; Springer Science and Business Media LLC.: Berlin/Heidelberg, Germany, 1997; ISBN 978-1-4613-7933-1. [Google Scholar]
  185. Pereira, L.; Correia, F. Algas Marinhas da Costa Portuguesa-Ecologia, Biodiversidade e Utilizações; Nota de Rodapé Editores: Paris, France, 2015; ISBN 978-989-20-5754-5. [Google Scholar]
  186. Dolganyuk, V.; Belova, D.; Babich, O.; Prosekov, A.; Ivanova, S.; Katserov, D.; Patyukov, N.; Sukhikh, S. Microalgae: A promising source of valuable bioproducts. Biomolecules 2020, 10, 1153. [Google Scholar] [CrossRef] [PubMed]
  187. Probst, Y. A review of the nutrient composition of selected Rubus berries. Nutr. Food Sci. 2015, 45, 242–254. [Google Scholar] [CrossRef]
  188. Rasmussen, R.S.; Morrissey, M.T. Marine biotechnology for production of food ingredients. Adv. Food Nutr. Res. 2007, 52, 237–292. [Google Scholar] [PubMed]
  189. Ouyang, Q.-Q.; Hu, Z.; Li, S.-D.; Quan, W.-Y.; Wen, L.-L.; Yang, Z.-M.; Li, P. Thermal degradation of agar: Mechanism and toxicity of products. Food Chem. 2018, 264, 277–283. [Google Scholar] [CrossRef] [PubMed]
  190. Cosenza, V.A.; Navarro, D.A.; Ponce, N.M.; Stortz, C.A. Seaweed polysaccharides: Structure and applications. In Industrial Applications of Renewable Biomass Products: Past, Present and Future; Springer: Berlin/Heidelberg, Germany, 2017; pp. 75–116. [Google Scholar]
  191. Alp Arıcı, T.; Özcan, A.S.; Özcan, A. Biosorption characteristics of Cu (II) and Cd (II) ions by modified alginate. J. Polym. Environ. 2020, 28, 3221–3234. [Google Scholar] [CrossRef]
  192. Chen, X.; Fu, X.; Huang, L.; Xu, J.; Gao, X. Agar oligosaccharides: A review of preparation, structures, bioactivities and application. Carbohydr. Polym. 2021, 265, 118076. [Google Scholar] [CrossRef] [PubMed]
  193. Aziz, E.; Batool, R.; Khan, M.U.; Rauf, A.; Akhtar, W.; Heydary, M.; Rehman, S.; Shahzad, T.; Malik, A.; Mosavat, S.-H.; et al. An overview on red algae bioactive compounds and their pharmaceutical applications. J. Altern. Complement. Med. 2021, 17, 20190203. [Google Scholar] [CrossRef] [PubMed]
  194. Laurienzo, P. Marine polysaccharides in pharmaceutical applications: An overview. Mar. Drugs 2010, 8, 2435–2465. [Google Scholar] [CrossRef] [PubMed]
  195. Di, T.; Chen, G.; Sun, Y.; Ou, S.; Zeng, X.; Ye, H. Antioxidant and immunostimulating activities in vitro of sulfated polysaccharides isolated from Gracilaria rubra. J. Funct. Foods 2017, 28, 64–75. [Google Scholar] [CrossRef]
  196. Isaka, S.; Cho, K.; Nakazono, S.; Abu, R.; Ueno, M.; Kim, D.; Oda, T. Antioxidant and anti-inflammatory activities of porphyrin isolated from discolored nori (Porphyra yezoensis). Int. J. Biol. Macromol. 2015, 74, 68–75. [Google Scholar] [CrossRef] [PubMed]
  197. Beaumont, M.; Tran, R.; Vera, G.; Niedrist, D.; Rousset, A.; Pierre, R.; Shastri, V.P.; Forget, A. Hydrogel-Forming Algae Polysaccharides: From Seaweed to Biomedical Applications. Biomacromolecules 2021, 22, 1027–1052. [Google Scholar] [CrossRef] [PubMed]
  198. Ghanbarzadeh, M.; Golmoradizadeh, A.; Homaei, A. Carrageenans and carrageenases: Versatile polysaccharides and promising marine enzymes. Phytochem. Rev. 2018, 17, 535–571. [Google Scholar] [CrossRef]
  199. Kwon, M.; Nam, T. Porphyran induces apoptosis related signal pathway in AGS gastric cancer cell lines. Life Sci. 2006, 79, 1956–1962. [Google Scholar] [CrossRef] [PubMed]
  200. Jafari, A.; Farahani, M.; Sedighi, M.; Rabiee, N.; Savoji, H. Carrageenans for tissue engineering and regenerative medicine applications: A review. Carbohydr. Polym. 2022, 281, 119045. [Google Scholar] [CrossRef] [PubMed]
  201. Carvalho, C.M.B.; Brocksom, T.J.; de Oliveira, K.T. Tetrabenzoporphyrins: Synthetic developments and applications. Chem. Soc. Rev. 2013, 42, 3302–3317. [Google Scholar] [CrossRef] [PubMed]
  202. Araujo, I.W.F.; Vanderlei, E.S.O.; Rodrigues, J.A.G.; Coura, C.O.; Quindere, A.L.G.; Fontes, B.P.; Queiroz, I.N.L.; Jorge, R.J.B.; Bezerra, M.M.; Silva, A.A.R.; et al. Effects of a sulfated polysaccharide isolated from the red seaweed Solieria filiformis on models of nociception and inflammation. Carbohydr. Polym. 2011, 86, 1207–1215. [Google Scholar] [CrossRef]
  203. Zhang, Y.; Tian, R.; Wu, H.; Li, X.; Li, S.; Bian, L. Evaluation of acute and sub-chronic toxicity of Lithothamnion sp. in mice and rats. Toxicol. Rep. 2020, 7, 852–858. [Google Scholar] [CrossRef] [PubMed]
  204. Wang, Y.S.; Cho, J.G.; Hwang, E.S.; Yang, J.E.; Gao, W.; Fang, M.Z.; Zheng, S.D.; Yi, T.H. Enhancement of protective effects of radix scutellariae on UVB-induced photo damage in human HaCaT keratinocytes. Appl. Biochem. Biotechnol. 2018, 184, 1073–1093. [Google Scholar] [CrossRef] [PubMed]
  205. Xie, X.-T.; Zhang, X.; Liu, Y.; Chen, X.-Q.; Cheong, K.-L. Quantification of 3,6-anhydro-galactose in red seaweed polysaccharides and their potential skin-whitening activity. 3 Biotech 2020, 10, 189. [Google Scholar] [CrossRef] [PubMed]
  206. Gaspar, R.; Pereira, L.; Sousa-Pinto, I. The seaweed resources of Portugal. Bot. Mar. 2019, 62, 499–525. [Google Scholar] [CrossRef]
  207. Armisén, R. Agar and agarose biotechnological applications. In International Workshop on Gelidium: Proceedings of the International Workshop on Gelidium Held in Santander, Spain, 3–8 September 1991; Springer: Dordrecht, The Netherlands, 1991; pp. 157–166. [Google Scholar]
  208. Yalçın, S. The mechanism of heavy metal biosorption on green marine macroalga Enteromorpha linza. Clean–Soil Air Water. 2014, 42, 251–259. [Google Scholar] [CrossRef]
  209. Araki, C. Some recent studies on the polysaccharides of agarophytes. In Proceedings of the Fifth International Seaweed Symposium, Halifax, Nova Scotia, 25–28 August 1966; pp. 3–17. [Google Scholar]
  210. Li, S.S.; Song, Y.L.; Yang, H.R.; An, Q.D.; Xiao, Z.Y.; Zhai, S.R. Modifying alginate beads using polycarboxyl component for enhanced metal ions removal. Int. J. Biol. Macromol. 2020, 158, 493–501. [Google Scholar] [CrossRef] [PubMed]
  211. Ibañez, E.; Cifuentes, A. Benefits of using algae as natural sources of functional ingredients. J. Sci. Food Agric. 2013, 93, 703–709. [Google Scholar] [CrossRef] [PubMed]
  212. Yu, B.; Bi, D.; Yao, L.; Li, T.; Gu, L.; Xu, H.; Li, X.; Li, H.; Hu, Z.; Xu, X. The inhibitory activity of alginate against allergic reactions in an ovalbumin-induced mouse model. Food Funct. 2020, 11, 2704–2713. [Google Scholar] [CrossRef] [PubMed]
  213. Szekalska, M.; Sosnowska, K.; Tomczykowa, M.; Winnicka, K.; Kasacka, I.; Tomczyk, M. In vivo anti-inflammatory and antiallergic activities of cynaroside evaluated by using hydrogel formulations. Biomed. Pharmacother. 2020, 121, 109681. [Google Scholar] [CrossRef] [PubMed]
  214. Mišurcová, L. Chemical composition of seaweeds. In Handbook of Marine Macroalgae: Biotechnology and Applied Phycology; John Wiley & Sons: Hoboken, NJ, USA, 2011; pp. 171–192. [Google Scholar]
  215. Burtin, P. Nutritional value of seaweeds. Electron. J. Environ. Agric. Food Chem. 2003, 2, 498–503. [Google Scholar]
  216. Xing, M.; Cao, Q.; Wang, Y.; Xiao, H.; Zhao, J.; Zhang, Q.; Ji, A.; Song, S. Advances in Research on the Bioactivity of Alginate Oligosaccharides. Mar. Drugs 2020, 18, 144. [Google Scholar] [CrossRef] [PubMed]
  217. Li, C.; Tian, Y.; Pei, J.; Zhang, Y.; Hao, D.; Han, T.; Wang, X.; Song, S.; Huang, L.; Wang, Z. Sea cucumber chondroitin sulfate polysaccharides attenuate OVA-induced food allergy in BALB/c mice associated with gut microbiota metabolism and Treg cell differentiation. Food Funct. 2023, 14, 7375–7386. [Google Scholar] [CrossRef] [PubMed]
  218. Qureshi, D.; Nayak, S.K.; Maji, S.; Kim, D.; Banerjee, I.; Pal, K. Carrageenan: A wonder polymer from marine algae for potential drug delivery applications. Curr. Pharm. Des. 2019, 25, 1172–1186. [Google Scholar] [CrossRef] [PubMed]
  219. Pacheco-Quito, E.-M.; Ruiz-Caro, R.; Veiga, M.-D. Carrageenan: Drug Delivery Systems and Other Biomedical Applications. Mar. Drugs 2020, 18, 583. [Google Scholar] [CrossRef] [PubMed]
  220. Matsuhiro, B. Vibrational spectroscopy of seaweed galactans. In Fifteenth International Seaweed Symposium: Proceedings of the Fifteenth International Seaweed Symposium Held in Valdivia, Chile, in January 1995; Springer: Dordrecht, The Netherlands, 1995; pp. 481–489. [Google Scholar]
  221. Rinaudo, M. Seaweed Polysaccharides. In Comprehensive Glycoscience; Kamerling, J.P., Ed.; Elsevier: Oxford, UK, 2007; pp. 691–735. ISBN 978-0-444-51967-2. [Google Scholar]
  222. Usov, A.I.; Zelinsky, N.D. Chemical structures of algal polysaccharides. In Functional Ingredients from Algae for Foods and Nutraceuticals; Woodhead Publishing: Cambridge, UK, 2013; pp. 23–86. [Google Scholar]
  223. Charlier, R.H.; Chaineux, M.-C.P. The Healing Sea: A Sustainable Coastal Ocean Resource: Thalassotherapy. J. Coast. Res. 2009, 254, 838–856. [Google Scholar] [CrossRef]
  224. Fernando, I.S.; Sanjeewa, K.A.; Kim, S.-Y.; Lee, J.-S.; Jeon, Y.-J. Reduction of heavy metal (Pb2+) biosorption in zebrafish model using alginic acid purified from Ecklonia cava and two of its synthetic derivatives. Int. J. Biol. Macromol. 2018, 106, 330–337. [Google Scholar] [CrossRef] [PubMed]
  225. Liu, J.; Zhan, X.; Wan, J.; Wang, Y.; Wang, C. Review for carrageenan-based pharmaceutical biomaterials: Favourable physical features versus adverse biological effects. Carbohydr. Polym. 2015, 121, 27–36. [Google Scholar] [CrossRef] [PubMed]
  226. Villarroel, L.H.; Zanlungo, A.B. Structural studies on the porphyran from Porphyra columbina (Montagne). Carbohydr. Res. 1981, 88, 139–145. [Google Scholar] [CrossRef]
  227. Bhatia, S.; Sharma, A.; Sharma, K.; Kavale, M.; Chaugule, B.; Dhalwal, K.; Namdeo, A.; Mahadik, K. Novel algal polysaccharides from marine source: Porphyran. Pharmacogn. Rev. 2008, 2, 271. [Google Scholar]
  228. Bhatia, S.; Sharma, K.; Sharma, A.; Nagpal, K.; Bera, T. Anti-inflammatory, Analgesic and Antiulcer properties of Porphyra vietnamensis. Avicenna J. Phytomed. 2015, 5, 69. [Google Scholar] [PubMed]
  229. Mondal, S.; Pain, T.; Sahu, K.; Kar, S. Large-scale Green synthesis of porphyrins. ACS Omega 2021, 6, 22922–22936. [Google Scholar] [CrossRef]
  230. Usman, A.; Khalid, S.; Usman, A.; Hussain, Z.; Wang, Y. Algal polysaccharides, novel application, and outlook. In Algae Based Polymers, Blends, and Composites; Elsevier: Amsterdam, The Netherlands, 2017; pp. 115–153. [Google Scholar]
  231. Šimat, V.; Elabed, N.; Kulawik, P.; Ceylan, Z.; Jamroz, E.; Yazgan, H.; Čagalj, M.; Regenstein, J.M.; Özogul, F. Recent advances in marine-based nutraceuticals and their health benefits. Mar. Drugs 2020, 18, 627. [Google Scholar] [CrossRef] [PubMed]
  232. Gautam, S.; Mannan, M.A.U. The role of algae in nutraceutical and pharmaceutical production. In Bioactive Natural Products in Drug Discovery; Springer: Berlin/Heidelberg, Germany, 2020; pp. 665–685. [Google Scholar]
  233. Ughy, B.; Nagy, C.I.; Kós, P.B. Biomedical potential of cyanobacteria and algae. Acta Biol. Szeged. 2015, 59 (Suppl. S2), 203–224. [Google Scholar]
  234. Lomartire, S.; Cotas, J.; Pacheco, D.; Marques, J.C.; Pereira, L.; Gonçalves, A.M. Environmental impact on seaweed phenolic production and activity: An important step for compound exploitation. Mar. Drugs 2021, 19, 245. [Google Scholar] [CrossRef] [PubMed]
  235. Kadam, S.U.; Tiwari, B.K.; O’Donnell, C.P. Extraction, structure and biofunctional activities of laminarin from brown algae. Int. J. Food Sci. Technol. 2015, 50, 24–31. [Google Scholar] [CrossRef]
  236. Costa, A.M.S.; Rodrigues, J.M.M.; Pérez-Madrigal, M.M.; Dove, A.P.; Mano, J.F. Modular Functionalization of Laminarin to Create Value-Added Naturally Derived Macromolecules. J. Am. Chem. Soc. 2020, 142, 19689–19697. [Google Scholar] [CrossRef] [PubMed]
  237. Li, J.; Cai, C.; Yang, C.; Li, J.; Sun, T.; Yu, G. Recent advances in pharmaceutical potential of brown algal polysaccharides and their derivatives. Curr. Pharm. Des. 2019, 25, 1290–1311. [Google Scholar] [CrossRef] [PubMed]
  238. Tümen Erden, S.; Ekentok Atici, C.; Cömez, B.; Sezer, A.D. Preparation and in vitro characterization of laminarin based hydrogels. J. Res. Pharm. 2021, 25, 164–172. [Google Scholar]
  239. Zargarzadeh, M.; Amaral, A.J.R.; Custódio, C.A.; Mano, J.F. Biomedical applications of laminarin. Carbohydr. Polym. 2020, 232, 115774. [Google Scholar] [CrossRef] [PubMed]
  240. Murapa, P.; Dai, J.; Chung, M.; Mumper, R.J.; D’Orazio, J. Anthocyanin-rich fractions of blackberry extracts reduce UV-induced free radicals and oxidative damage in keratinocytes. Phytother. Res. 2012, 26, 106–112. [Google Scholar] [CrossRef] [PubMed]
  241. Cao, L.; Lee, S.G.; Lim, K.T.; Kim, H.R. Potential anti-aging substances derived from seaweeds. Marine Drugs 2020, 18, 564. [Google Scholar] [CrossRef] [PubMed]
  242. Chevolot, L.; Mulloy, B.; Ratiskol, J.; Foucault, A.; Colliec-Jouault, S. A disaccharide repeat unit is the major structure in fucoidans from two species of brown algae. Carbohydr. Res. 2001, 330, 529–535. [Google Scholar] [CrossRef] [PubMed]
  243. Saravana, P.S.; Cho, Y.-N.; Patil, M.P.; Cho, Y.-J.; Kim, G.-D.; Park, Y.B.; Woo, H.-C.; Chun, B.-S. Hydrothermal degradation of seaweed polysaccharide: Characterization and biological activities. Food Chem. 2018, 268, 179–187. [Google Scholar] [CrossRef] [PubMed]
  244. Wu, L.; Sun, J.; Su, X.; Yu, Q.; Yu, Q.; Zhang, P. A review about the development of fucoidan in antitumor activity: Progress and challenges. Carbohydr. Polym. 2016, 154, 96–111. [Google Scholar] [CrossRef] [PubMed]
  245. Deepika, C.; Ravishankar, G.A.; Rao, A.R. Potential products from macroalgae: An overview. In Sustainable Global Resources of Seaweeds Volume 1: Bioresources, Cultivation, Trade and Multifarious Applications; Springer: Berlin/Heidelberg, Germany, 2022; pp. 17–44. [Google Scholar]
  246. Kuznetsova, T.A.; Besednova, N.N.; Mamaev, A.; Momot, A.P.; Shevchenko, N.M.; Zvyagintseva, T.N. Anticoagulant Activity of Fucoidan from Brown Algae Fucus evanescens of the Okhotsk Sea. Bull. Exp. Biol. Med. 2003, 136, 471–473. [Google Scholar] [CrossRef]
  247. Obluchinsksya, E.D.; Makarova, M.N.; Pozharitskaya, O.N.; Shikov, A.N. Effects of Ultrasound Treatment on the Chemical Composition and Anticoagulant Properties of Dry Fucus Extract. Pharm. Chem. J. 2015, 49, 183–186. [Google Scholar] [CrossRef]
  248. Wijesekara, I.; Pangestuti, R.; Kim, S.K. Biological activities and potential health benefits of sulfated polysaccharides derived from marine algae. Carbohydr. Polym. 2011, 84, 14–21. [Google Scholar] [CrossRef]
  249. Holtkamp, A.D.; Kelly, S.; Ulber, R.; Lang, S. Fucoidans and fucoidanases-focus on techniques for molecular structure elucidation and modification of marine polysaccharides. Appl. Microbiol. Biotechnol. 2009, 82, 1–11. [Google Scholar] [CrossRef] [PubMed]
  250. Li, B.; Lu, F.; Wei, X.; Zhao, R. Fucoidan: Structure and Bioactivity. Molecules 2008, 13, 1671–1695. [Google Scholar] [CrossRef]
  251. Rupérez, P.; Gómez-Ordóñez, E.; Jiménez-Escrig, A. Biological activity of algal sulfated and nonsulfated polysaccharides. In Bioactive Compounds from Marine Foods: Plant and Animal Sources; John Wiley & Sons: New York, NY, USA, 2013; pp. 219–247. [Google Scholar]
  252. Moon, H.E.; Islam, 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]
  253. Thomas, N.V.; Kim, S.-K. Beneficial Effects of Marine Algal Compounds in Cosmeceuticals. Mar. Drugs 2013, 11, 146–164. [Google Scholar] [CrossRef] [PubMed]
  254. Elzoghby, A.O.; Freag, M.S.; Elkhodairy, K.A. Biopolymeric Nanoparticles for Targeted Drug Delivery to Brain Tumors; Kesharwani, P., Gupta, U., Eds.; Elsevier: Amsterdam, The Netherlands, 2018; pp. 169–190. ISBN 978-0-12-812218-1. [Google Scholar]
  255. Pozharitskaya, O.N.; Obluchinskaya, E.D.; Shikov, A.N. Mechanisms of Bioactivities of Fucoidan from the Brown Seaweed Fucus vesiculosus L. of the Barents Sea. Mar. Drugs 2020, 18, 275. [Google Scholar] [CrossRef] [PubMed]
  256. Shan, X.; Liu, X.; Hao, J.; Cai, C.; Fan, F.; Dun, Y.; Zhao, X.; Liu, X.; Li, C.; Yu, G. In vitro and in vivo hypoglycemic effects of brown algal fucoidans. Int. J. Biol. Macromol. 2016, 82, 249–255. [Google Scholar] [CrossRef] [PubMed]
  257. Kumar, C.S.; Ganesan, P.; Suresh, P.V.; Bhaskar, N. Seaweeds as a source of nutritionally beneficial compounds-a review. J. Food Sci. Technol. 2008, 45, 1. [Google Scholar]
  258. Narayanaswamy, R.; Jo, B.W.; Choi, S.K.; Ismail, I.S. Fucoidan: Versatile cosmetic ingredient. An overview. J. Appl. Cosmetol. 2013, 31, 131–138. [Google Scholar]
  259. Baweja, P.; Kumar, S.; Sahoo, D.; Levine, I. Biology of Seaweeds in Seaweed. In Health and Disease Prevention; Fleurence, J., Levine, I., Eds.; Nikki Levy Elsevier Inc.: Amsterdam, The Netherlands, 2016; pp. 41–106. [Google Scholar]
  260. Li, N.; Zhang, Q.; Song, J. Toxicological evaluation of fucoidan extracted from Laminaria japonica in Wistar rats. Food Chem. Toxicol. 2005, 43, 421–426. [Google Scholar] [CrossRef] [PubMed]
  261. Citkowska, A.; Szekalska, M.; Winnicka, K. Possibilities of fucoidan utilization in the development of pharmaceutical dosage forms. Mar. Drugs 2019, 17, 458. [Google Scholar] [CrossRef] [PubMed]
  262. 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] [PubMed]
  263. Thiyagarasaiyar, K.; Goh, B.H.; Jeon, Y.J.; Yow, Y.Y. Algae metabolites in cosmeceutical: An overview of current applications and challenges. Mar. Drugs 2020, 18, 323. [Google Scholar] [CrossRef] [PubMed]
  264. Fernando, I.P.S.; Jayawardena, T.U.; Kim, H.S.; Vaas, A.; De Silva, H.I.C.; Nanayakkara, C.M.; Abeytunga, D.T.U.; Lee, W.; Ahn, G.; Lee, D.S.; et al. A keratinocyte and integrated fibroblast culture model for studying particulate matter-induced skin lesions and therapeutic intervention of fucosterol. Life Sci. 2019, 233, 116714. [Google Scholar] [CrossRef]
  265. Manlusoc, J.K.T.; Hsieh, C.L.; Hsieh, C.Y.; Salac, E.S.N.; Lee, Y.T.; Tsai, P.W. Pharmacologic application potentials of sulfated polysaccharide from marine algae. Polymers 2019, 11, 1163. [Google Scholar] [CrossRef] [PubMed]
  266. Wijesinghe, W.; Jeon, Y. Biological activities and potential industrial applications of fucose rich sulfated polysaccharides and fucoidans isolated from brown seaweeds: A review. Carbohydr. Polym. 2012, 88, 13–20. [Google Scholar] [CrossRef]
  267. Kartik, A.; Akhil, D.; Lakshmi, D.; Panchamoorthy Gopinath, K.; Arun, J.; Sivaramakrishnan, R.; Pugazhendhi, A. A critical review on production of biopolymers from algae biomass and their applications. Bioresour. Technol. 2021, 329, 124868. [Google Scholar] [CrossRef] [PubMed]
  268. Xue, C.-H.; Fang, Y.; Lin, H.; Chen, L.; Li, Z.-J.; Deng, D.; Lu, C.-X. Chemical characters and antioxidative properties of sulphated polysaccharides from Laminaria japonica. J. Appl. Phycol. 2001, 13, 67–70. [Google Scholar] [CrossRef]
  269. Premarathna, A.D.; Ahmed, T.A.; Rjabovs, V.; Hammami, R.; Critchley, A.T.; Tuvikene, R.; Hincke, M.T. Immunomodulation by xylan and carrageenan-type polysaccharides from red seaweeds: Anti-inflammatory, wound healing, cytoprotective, and anticoagulant activities. Int. J. Biol. Macromol. 2024, 260, 129433. [Google Scholar] [CrossRef] [PubMed]
  270. 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]
  271. Li, Y.; Wang, X.; Jiang, Y.; Wang, J.; Hwang, H.; Yang, X.; Wang, P. Structure characterization of low molecular weight sulfate Ulva polysaccharide and the effect of its derivative on iron deficiency anemia. Int. J. Biol. Macromol. 2019, 126, 747–754. [Google Scholar] [CrossRef] [PubMed]
  272. Ustyuzhanina, N.E.; Bilan, M.I.; Ushakova, N.A.; Usov, A.I.; Kiselevskiy, M.V.; Nifantiev, N.E. Fucoidans: Pro-or antiangiogenic agents? Glycobiology 2014, 24, 1265–1274. [Google Scholar] [CrossRef] [PubMed]
  273. Ali Karami, M.; Sharif Makhmalzadeh, B.; Pooranian, M.; Rezai, A. Preparation and optimization of silibinin-loaded chitosan–fucoidan hydrogel: An in vivo evaluation of skin protection against UVB. Pharm. Dev. Technol. 2021, 26, 209–219. [Google Scholar] [CrossRef] [PubMed]
  274. Dai, Y.L.; Jiang, Y.F.; Lu, Y.A.; Kang, M.C.; Jeon, Y.J. Fucoidan from acid-processed Hizikia fusiforme attenuates oxidative damage and regulate apoptosis. Int. J. Biol. Macromol. 2020, 160, 390–397. [Google Scholar] [CrossRef] [PubMed]
  275. Lim, J.K. A review of the usability of fucoidan extracted from brown seaweed as a functional ingredient of cosmetics. Kor J Aesthet Cosmetol. 2014, 12, 447–452. [Google Scholar]
  276. Ji, D.; You, L.; Ren, Y.; Wen, L.; Zheng, G.; Li, C. Protective effect of polysaccharides from Sargassum fusiforme against UVB-induced oxidative stress in HaCaT human keratinocytes. J. Funct. Foods 2017, 36, 332–340. [Google Scholar] [CrossRef]
  277. Yarkent, Ç.; Gürlek, C.; Oncel, S.S. Potential of microalgal compounds in trending natural cosmetics: A review. Sustainable Chemistry and Pharmacy. 2020, 17, 100304. [Google Scholar] [CrossRef]
  278. Vo, T.; Ngo, D.; Kang, K.; Jung, W.; Kim, S. The beneficial properties of marine polysaccharides in alleviation of allergic responses. Mol. Nutr. Food Res. 2015, 59, 129–138. [Google Scholar] [CrossRef] [PubMed]
  279. Kim, J.H.; Lee, J.-E.; Kim, K.H.; Kang, N.J. Beneficial effects of marine algae-derived carbohydrates for skin health. Mar. Drugs 2018, 16, 459. [Google Scholar] [CrossRef] [PubMed]
  280. Ray, B.; Lahaye, M. Cell-wall polysaccharides from the marine green alga Ulva “rigida” (Ulvales, Chlorophyta). Chemical structure of ulvan. Carbohydr. Res. 1995, 274, 313–318. [Google Scholar] [CrossRef]
  281. Yaich, H.; Ben Amira, A.; Abbes, F.; Bouaziz, M.; Besbes, S.; Richel, A.; Blecker, C.; Attia, H.; Garna, H. Effect of extraction procedures on structural, thermal and antioxidant properties of ulvan from Ulva lactuca collected in Monastir coast. Int. J. Biol. Macromol. 2017, 105, 1430–1439. [Google Scholar] [CrossRef]
  282. Pereira, L. Biological and therapeutic properties of the seaweed polysaccharides. Int. Biol. Rev. 2018, 2, 1–50. [Google Scholar] [CrossRef]
  283. Lahaye, M.; Rochas, C. Chemical structure and physico-chemical properties of agar. Hydrobiology 1991, 221, 137–148. [Google Scholar] [CrossRef]
  284. Zanella, L.; Alam, M.A. Extracts and bioactives from microalgae (sensu stricto): Opportunities and challenges for a new generation of cosmetics. In Microalgae Biotechnology for Food, Health and High Value Products; Springer: Berlin/Heidelberg, Germany, 2020; pp. 295–349. [Google Scholar]
  285. Robic, A.; Rondeau-Mouro, C.; Sassi, J.-F.; Lerat, Y.; Lahaye, M. Structure and interactions of ulvan in the cell wall of the marine green algae Ulva rotundata (Ulvales, Chlorophyceae). Carbohydr. Polym. 2009, 77, 206–216. [Google Scholar] [CrossRef]
  286. Basti, D.; Bricknell, I.; Beane, D.; Bouchard, D. Recovery from a near-lethal exposure to ultraviolet-C radiation in a scleractinian coral. J. Invertebr. Pathol. 2009, 101, 43–48. [Google Scholar] [CrossRef] [PubMed]
  287. Morelli, A.; Massironi, A.; Puppi, D.; Creti, D.; Martinez, E.D.; Bonistalli, C.; Fabroni, C.; Morgenni, F.; Chiellini, F. Development of ulvan-based emulsions containing flavour and fragrances for food and cosmetic applications. Flavour Fragr. J. 2019, 34, 411–425. [Google Scholar] [CrossRef]
  288. Duraikkannu, S.L.; Sankaranarayanan, S.; Gajaria, T.K.; Li, G.; Kujawski, W.; Kujawa, J.; Navia, R. A Short Review on the Valorization of Green Seaweeds and Ulvan: FEEDSTOCK for Chemicals and Biomaterials. Biomolecules 2020, 10, 991. [Google Scholar]
  289. Bai, L.; Xu, D.; Zhou, Y.M.; Zhang, Y.B.; Zhang, H.; Chen, Y.B.; Cui, Y.L. Antioxidant activities of natural polysaccharides and their derivatives for biomedical and medicinal applications. Antioxidants 2022, 11, 2491. [Google Scholar] [CrossRef]
  290. Oucif, H.; Benaissa, M.; Ali Mehidi, S.; Prego, R.; Aubourg, S.P.; Abi-Ayad, S.M.E.A. Chemical composition and nutritional value of different seaweeds from the west Algerian coast. J. Aquat. Food Prod. Technol. 2020, 29, 90–104. [Google Scholar] [CrossRef]
  291. Lahaye, M.; Robic, A. Structure and function properties of Ulvan, a polysaccharide from green seaweeds. Biomacromolecules 2007, 8, 1765–1774. [Google Scholar] [CrossRef] [PubMed]
  292. Li, C.; Tang, T.; Du, Y.; Jiang, L.; Yao, Z.; Ning, L.; Zhu, B. Ulvan and Ulva oligosaccharides: A systematic review of structure, preparation, biological activities and applications. Bioresour. Bioprocess. 2023, 10, 66. [Google Scholar] [CrossRef] [PubMed]
  293. Kidgell, J.T.; Carnachan, S.M.; Magnusson, M.; Lawton, R.J.; Sims, I.M.; Hinkley, S.F.R.; de Nys, R.; Glasson, C.R.K. Are all ulvans equal? A comparative assessment of the chemical and gelling properties of ulvan from blade and filamentous Ulva. Carbohyd. Polym. 2021, 264, 118010. [Google Scholar] [CrossRef] [PubMed]
  294. Li, B.; Xu, H.; Wang, X.; Wang, Y.; Jiang, N.; Qi, H.; Liu, X. Antioxidant and antihyperlipidemic activities of high sulfate content purified polysaccharide from Ulva pertusa. Int. J. Biol. Macromol. 2020, 146, 756–762. [Google Scholar] [CrossRef] [PubMed]
  295. Jiao, G.L.; Yu, G.L.; Zhang, J.Z.; Ewart, H.S. Chemical structures and bioactivities of sulfated polysaccharides from marine algae. Mar. Drugs 2011, 9, 196–223. [Google Scholar] [CrossRef] [PubMed]
  296. Fournière, M.; Bedoux, G.; Lebonvallet, N.; Lescchiera, R.; Goff-Pain, C.L.; Bourgougnon, N.; Latire, T. Poly-and oligosaccharide ulva sp. Fractions from enzyme-assisted extraction modulate the metabolism of extracellular matrix in human skin fibroblasts: Potential in anti-aging dermo-cosmetic applications. Mar. Drugs 2021, 19, 156. [Google Scholar] [CrossRef] [PubMed]
  297. Rupérez, P.; Ahrazem, O.; Leal, J.A. Potential antioxidant capacity of sulfated polysaccharides from the edible marine brown seaweed Fucus vesiculosus. J. Agric. Food Chem. 2002, 50, 840–845. [Google Scholar] [CrossRef] [PubMed]
  298. Wang, J.; Zhang, Q.; Zhang, Z.; Li, Z. Antioxidant activity of sulfated polysaccharide fractions extracted from Laminaria japonica. Int. J. Biol. Macromol. 2008, 42, 127–132. [Google Scholar] [CrossRef]
  299. Wang, J.; Wang, F.; Zhang, Q.; Zhang, Z.; Shi, X.; Li, P. Synthesized different derivatives of low molecular fucoidan extracted from Laminaria japonica and their potential antioxidant activity in vitro. Int. J. Biol. Macromol. 2009, 44, 379–384. [Google Scholar] [CrossRef]
  300. Marudhupandi, T.; Kumar, T.T.; Senthil, S.L.; Devi, K.N. In vitro antioxidant properties of fucoidan fractions from Sargassum tenerrimum. Pak. J. Biol. Sci. 2014, 17, 402–407. [Google Scholar] [CrossRef] [PubMed]
  301. Moon, H.J.; Lee, S.R.; Shim, S.N.; Jeong, S.H.; Stonik, V.A.; Rasskazov, V.A.; Zvyagintseva, T.; Lee, Y.H. Fucoidan inhibits UVB-induced MMP-1 expression in human skin fibroblasts. Biol. Pharm. Bull. 2008, 31, 284–289. [Google Scholar] [CrossRef] [PubMed]
  302. Leirós, G.J.; Kusinsky, A.G.; Balañá, M.E.; Hagelin, K. Triolein reduces MMP-1 upregulation in dermal fibroblasts generated by ROS production in UVB-irradiated keratinocytes. J. Dermatol. Sci. 2017, 85, 124–130. [Google Scholar] [CrossRef] [PubMed]
  303. Yang, J.H. Topical application of fucoidan improves atopic dermatitis symptoms in NC/Nga mice. Phytother. Res. 2012, 26, 1898–1903. [Google Scholar] [CrossRef] [PubMed]
  304. Moon, H.J.; Park, K.S.; Ku, M.J.; Lee, M.S.; Jeong, S.H.; Imbs, T.I.; Zvyagintseva, T.N.; Ermakova, S.P.; Lee, Y.H. Effect of Costaria costata fucoidan on expression of matrix metalloproteinase-1 promoter, mRNA, and protein. J. Nat. Prod. 2009, 72, 1731–1734. [Google Scholar] [CrossRef] [PubMed]
  305. Song, Y.S.; Balcos, M.C.; Yun, H.Y.; Baek, K.J.; Kwon, N.S.; Kim, M.K.; Kim, D.S. ERK activation by fucoidan leads to inhibition of melanogenesis in Mel-Ab cells. Korean J. Physiol. Pharmacol. Off. J. Korean Physiol. Soc. Korean Soc. Pharmacol. 2015, 19, 29. [Google Scholar] [CrossRef] [PubMed]
  306. Maruyama, H.; Tamauchi, H.; Kawakami, F.; Yoshinaga, K.; Nakano, T. Suppressive effect of dietary fucoidan on proinflammatory immune response and MMP-1 expression in UVB-irradiated mouse skin. Planta Medica 2015, 81, 1370–1374. [Google Scholar] [PubMed]
  307. Rocha, H.A.; Franco, C.R.; Trindade, E.S.; Veiga, S.S.; Leite, E.L.; Nader, H.B.; Dietrich, C.P. Fucan inhibits Chinese hamster ovary cell (CHO) adhesion to fibronectin by binding to the extracellular matrix. Planta Medica 2005, 71, 628–633. [Google Scholar] [CrossRef] [PubMed]
  308. Senni, K.; Gueniche, F.; Foucault-Bertaud, A.; Igondjo-Tchen, S.; Fioretti, F.; Colliec-Jouault, S.; Durand, P.; Guezennec, J.; Godeau, G.; Letourneur, D. Fucoidan a sulfated polysaccharide from brown algae is a potent modulator of connective tissue proteolysis. Arch. Biochem. Biophys. 2006, 445, 56–64. [Google Scholar] [CrossRef] [PubMed]
  309. Tian, T.; Chang, H.; He, K.; Ni, Y.; Li, C.; Hou, M.; Chen, L.; Xu, Z.; Chen, B.; Ji, M. Fucoidan from seaweed Fucus vesiculosus inhibits 2, 4-dinitrochlorobenzene-induced atopic dermatitis. Int. Immunopharmacol. 2019, 75, 105823. [Google Scholar] [CrossRef]
  310. Oomizu, S.; Yanase, Y.; Suzuki, H.; Kameyoshi, Y.; Hide, M. Fucoidan prevents Cε germline transcription and NFκB p52 translocation for IgE production in B cells. Biochem. Biophys. Res. Commun. 2006, 350, 501–507. [Google Scholar] [CrossRef] [PubMed]
  311. Iwamoto, K.; Hiragun, T.; Takahagi, S.; Yanase, Y.; Morioke, S.; Mihara, S.; Kameyoshi, Y.; Hide, M. Fucoidan suppresses IgE production in peripheral blood mononuclear cells from patients with atopic dermatitis. Arch. Dermatol. Res. 2011, 303, 425–431. [Google Scholar] [CrossRef] [PubMed]
  312. Zvyagintseva, T.N.; Shevchenko, N.M.; Popivnich, I.B.; Isakov, V.V.; Scobun, A.S.; Sundukova, E.V.; Elyakova, L.A. A new procedure for the separation of water-soluble polysaccharides from brown seaweeds. Carbohydr. Res. 1999, 322, 32–39. [Google Scholar] [CrossRef]
  313. Lee, N.Y.; Ermakova, S.P.; Choi, H.K.; Kusaykin, M.I.; Shevchenko, N.M.; Zvyagintseva, T.N.; Choi, H.S. Fucoidan from Laminaria cichorioides inhibits AP-1 transactivation and cell transformation in the mouse epidermal JB6 cells. Mol. Carcinog. Publ. Coop. Univ. Tex. MD Anderson Cancer Cent. 2008, 47, 629–637. [Google Scholar] [CrossRef] [PubMed]
  314. Qi, H.; Zhang, Q.; Zhao, T.; Chen, R.; Zhang, H.; Niu, X.; Li, Z. Antioxidant activity of different sulfate content derivatives of polysaccharide extracted from Ulva pertusa (Chlorophyta) in vitro. Int. J. Biol. Macromol. 2005, 37, 195–199. [Google Scholar] [CrossRef] [PubMed]
  315. Qi, H.; Zhang, Q.; Zhao, T.; Hu, R.; Zhang, K.; Li, Z. In vitro antioxidant activity of acetylated and benzoylated derivatives of polysaccharide extracted from Ulva pertusa (Chlorophyta). Bioorganic Med. Chem. Lett. 2006, 16, 2441–2445. [Google Scholar] [CrossRef] [PubMed]
  316. Li, J.; Xie, L.; Qin, Y.; Liang, W.H.; Mo, M.Q.; Liu, S.L.; Liang, F.; Wang, Y.; Tan, W.; Liang, Y. Effect of laminarin polysaccharide on activity of matrix metalloproteinase in photoaging skin. Zhongguo Zhong Yao Za Zhi = Zhongguo Zhongyao Zazhi = China J. Chin. Mater. Medica 2013, 38, 2370–2373. [Google Scholar]
  317. Adrien, A.; Bonnet, A.; Dufour, D.; Baudouin, S.; Maugard, T.; Bridiau, N. Pilot production of ulvans from Ulva sp. and their effects on hyaluronan and collagen production in cultured dermal fibroblasts. Carbohydr. Polym. 2017, 157, 1306–1314. [Google Scholar] [CrossRef] [PubMed]
  318. Thevanayagam, H.; Mohamed, S.M.; Chu, W.L. Assessment of UVB-photoprotective and antioxidative activities of carrageenan in keratinocytes. J. Appl. Phycol. 2014, 26, 1813–1821. [Google Scholar] [CrossRef]
  319. Kuda, T.; Tsunekawa, M.; Goto, H.; Araki, Y. Antioxidant properties of four edible algae harvested in the Noto Peninsula, Japan. J. Food Compos. Anal. 2005, 18, 625–633. [Google Scholar] [CrossRef]
  320. Takematsu, H.; Seiji, M. Effect of macrophages on elimination of dermal melanin from the dermis. Arch. Dermatol. Res. 1984, 276, 96–98. [Google Scholar] [CrossRef]
  321. Zhang, Q.; Li, N.; Zhou, G.; Lu, X.; Xu, Z.; Li, Z. In vivo antioxidant activity of polysaccharide fraction from Porphyra haitanesis (Rhodephyta) in aging mice. Pharmacol. Res. 2003, 48, 151–155. [Google Scholar] [CrossRef]
  322. Zhang, Q.; Li, N.; Liu, X.; Zhao, Z.; Li, Z.; Xu, Z. The structure of a sulfated galactan from Porphyra haitanensis and its in vivo antioxidant activity. Carbohydr. Res. 2004, 339, 105–111. [Google Scholar] [CrossRef]
  323. Zhao, T.; Zhang, Q.; Qi, H.; Zhang, H.; Niu, X.; Xu, Z.; Li, Z. Degradation of porphyran from Porphyra haitanensis and the antioxidant activities of the degraded porphyrans with different molecular weight. Int. J. Biol. Macromol. 2006, 38, 45–50. [Google Scholar] [CrossRef] [PubMed]
  324. Jiang, Z.; Hama, Y.; Yamaguchi, K.; Oda, T. Inhibitory effect of sulphated polysaccharide porphyran on nitric oxide production in lipopolysaccharide-stimulated RAW264. 7 macrophages. J. Biochem. 2012, 151, 65–74. [Google Scholar] [CrossRef]
  325. Liu, X.; Yuan, W.; Sharma-Shivappa, R.; van Zanten, J. Antioxidant activity of phlorotannins from brown algae. International J. Agric. Biol. Eng. 2017, 10, 184–191. [Google Scholar] [CrossRef]
  326. Zhang, Z.; Zhang, Q.; Wang, J.; Zhang, H.; Niu, X.; Li, P. Preparation of the different derivatives of the low-molecular-weight porphyran from Porphyra haitanensis and their antioxidant activities in vitro. Int. J. Biol. Macromol. 2009, 45, 22–26. [Google Scholar] [CrossRef] [PubMed]
  327. Shanura Fernando, I.P.; Asanka Sanjeewa, K.K.; Samarakoon, K.W.; Kim, H.S.; Gunasekara, U.K.D.S.S.; Park, Y.J.; Abeytungaa, 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]
  328. Han, J.; Liu, B.; Liu, Q.-M.; Zhang, Y.-F.; Liu, Y.-X.; Liu, H.; Cao, M.-J.; Liu, G.-M. Red algae sulfated polysaccharides effervescent tablets attenuated ovalbumin-induced anaphylaxis by upregulating regulatory T cells in mouse models. J. Agric. Food Chem. 2019, 67, 11911–11921. [Google Scholar] [CrossRef] [PubMed]
  329. Pradhan, B.; Patra, S.; Nayak, R.; Behera, C.; Dash, S.R.; Nayak, S.; Sahu, B.B.; Bhutia, S.K.; Jena, M. Multifunctional role of fucoidan, sulfated polysaccharides in human health and disease: A journey under the sea in pursuit of potent therapeutic agents. Int. J. Biol. Macromol. 2020, 164, 4263–4278. [Google Scholar] [CrossRef] [PubMed]
  330. Casas, M.P.; Rodríguez-Hermida, V.; Pérez-Larrán, P.; Conde, E.; Liveri, M.T.; Ribeiro, D.; Fernandes, E.; Domínguez, H. In vitro bioactive properties of phlorotannins recovered from hydrothermal treatment of Sargassum muticum. Sep. Purif. Technol. 2016, 167, 117–126. [Google Scholar] [CrossRef]
  331. Ruocco, N.; Costantini, S.; Guariniello, S.; Costantini, M. Polysaccharides from the marine environment with pharmacological, cosmeceutical and nutraceutical potential. Molecules 2016, 21, 551. [Google Scholar] [CrossRef] [PubMed]
  332. Lee, Y.-E.; Kim, H.; Seo, C.; Park, T.; Lee, K.B.; Yoo, S.-Y.; Hong, S.-C.; Kim, J.T.; Lee, J. Marine polysaccharides: Therapeutic efficacy and biomedical applications. Arch. Pharm. Res. 2017, 40, 1006–1020. [Google Scholar] [CrossRef] [PubMed]
  333. Shao, P.; Chen, X.; Sun, P. In vitro antioxidant and antitumor activities of different sulfated polysaccharides isolated from three algae. Int. J. Biol. Macromol. 2013, 62, 155–161. [Google Scholar] [CrossRef] [PubMed]
  334. Cikoš, A.M.; Jerković, I.; Molnar, M.; Šubarić, D.; Jokić, S. New trends for macroalgal natural products applications. Nat. Prod. Res. 2021, 35, 1180–1191. [Google Scholar] [CrossRef] [PubMed]
  335. Alves, A.; Sousa, R.A.; Reis, R.L. A practical perspective on ulvan extracted from green algae. J. Appl. Phycol. 2013, 25, 407–424. [Google Scholar] [CrossRef]
  336. Wong, T.; Brault, L.; Gasparotto, E.; Vallée, R.; Morvan, P.Y.; Ferrières, V.; Nugier-Chauvin, C. Formation of Amphiphilic Molecules from the Most Common Marine Polysaccharides, toward a Sustainable Alternative? Molecules 2021, 26, 4445. [Google Scholar] [CrossRef]
  337. Costa, L.S.; Fidelis, G.P.; Telles, C.B.S.; Dantas-Santos, N.; Camara, R.B.G.; Cordeiro, S.L.; Costa, M.S.; Almetida-Lima, J.; Oliveira, R.M.; Alburquerque, I.R.; et al. Antioxidant and antiproliferative activities of heterofucans from the seaweed Sargassum filipendula. Mar. Drugs 2011, 9, 952–966. [Google Scholar] [CrossRef] [PubMed]
  338. Qiao, L.; Li, Y.; Chi, Y.; Ji, Y.; Gao, Y.; Hwang, H.; Aker, W.G.; Wang, P. Rheological properties, gelling behavior and texture characteristics of polysaccharide from Enteromorpha prolifera. Carbohydr. Polym. 2016, 136, 1307–1314. [Google Scholar] [CrossRef] [PubMed]
  339. Lin, G.-P.; Wu, D.-S.; Xiao, X.-W.; Huang, Q.-Y.; Chen, H.-B.; Liu, D.; Fu, H.; Chen, X.-H.; Zhao, C. Structural characterization and antioxidant effect of green alga Enteromorpha prolifera polysaccharide in Caenorhabditis elegans via modulation of microRNAs. Int. J. Biol. Macromol. 2020, 150, 1084–1092. [Google Scholar] [CrossRef]
  340. Cindana Mo’o, F.R.; Wilar, G.; Devkota, H.P.; Wathoni, N. Ulvan, a polysaccharide from macroalga Ulva sp.: A review of chemistry, biological activities and potential for food and biomedical applications. Appl. Sci. 2020, 10, 5488. [Google Scholar] [CrossRef]
  341. Qiu, S.M.; Aweya, J.J.; Liu, X.; Liu, Y.; Tang, S.; Zhang, W.; Cheong, K.L. Bioactive polysaccharides from red seaweed as potent food supplements: A systematic review of their extraction, purification, and biological activities. Carbohydr. Polym. 2022, 275, 118696. [Google Scholar] [CrossRef] [PubMed]
  342. Ahmed, A.B.A.; Adel, M.; Karimi, P.; Peidayesh, M. Pharmaceutical, cosmeceutical, and traditional applications of marine carbohydrates. Adv. Food Nutr. Res. 2014, 73, 197–220. [Google Scholar] [PubMed]
  343. Schuster, S.; Fell, D.A.; Dandekar, T. A general definition of metabolic pathways useful for systematic organization and analysis of complex metabolic networks. Nat. Biotechnol. 2000, 18, 326–332. [Google Scholar] [CrossRef] [PubMed]
  344. Samarakoon, K.; Jeon, Y.J. Bio-functionalities of proteins derived from marine algae—A review. Food Res. Int. 2012, 48, 948–960. [Google Scholar] [CrossRef]
  345. 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]
  346. Fabrowska, J.; Łęska, B.; Schroeder, G.; Messyasz, B.; Pikosz, M. Biomass and Extracts of Algae as Material for Cosmetics. In Marine Algae Extracts; Kim, S., Chojnacka, K., Eds.; Wiley: Hoboken, NJ, USA, 2015; pp. 681–706. ISBN 9783527679577. [Google Scholar]
  347. Wells, M.L.; Potin, P.; Craigie, J.S.; Raven, J.A.; Merchant, S.S.; Helliwell, K.E.; Smith, A.G.; Camire, M.E.; Brawley, S.H. Algae as nutritional and functional food sources: Revisiting our understanding. J. Appl. Phycol. 2017, 29, 949–982. [Google Scholar] [CrossRef] [PubMed]
  348. Houston, M.C. Nutrition and nutraceutical supplements in the treatment of hypertension. Expert Rev. Cardiovasc. Ther. 2010, 8, 821–833. [Google Scholar] [CrossRef]
  349. Bedoux, G.; Hardouin, K.; Burlot, A.S.; Bourgougnon, N. Bioactive Components from Seaweeds: Cosmetic Applications and Future Development. In Advances in Botanical Research; Bourgougnon, N., Ed.; Academic Press Inc.: Cambridge, MA, USA, 2014; pp. 345–378. [Google Scholar]
  350. Brown, B.E.; Dunne, R.P. Solar radiation modulates bleaching and damage protection in a shallow water coral. Mar. Ecol. Prog. Ser. 2008, 362, 99–107. [Google Scholar] [CrossRef]
  351. Morais, T.; Cotas, J.; Pacheco, D.; Pereira, L. Seaweeds compounds: An ecosustainable source of cosmetic ingredients? Cosmetics 2021, 8, 8. [Google Scholar] [CrossRef]
  352. Echave, J.; Otero, P.; Garcia-Oliveira, P.; Munekata, P.E.; Pateiro, M.; Lorenzo, J.M.; Simal-Gandara, J.; Prieto, M.A. Seaweed-derived proteins and peptides: Promising marine bioactives. Antioxidants 2022, 11, 176. [Google Scholar] [CrossRef] [PubMed]
  353. Bleakley, S.; Hayes, M. Algal proteins: Extraction, application, and challenges concerning production. Foods 2017, 6, 33. [Google Scholar] [CrossRef] [PubMed]
  354. Samarathunga, J.; Wijesekara, I.; Jayasinghe, M. Seaweed proteins as a novel protein alternative: Types, extractions, and functional food applications. Food Rev. Int. 2023, 39, 4236–4261. [Google Scholar] [CrossRef]
  355. Gressler, V.; Fujii, M.T.; Martins, A.P.; Colepicolo, P.; Mancini-Filho, J.; Pinto, E. Biochemical composition of two red seaweed species grown on the Brazilian coast. J. Sci. Food Agric. 2011, 91, 1687–1692. [Google Scholar] [CrossRef] [PubMed]
  356. Cheung, R.C.F.; Ng, T.B.; Wong, J.H. Marine peptides: Bioactivities and applications. Mar. Drugs 2015, 13, 4006–4043. [Google Scholar] [CrossRef] [PubMed]
  357. Holdt, S.L.; Kraan, S. Bioactive compounds in seaweed: Functional food applications and legislation. J. Appl. Phycol. 2011, 23, 543–597. [Google Scholar] [CrossRef]
  358. Qasim, R. Amino acids composition of some common seaweeds. Pak. J. Pharmac. Sci. 1991, 4, 49–54. [Google Scholar]
  359. Wong, K.H.; Cheung, P.C.K. Nutritional evaluation of some subtropical red and green seaweeds. Part II. In vitro protein digestibility and amino acid profiles of protein concentrates. Food Chem. 2001, 72, 11–17. [Google Scholar] [CrossRef]
  360. Matanjun, P.; Mohamed, S.; Mustapha, N.M.; Muhammad, K. Nutrient content of tropical edible seaweeds, Eucheuma cottoni, Caulerpa lentillifera and Sargassum polycysum. J. Appl. Phycol. 2009, 21, 75–80. [Google Scholar] [CrossRef]
  361. Lordan, S.; Ross, R.P.; Stanton, C. Marine bioactives as functional food ingredients: Potential to reduce the incidence of chronic diseases. Mar. Drugs 2011, 9, 1056–1100. [Google Scholar] [CrossRef] [PubMed]
  362. Aondona, M.M.; Ikya, J.K.; Ukeyima, M.T.; Gborigo TW, J.; Aluko, R.E.; Girgih, A.T. In vitro antioxidant and antihypertensive properties of sesame seed enzymatic protein hydrolysate and ultrafiltration peptide fractions. J. Food Biochem. 2021, 45, e13587. [Google Scholar] [CrossRef] [PubMed]
  363. Meisel, H. Multifunctional peptides encrypted in milk proteins. Biofactors 2004, 21, 55–61. [Google Scholar] [CrossRef] [PubMed]
  364. Murray, B.A.; FitzGerald, R.J. Angiotensin converting enzyme inhibitory peptides derived from food proteins: Biochemistry, bioactivity and production. Curr. Pharmac. Des. 2007, 13, 773–791. [Google Scholar] [CrossRef] [PubMed]
  365. Harnedy, P.A.; FitzGerald, R.J. Bioactive proteins, peptides, and amino acids from macroalgae. J. Phycol. 2011, 47, 218–232. [Google Scholar] [CrossRef] [PubMed]
  366. Orfanoudaki, M.; Hartmann, A.; Karsten, U.; Ganzera, M. Chemical profiling of mycosporine-like amino acids in twenty-three red algal species. J. Phycol. 2019, 55, 393–403. [Google Scholar] [CrossRef] [PubMed]
  367. Mercurio, D.G.; Wagemaker, T.A.L.; Alves, V.M.; Benevenuto, C.G.; Gaspar, L.R.; Campos, P.M. In vivo photoprotective effects of cosmetic formulations containing UV filters, vitamins, Ginkgo biloba and red algae extracts. J. Photochem. Photobiol. B: Biol. 2015, 153, 121–126. [Google Scholar] [CrossRef] [PubMed]
  368. Rangel, K.C.; Villela, L.Z.; de Castro Pereira, K.; Colepicolo, P.; Debonsi, H.M.; Gaspar, L.R. Assessment of the photoprotective potential and toxicity of Antarctic red macroalgae extracts from Curdiea racovitzae and Iridaea cordata for cosmetic use. Algal Res. 2020, 50, 101984. [Google Scholar] [CrossRef]
  369. Barceló-Villalobos, M.; Figueroa, F.L.; Korbee, N.; Álvarez-Gómez, F.; Abreu, M.H. Production of mycosporine-like amino acids from Gracilaria vermiculophylla (Rhodophyta) cultured through one year in an integrated multi-trophic aquaculture (IMTA) system. Mar. Biotechnol. 2017, 19, 246–254. [Google Scholar] [CrossRef] [PubMed]
  370. Athukorala, Y.; Trang, S.; Kwok, C.; Yuan, Y.V. Antiproliferative and antioxidant activities and mycosporine-like amino acid profiles of wild-harvested and cultivated edible Canadian marine red macroalgae. Molecules 2016, 21, 119. [Google Scholar] [CrossRef] [PubMed]
  371. Pliego-Cortés, H.; Bedoux, G.; Boulho, R.; Taupin, L.; Freile-Pelegrín, Y.; Bourgougnon, N.; Robledo, D. Stress tolerance and photoadaptation to solar radiation in Rhodymenia pseudopalmata (Rhodophyta) through mycosporine-like amino acids, phenolic compounds, and pigments in an Integrated Multi-Trophic Aquaculture system. Algal Res. 2019, 41, 101542. [Google Scholar] [CrossRef]
  372. Leelapornpisid, P.; Mungmai, L.; Sirithunyalug, B.; Jiranusornkul, S.; Peerapornpisal, Y. A novel moisturizer extracted from freshwater macroalga [Rhizoclonium hieroglyphicum (C. Agardh) Kützing] for skin care cosmetic. Chiang Mai J. Sci. 2014, 41, 1195–1207. [Google Scholar]
  373. Lee, H.A.; Kim, I.H.; Nam, T.J. Bioactive peptide from Pyropia yezoensis and its anti-inflammatory activities. Int. J. Mol. Med. 2015, 36, 1701–1706. [Google Scholar] [CrossRef] [PubMed]
  374. Ryu, J.; Park, S.J.; Kim, I.H.; Choi, Y.H.; Nam, T.J. Protective effect of porphyra-334 on UVA-induced photoaging in human skin fibroblasts. Int. J. Mol. Med. 2014, 34, 796–803. [Google Scholar] [CrossRef] [PubMed]
  375. Gianeti, M.D.; Maia Campos, P.M. Efficacy evaluation of a multifunctional cosmetic formulation: The benefits of a combination of active antioxidant substances. Molecules. 2014, 19, 18268–18282. [Google Scholar] [CrossRef] [PubMed]
  376. Francavilla, M.; Franchi, M.; Monteleone, M.; Caroppo, C. The red seaweed Gracilaria gracilis as a multi products source. Mar. Drugs 2013, 11, 3754–3776. [Google Scholar] [CrossRef] [PubMed]
  377. Fan, X.; Bai, L.; Mao, X.; Zhang, X. Novel peptides with anti-proliferation activity from the Porphyra haitanesis hydrolysate. Process Biochem. 2017, 60, 98–107. [Google Scholar] [CrossRef]
  378. Verdy, C.; Branka, J.E.; Mekideche, N. Quantitative assessment of lactate and progerin production in normal human cutaneous cells during normal ageing: Effect of an Alaria esculenta extract. Int. J. Cosmet. Sci. 2011, 33, 462–466. [Google Scholar] [CrossRef] [PubMed]
  379. Indumathi, P.; Mehta, A. A novel anticoagulant peptide from the nori hydrolysate. J. Funct. Foods 2016, 20, 606–617. [Google Scholar] [CrossRef]
  380. Jo, C.; Khan, F.F.; Khan, M.I.; Iqbal, J. Marine bioactive peptides: Types, structures, and physiological functions. Food Rev. Int. 2016, 33, 44–61. [Google Scholar] [CrossRef]
  381. Lafarga, T.; Acién-Fernández, F.G.; Garcia-Vaquero, M. Bioactive peptides and carbohydrates from seaweed for food applications: Natural occurrence, isolation, purification, and identification. Algal Res. 2020, 48, 101909. [Google Scholar] [CrossRef]
  382. Sridhar, K.; Inbaraj, B.S.; Chen, B.-H. Recent developments on production, purification and biological activity of marine peptides. Food Res. Int. 2021, 47, 110468. [Google Scholar] [CrossRef] [PubMed]
  383. Pangestuti, R.; Kim, S.K. Seaweed Proteins, Peptides, and Amino Acids. In Seaweed Sustainability: Food and Non-Food Applications; Tiwari, B.K., Toy, D.J., Eds.; Academic Press: San Diego, CA, USA, 2015; pp. 125–140. [Google Scholar]
  384. Yanshin, N.; Kushnareva, A.; Lemesheva, V.; Birkemeyer, C.; Tarakhovskaya, E. Chemical composition and potential practical application of 15 red algal species from the White Sea Coast (the Arctic Ocean). Molecules 2021, 26, 2489. [Google Scholar] [CrossRef] [PubMed]
  385. Subramaniyan, S.A.; Begum, N.; Kim, S.J.; Choi, Y.H.; Nam, T.J. Biopeptides of Pyropia yezoensis and their potential health benefits: A review. Asian Pac. J. Trop. Biomed. 2021, 11, 375–384. [Google Scholar]
  386. De la Coba, F.; Aguilera, J.; De Galvez, M.V.; Alvarez, M.; Gallego, E.; Figueroa, F.L.; Herrera, E. Prevention of the ultraviolet effects on clinical and histopathological changes, as well as the heat shock protein-70 expression in mouse skin by topical application of algal UV-absorbing compounds. J. Dermatol. Sci. 2009, 55, 161–169. [Google Scholar] [CrossRef] [PubMed]
  387. Romarís-Hortas, V.; Bermejo-Barrera, P.; Moreda-Piñeiro, A. Ultrasound-assisted enzymatic hydrolysis for iodinated amino acid extraction from edible seaweed before reversed-phase high performance liquid chromatography-inductively coupled plasma-mass spectrometry. J. Chromatogr. A 2013, 1309, 33–40. [Google Scholar] [CrossRef] [PubMed]
  388. Garcia-Vaquero, M.; Mora, L.; Hayes, M. In vitro and in silico approaches to generating and identifying angiotensin-converting enzyme I inhibitory peptides from green macroalga Ulva lactuca. Mar. Drugs 2019, 17, 204. [Google Scholar] [CrossRef] [PubMed]
  389. Chen, J.C.; Wang, J.; Zheng, B.D.; Pang, J.; Chen, L.J.; Lin, H.T.; Guo, X. Simultaneous Determination of 8 Small Antihypertensive Peptides with Tyrosine at the C-Terminal in Laminaria japonica Hydrolysates by RP-HPLC Method. J. Food Process. Preserv. 2016, 40, 492–501. [Google Scholar] [CrossRef]
  390. Pan, S.; Wang, S.; Jing, L.; Yao, D. Purification and characterisation of a novel angiotensin-I converting enzyme (ACE)-inhibitory peptide derived from the enzymatic hydrolysate of Enteromorpha clathrata protein. Food Chem. 2016, 211, 423–430. [Google Scholar] [CrossRef] [PubMed]
  391. Beaulieu, L.; Bondu, S.; Doiron, K.; Rioux, L.E.; Turgeon, S.L. Characterization of antibacterial activity from protein hydrolysates of the macroalga Saccharina longicruris and identification of peptides implied in bioactivity. J. Funct. Foods 2015, 17, 685–697. [Google Scholar] [CrossRef]
  392. Cermeño, M.; Stack, J.; Tobin, P.R.; O’Keeffe, M.B.; Harnedy, P.A.; Stengel, D.B.; FitzGerald, R.J. Peptide identification from a Porphyra dioica protein hydrolysate with antioxidant, angiotensin converting enzyme and dipeptidyl peptidase IV inhibitory activities. Food Funct. 2019, 10, 3421–3429. [Google Scholar] [CrossRef] [PubMed]
  393. Zheng, Y.; Zhang, Y.; San, S. Efficacy of a novel ACE-inhibitory peptide from Sargassum maclurei in hypertension and reduction of intracellular endothelin-1. Nutrients 2020, 12, 653. [Google Scholar] [CrossRef] [PubMed]
  394. Joel, C.H.; Sutopo, C.C.; Prajitno, A.; Su, J.H.; Hsu, J.L. Screening of angiotensin-I converting enzyme inhibitory peptides derived from Caulerpa lentillifera. Molecules 2018, 23, 3005. [Google Scholar] [CrossRef]
  395. Zhang, X.; Cao, D.; Sun, X.; Sun, S.; Xu, N. Preparation and identification of antioxidant peptides from protein hydrolysate of marine alga Gracilariopsis lemaneiformis. J. Appl. Phycol. 2019, 31, 2585–2596. [Google Scholar] [CrossRef]
  396. Pimentel, F.B.; Machado, M.; Cermeño, M.; Kleekayai, T.; Machado, S.; Rego, A.M.; Abreu, M.H.; Alves, R.C.; Oliveira, M.B.P.; FitzGerald, R.J. Enzyme-Assisted Release of Antioxidant Peptides from Porphyra dioica Conchocelis. Antioxidants 2021, 10, 249. [Google Scholar] [CrossRef] [PubMed]
  397. Park, S.J.; Ryu, J.; Kim, I.H.; Choi, Y.H.; Nam, T.J. Activation of the mTOR signaling pathway in breast cancer MCF-7 cells by a peptide derived from Porphyra yezoensis. Oncol. Rep. 2015, 33, 19–24. [Google Scholar] [CrossRef] [PubMed]
  398. Cian, R.E.; Martínez-Augustin, O.; Drago, S.R. Bioactive properties of peptides obtained by enzymatic hydrolysis from protein byproducts of Porphyra columbina. Food Res. Int. 2012, 49, 364–372. [Google Scholar] [CrossRef]
  399. Beaulieu, L.; Sirois, M.; Tamigneaux, É. Evaluation of the in vitro biological activity of protein hydrolysates of the edible red alga, Palmaria palmata (dulse) harvested from the Gaspe coast and cultivated in tanks. J. Appl. Phycol. 2016, 28, 3101–3115. [Google Scholar] [CrossRef]
  400. Cian, R.E.; López-Posadas, R.; Drago, S.R.; De Medina, F.S.; Martínez-Augustin, O.A. Porphyra columbina hydrolysate upregulates IL-10 production in rat macrophages and lymphocytes through an NF-κB, and p38 and JNK dependent mechanism. Food Chem. 2012, 134, 1982–1990. [Google Scholar] [CrossRef]
  401. Song, J.; Li, T.; Cheng, X.; Ji, X.; Gao, D.; Du, M.; Jiang, N.; Liu, X.; Mao, X. Sea cucumber peptides exert anti-inflammatory activity through suppressing NF-κB and MAPK and inducing HO-1 in RAW264. 7 macrophages. Food Funct. 2016, 7, 2773–2779. [Google Scholar] [CrossRef] [PubMed]
  402. Wu, Q.; Cai, Q.F.; Yoshida, A.; Sun, L.C.; Liu, Y.X.; Liu, G.M.; Su, W.J.; Cao, M.J. Purification and characterization of two novel angiotensin I-converting enzyme inhibitory peptides derived from R-phycoerythrin of red algae (Bangia fusco-purpurea). Eur. Food Res. Technol. 2017, 243, 779–789. [Google Scholar] [CrossRef]
  403. Rosic, N.N. Mycosporine-like amino acids: Making the foundation for organic personalised sunscreens. Mar. Drugs 2019, 17, 638. [Google Scholar] [CrossRef] [PubMed]
  404. Bedoux, G.; Hardouin, K.; Marty, C.; Taupin, L.; Vandanjon, L.; Bourgougnon, N. Chemical characterization and photoprotective activity measurement of extracts from the red macroalga Solieria chordalis. Bot. Mar. 2014, 57, 291–301. [Google Scholar] [CrossRef]
  405. 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] [PubMed]
  406. Shick, J.M.; Dunlap, W.C.; Buettner, G.R. Ultraviolet (UV) Protection in Marine Organisms II, Biosynthesis, Accumulation, and Sunscreening Function of Mycosporine-Like Amino Acids. In Free Radicals in Chemistry, Biology and Medicine; Yoshikawa, S., Toyokuni, S., Yamamoto, Y., Naito, Y., Eds.; OICA International: London, UK, 2000; pp. 215–228. [Google Scholar]
  407. Dunlap, W.C.; Yamamoto, Y. Small-molecule antioxidants in marine organisms: Antioxidant activity of mycosporine-glycine. Compar. Biochem. Physiol. 1995, 112, 105–114. [Google Scholar] [CrossRef]
  408. Bandaranayake, W.M. Mycosporines: Are they nature’s sunscreens? Nat. Prod. Rep. 1998, 15, 159–172. [Google Scholar] [CrossRef] [PubMed]
  409. Conde, F.R.; Churio, M.S.; Previtali, C.M. The photoprotector mechanism of mycosporine-like amino acids. Excited-state properties and photostability of porphyra-334 in aqueous solution. J. Photochem. Photobiol. B Biol. 2000, 56, 139–144. [Google Scholar] [CrossRef] [PubMed]
  410. Gröniger, A.; Sinha, R.P.; Klisch, M.; Häder, D.P. Photoprotective compounds in cyanobacteria, phytoplankton and macroalgae—A database. J. Photochem. Photobiol. B Biol. 2000, 58, 115–122. [Google Scholar] [CrossRef] [PubMed]
  411. Adams, N.L.; Shick, J.M. Mycosporine-like amino acids prevent UVB-induced abnormalities during early development of the green sea urchin Strongylocentrotus droebachiensis. Mar. Biol. 2001, 138, 267–280. [Google Scholar] [CrossRef]
  412. Singh, S.P.; Kumari, S.; Rastogi, R.P.; Singh, K.L.; Sinha, R.P. Mycosporine-like amino acids (MAAs): Chemical structure, biosynthesis and significance as UV-absorbing/screening compounds. Indian J. Experim. Biol. 2008, 46, 7–17. [Google Scholar]
  413. Karsten, U.; Sawall, T.; Wiencke, C. A survey of the distribution of UV-absorbing substances in tropical macroalgae. Phycol. Res. 2006, 46, 271–279. [Google Scholar]
  414. Rastogi, R.P.; Sonani, R.R.; Madamwar, D. UV Photoprotectants from Algae—Synthesis and Bio-Functionalities. In Algal Green Chemistry; Rastogi, R.P., Madamwar, D., Pandey, A., Eds.; Elsevier: Amsterdam, The Netherlands, 2017; pp. 17–38. [Google Scholar]
  415. Sinha, R.P.; Klisch, M.; Gröniger, A.; Häder, D.P. Mycosporine-like amino acids in the marine red alga Gracilaria cornea—effects of UV and heat. Environ. Exp. Bot. 2000, 43, 33–43. [Google Scholar] [CrossRef]
  416. Murphy, C.; Hotchkiss, S.; Worthington, J.; McKeown, S.R. The potential of seaweed as a source of drugs for use in cancer chemotherapy. J. Appl. Phycol. 2014, 26, 2211–2264. [Google Scholar] [CrossRef]
  417. Chrapusta, E.; Kaminski, A.; Duchnik, K.; Bober, B.; Adamski, M.; Bialczyk, J. Mycosporine-like amino acids: Potential health and beauty ingredients. Mar. Drugs 2017, 15, 326. [Google Scholar] [CrossRef] [PubMed]
  418. Arad, S.; Yaron, A. Natural pigments from red microalgae for use in foods and cosmetics. Trends Food Sci. Technol. 1992, 3, 92–97. [Google Scholar] [CrossRef]
  419. Bermejo, R.; Talavera, E.M.; del Valle, C.; Alvarez-Pez, J.M. C-phycocyanin incorporated into reverse micelles: A fluorescence study. Colloids Surf. B Biointerfaces 2000, 18, 51–59. [Google Scholar] [CrossRef]
  420. Chronakis, I.S. Biosolar proteins from aquatic algae. Dev. Food Sci. 2000, 41, 39–75. [Google Scholar]
  421. Rossano, R.; Ungaro, N.; D’Ambrosio, A.; Liuzzi, G.M.; Riccio, P. Extracting and purifying R-phycoeythrin from Mediterranean red algae Corallina elongata Ellis and Solander. J. Biotecnol. 2003, 101, 289–296. [Google Scholar] [CrossRef] [PubMed]
  422. Fleurence, J. Seaweed Proteins. In Proteins in Food Processing; Yada, R.Y., Ed.; Woodhead Publishing Limited: Cambridge, UK, 2004; pp. 197–213. [Google Scholar]
  423. Sekar, S.; Chandramohan, M. Phycobiliproteins as a commodity: Trends in applied research, patents and commercialization. J. Appl. Phycol. 2008, 20, 113–136. [Google Scholar] [CrossRef]
  424. Viskari, P.J.; Colyer, C.L. Rapid extraction of phycobiliproteins from cultures cyanobacteria samples. Anal. Biochem. 2003, 319, 263–271. [Google Scholar] [CrossRef] [PubMed]
  425. Martins, M.; Vieira, F.A.; Correia, I.; Ferreira, R.A.S.; Abreu, H.; Coutinho, J.A.P.; Ventura, S.P.M. Recovery of phycobiliproteins from the red macroalga Gracilaria sp. using ionic liquid aqueous solutions. Green Chem. 2016, 18, 4287–4296. [Google Scholar] [CrossRef]
  426. Capillo, G.; Savoca, S.; Costa, R.; Sanfilippo, M.; Rizzo, C.; Lo Giudice, A.; Albergamo, A.; Rando, R.; Bartolomeo, G.; Spanò, N.; et al. New insights into the culture method and antibacterial potential of Gracilaria gracilis. Marine Drugs 2018, 16, 492. [Google Scholar] [CrossRef] [PubMed]
  427. Saluri, M.; Kaldmäe, M.; Rospu, M.; Sirkel, H.; Paalme, T.; Landreh, M.; Tuvikene, R. Spatial variation and structural characteristics of phycobiliproteins from the red algae Furcellaria lumbricalis and Coccotylus truncatus. Algal Res. 2020, 52, 102058. [Google Scholar] [CrossRef]
  428. Osório, C.; Machado, S.; Peixoto, J.; Bessada, S.; Pimentel, F.B.; Alves, R.C.; Oliveira, M.B.P.P. Pigments content (chlorophylls, fucoxanthin and phycobiliproteins) of different commercial dried algae. Separations 2020, 7, 33. [Google Scholar] [CrossRef]
  429. Tovar-Pérez, E.G.; Guerrero-Legarreta, I.; Farrés-González, A.; Soriano-Santos, J. Angiotensin I-converting enzyme-inhibitory peptide fractions from albumin 1 and globulin as obtained of amaranth grain. Food Chem. 2009, 116, 437–444. [Google Scholar] [CrossRef]
  430. Balti, R.; Bougatef, A.; Sila, A.; Guillochon, D.; Dhulster, P.; Nedjar-Arroume, N. Nine novel angiotensin I-converting enzyme (ACE) inhibitory peptides from cuttlefish (Sepia officinalis) muscle protein hydrolysates and antihypertensive effect of the potent active peptide in spontaneously hypertensive rats. Food Chem. 2015, 170, 519–525. [Google Scholar] [CrossRef]
  431. 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] [PubMed]
  432. Borghi, C.; Tsioufis, K.; Agabiti-Rosei, E.; Burnier, M.; Cicero, A.F.G.; Clement, D.; Coca, A.; Desideri, G.; Grassi, G.; Lovic, D.; et al. Nutraceuticals and blood pressure control: A European Society of Hypertension position document. J. Hypertens. 2020, 38, 799–812. [Google Scholar] [PubMed]
  433. Corinaldesi, C.; Barone, G.; Marcellini, F.; Dell’Anno, A.; Danovaro, R. Marine microbial-derived molecules and their potential use in cosmeceutical and cosmetic products. Marine Drugs. 2017, 15, 118. [Google Scholar] [CrossRef] [PubMed]
  434. Pangestuti, R.; Siahaan, E.A.; Kim, S.K. Photoprotective substances derived from marine algae. Mar. Drugs 2018, 16, 399. [Google Scholar] [CrossRef] [PubMed]
  435. Choi, J.W.; Kwon, M.J.; Kim, I.H.; Kim, Y.M.; Lee, M.K.; Nam, T.J. Pyropia yezoensis glycoprotein promotes the M1 to M2 macrophage phenotypic switch via the STAT3 and STAT6 transcription factors. Int. J. Mol. Med. 2016, 38, 666–674. [Google Scholar] [CrossRef] [PubMed]
  436. Cardozo, K.H.; Marques, L.G.; Carvalho, V.M.; Carignan, M.O.; Pinto, E.; Marinho-Soriano, E.; Colepicolo, P. Analyses of photoprotective compounds in red algae from the Brazilian coast. Rev. Bras. Farmacogn. 2011, 21, 202–208. [Google Scholar] [CrossRef]
  437. Hoyer, K.; Karsten, U.; Sawall, T.; Wiencke, C. Photoprotective substances in Antarctic macroalgae and their variation with respect to depth distribution, different tissues and developmental stages. Mar. Ecol. Prog. Ser. 2001, 211, 117–129. [Google Scholar] [CrossRef]
  438. Oren, A.; Gunde-Cimerman, N. Mycosporines and mycosporine-like amino acids: UV protectants or multipurpose secondary metabolites? FEMS Microbiol. Lett. 2007, 269, 1–10. [Google Scholar] [CrossRef] [PubMed]
  439. Oresajo, C.; Yatskayer, M.; Galdi, A.; Foltis, P.; Pillai, S. Complementary effects of antioxidants and sunscreens in reducing UV-induced skin damage as demonstrated by skin biomarker expression. J. Cosmet. Laser Ther. 2010, 12, 157–162. [Google Scholar] [CrossRef] [PubMed]
  440. Suh, S.S.; Hwang, J.; Park, M.; Seo, H.H.; Kim, H.S.; Lee, J.H.; Moh, S.H.; Lee, T.K. Anti-inflammation activities of mycosporine-like amino acids (MAAs) in response to UV radiation suggest potential anti-skin aging activity. Mar. Drugs 2014, 12, 5174–5187. [Google Scholar] [CrossRef] [PubMed]
  441. de Andrade, C.J.; de Andrade, L.M. An overview on the application of genus Chlorella in biotechnological processes. J. Adv. Res. Biotechnol. 2017, 2, 1–9. [Google Scholar] [CrossRef]
  442. 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]
  443. Kumar, M.S.; Sharma, S.A. Toxicological effects of marine seaweeds: A cautious insight for human consumption. Crit. Rev. Food Sci. Nutr. 2021, 61, 500–521. [Google Scholar] [CrossRef] [PubMed]
  444. Yan, X.; Chuda, Y.; Suzuki, M.; Nagata, T. Fucoxanthin as the major antioxidant in Hijikia fusiformis, a common edible seaweed. Biosci. Biotechnol. Biochem. 1999, 63, 605–607. [Google Scholar] [CrossRef] [PubMed]
  445. Kim, C.R.; Kim, Y.M.; Lee, M.K.; Kim, I.H.; Choi, Y.H.; Nam, T.J. Pyropia yezoensis peptide promotes collagen synthesis by activating the TGF-_/Smad signaling pathway in the human dermal fibroblast cell line Hs27. Int. J. Mol. Med. 2017, 39, 31–38. [Google Scholar] [CrossRef] [PubMed]
  446. Bjarnadóttir, M.; Aðalbjörnsson, B.V.; Nilsson, A.; Slizyte, R.; Roleda, M.Y.; Hreggviðsson, G.; Friðjónsson, H.; Jónsdóttir, R. Palmaria palmata as an alternative protein source: Enzymatic protein extraction, amino acid composition, and nitrogen-to-protein conversion factor. J. Appl. Phycol. 2018, 30, 2061–2070. [Google Scholar] [CrossRef]
  447. Houston, M.C. Nutraceuticals, Vitamins, Antioxidants, and Minerals in the Prevention and Treatment of Hypertension. Prog. Cardiovasc. Dis. 2005, 47, 396–449. [Google Scholar] [CrossRef] [PubMed]
  448. Galland-Irmouli, A.-V.; Fleurence, J.; Lamghari, R.; Luçon, M.; Rouxel, C.; Barbaroux, O.; Bronowicki, J.-P.; Villaume, C.; Guéant, J.-L. Nutritional value of proteins from edible seaweed Palmaria palmata (dulse). J. Nutr. Biochem. 1999, 10, 353–359. [Google Scholar] [CrossRef] [PubMed]
  449. Khairy, H.M.; El-Shafay, S.M. Seasonal variations in the biochemical composition of some common seaweed species from the coast of Abu Qir Bay, Alexandria, Egypt. Oceanologia. 2013, 55, 435–452. [Google Scholar] [CrossRef]
  450. Mohammed, H.O.; O’Grady, M.N.; O’Sullivan, M.G.; Hamill, R.M.; Kilcawley, K.N.; Kerry, J.P. An Assessment of Selected Nutritional, Bioactive, Thermal and Technological Properties of Brown and Red Irish Seaweed Species. Foods 2021, 10, 2784. [Google Scholar] [CrossRef] [PubMed]
  451. Vega, J.; Schneider, G.; Moreira, B.R.; Herrera, C.; Bonomi-Barufi, J.; Figueroa, F.L. Mycosporine-like amino acids from red macroalgae: UV-photoprotectors with potential cosmeceutical applications. Appl. Sci. 2021, 11, 5112. [Google Scholar] [CrossRef]
  452. Araújo, R.G.; Alcantar-Rivera, B.; Meléndez-Sánchez, E.R.; Martínez-Prado, M.A.; Sosa-Hernández, J.E.; Iqbal, H.M.N.; Parra-Saldivar, R.; Martínez-Ruiz, M. Effects of UV and UV-vis Irradiation on the Production of Microalgae and Macroalgae: New Alternatives to Produce Photobioprotectors and Biomedical Compounds. Molecules 2022, 27, 5334. [Google Scholar] [CrossRef] [PubMed]
  453. Chiang, H.M.; Lin, T.J.; Chiu, C.Y.; Chang, C.W.; Hsu, K.C.; Fan, P.C.; Wen, K.C. Coffea arabica extract and its constituents prevent photoaging by suppressing MMPs expression and MAP kinase pathway. Food Chem. Toxicol. 2011, 49, 309–318. [Google Scholar] [CrossRef] [PubMed]
  454. Shen, C.-T.; Chen, P.-Y.; Wu, J.-J.; Lee, T.-M.; Hsu, S.-L.; Chang, C.-M.J.; Young, C.-C.; Shieh, C.-J. Purification of algal antityrosinase zeaxanthin from Nannochloropsis oculata using supercritical anti-solvent precipitation. J. Supercrit. Fluids 2011, 55, 955–9625. [Google Scholar] [CrossRef]
Figure 1. Structural analysis of polysaccharide compounds isolated from seaweeds.
Figure 1. Structural analysis of polysaccharide compounds isolated from seaweeds.
Phycology 04 00015 g001aPhycology 04 00015 g001b
Figure 2. Diagrammatic representation of oligosaccharides derived from red seaweed. (A) Agaro–oligosaccharides and neoagaro–oligosaccharides; (B) κ−carrageenan oligosaccharides; (C) ι−carrageenan oligosaccharides; and (D) λ−carrageenan oligosaccharides. Source: Cheong et al. [342].
Figure 2. Diagrammatic representation of oligosaccharides derived from red seaweed. (A) Agaro–oligosaccharides and neoagaro–oligosaccharides; (B) κ−carrageenan oligosaccharides; (C) ι−carrageenan oligosaccharides; and (D) λ−carrageenan oligosaccharides. Source: Cheong et al. [342].
Phycology 04 00015 g002
Figure 3. Structural representation of bioactive compounds from seaweed: chromophore group of R-phycoerythrin, mycosporine-like amino acids, and peptides. Source: Echave et al. [352].
Figure 3. Structural representation of bioactive compounds from seaweed: chromophore group of R-phycoerythrin, mycosporine-like amino acids, and peptides. Source: Echave et al. [352].
Phycology 04 00015 g003
Table 1. Biological activities and properties of seaweed-derived polysaccharide compounds in skincare.
Table 1. Biological activities and properties of seaweed-derived polysaccharide compounds in skincare.
Seaweed/sCompoundProperties/ActivitiesReferences
Ulva pertusa, Ulva sp.UlvanAntiaging, antiherpetic[153,154]
Saccharina japonica, Chondrus crispus, Codium tomentosumPolysaccharidesHydration[155]
Saccharina longicruris, Laminarin (Sigma)LaminaranReconstructed dermis; skin cell anti-inflammation; antioxidant[156,157]
Ascophyllum nodosum, Chnoospora minima, Ecklonia maxima, Hizikia fusiforme, Saccharina japonica, Sargassum hemiphyllum, Sargassum horneri, Sargassum polycystum, Sargassum vachellianumFucoidansPhotoaging inhibition; minimized elastase activity; antioxidant, anti-inflammatory; collagenase and elastase inhibition; skin-whitening[158,159,160,161,162,163,164,165,166,167,168]
Red seaweeds, Porphyra haitanensis, Gracilaria chouae, Gracilaria blodgettiiCarrageenansAntioxidant, antitumor, antiaging, thickening properties, radiation protection[169,170,171,172,173,174,175]
Brown seaweedsAlginateHigh stability, thickening agent, gelling agent[176,177,178]
Pterocladia, Pterocladiella, Gelidium amansii, GracilariaAgarThickener; antioxidant[179,180,181]
Table 2. Therapeutic potential of seaweed compounds for skin health.
Table 2. Therapeutic potential of seaweed compounds for skin health.
Seaweed/sCompoundProperties/ActivitiesReferences
Sargassum tenerrimumFucoidanAntioxidant: Decrease in DPPH radical and superoxide radical, high total antioxidant and FRAP ability[297,298,299,300]
Costaria costataFucoidanSkin anti-aging: Decrease in UVB-induced mRNA and protein expression of MMP-1, increase in type 1 pro-collagen, and decrease in activation of ERK and JNK[301,302]
Laminaria cichorioidesFucoidanAnti-atopic dermatitis: Reduction in DNCB-induced atopic dermatitis symptoms, including clinical severity scores, scratching counts, and serum histamine levels[303]
Costaria costataFucoidanSkin anti-aging: Decrease in expression of MMP-1 and increase in type 1 pro-collagen[276,304]
Sargassum tenerrimumFucoidanAnti-melanogenesis: Activation of the ERK pathway leading to a decrease in melanin content[305]
MekabFucoidanSkin anti-aging: Decrease in UVB-induced edema, thickness of prickle cell layer, and MMP-1 activity and expression[306]
Ascophyllum nodosumFucoidan (16 kDa)Skin elasticity: Decrease in elastic fiber degradation and leukocyte elastase activity[307]
Ascophyllum nodosumFucoidan (16 kDa)Inflammatory response: Reduction in IL-1β-induced MMP-9 and MMP-3 expression/secretion, increase in TIMP-1[308]
Laminaria cichorioidesFucoidanAtopic dermatitis: Reduction in AD-associated chemokines, including TARC, MDC, and RANTES[309]
Laminaria cichorioidesFucoidanAtopic dermatitis: Decrease in IgE production in PBMC from patients with AD and immunoglobulin germline transcripts of B cells[310]
Saccharina longicrurisLaminaranSkin tissue engineering: Increase in deposition of matrix[311]
Saccharina japonicaFucoidanMoisturizing: Higher moisture absorption and retention ability compared to HA[312]
Laminaria cichorioidesFucoidan (water soluble)Skin cancer prevention: Decrease in EGF or TPA-induced neoplastic cell transformation and binding of EGF and EGFR[313]
Ulva pertusaUlvansAntioxidant: Reduction in superoxide and hydroxyl radicals, increase in reducing power and metal chelating ability[314,315]
Saccharina longicrurisLaminaranSkin anti-aging: Decrease in UVA+UVB-induced skin dermal thickness, increase in Hyp content, decrease in MMP-1 expression, and increase in TIMP-1[316]
Ulva sp.Crude ulvans (57 kDa)Skin anti-aging: Increase in hyaluronan production and decrease in collagen release[317]
Eucheuma spinosum (Eucheuma denticulatum) Eucheuma cottonii (Kappaphycus alvarezii)CarrageenanAntioxidant, Photoprotective: Decrease in UVB-induced cell death, intracellular ROS, and DPPH radical[318]
Porphyra sp.PorphyranAntioxidant: Ferrous ion chelating, increase in reducing power, decrease in DPPH radical and superoxide[319]
Porphyridium sp.CarrageenanAnti-melanogenesis: Decrease in level of melanosome[320]
Porphyra haitanensisPorphyranAntioxidant: Increase in antioxidant enzyme activity, decrease in lipid peroxidation, increase in total antioxidant capacity[321,322]
Porphyra haitanensisPorphyran with different MWAntioxidant: Decrease in DPPH radicals, increase in reducing power[323]
Porphyra yezoensisPorphyran| Anti-inflammation: Decrease in LPS-induced inflammatory markers including NO, iNOS, NF-κB activation, and TNF-α production[324,325]
Porphyra haitanensisLMW Porphyran SD, AD, PD, BDAntioxidant: Decrease in DPPH radicals, hydroxyl radicals, and superoxide[326]
Table 3. Bioactivity of seaweed extracts and seaweed-derived peptide compounds in dermatological applications.
Table 3. Bioactivity of seaweed extracts and seaweed-derived peptide compounds in dermatological applications.
Extract/CompoundActivitySeaweedReferences
Eleven mycosporine-like amino acidsUV-protective effect, antioxidantAgarophyton chilense, Pyropia plicata, Champia novaezelandiae[366]
Mycosporine-like amino acids extractAnti-agingPhorphyra umbilicalis[367]
Mycosporine-like amino acids extractAntioxidant, UV-protective effect, anti-agingCurdieara covitzae, Iridaea cordata[368]
Mycosporine-like amino acids extractAntioxidant; UV-protective effectGracilaria vermiculophylla[369]
Mycosporine-like amino acids extractAntioxidant, antiproliferativeChondrus crispus, Mastocarpus stellatus, Palmaria palmata[370]
Mycosporine-like amino acids extractAntioxidantRhodymenia pseudopalmata[371]
Aqueous extract from freshwater macroalgaSkin moisturizing effectRhizoclonium hieroglyphicum[372]
Peptide PPY1Anti-inflammatoryPyropia yezoensis[373]
Peptides PYP1-5 and porphyra 334Increase in the production of elastin and collagenPorphyra yezoensis f. coreana Ueda[374]
Methanol extract rich in proteins, vitamins, minerals, porphyra-334 and shinorineHydration, skin protective, anti-wrinkle, anti-roughnessPhorphyra umbilicalis[375]
Phycobiliproteins (R-phycoerythrin allophycocyanin and phycocyanin)AntioxidantGracilaria gracilis[376]
Hydrolyzed extractAntitumorPorphyra haitanensis[377]
Algae extractDecrease in progerin production, anti-elastase, and anti-collagenaseAlaria esculenta[378]
Table 4. Biological activities of enzymatically derived compounds from seaweed species.
Table 4. Biological activities of enzymatically derived compounds from seaweed species.
SpeciesHydrolysisBioactivityResultsRef.
Ulva lactucaPapainIn vitro, antihypertensiveACE inhibition (93%) in the >1 kDa hydrolysate fraction[388]
Laminaria japonicaAlcalase, papain, trypsin, and pepsinIn vitro, antihypertensiveACE inhibition. The hydrolysate of all combined proteases achieved an IC50 = 0.6 mg/mL[389]
Enteromorpha clathrataAlcalaseIn vitro, antihypertensiveACE inhibition. IC50 = 0.014 mg/mL[390]
Saccharina longicrurisTrypsinIn vitro, antimicrobialA mix of all peptides (2.50 mg/mL) inhibited 40% of Staphylococcus aureus growth[391]
Porphyra dioicaAlcalase and flavoenzymeIn vitro, antioxidantMost antioxidants on the ORAC assay: IC50 (AFIT) = 0.4 μg/mL, IC50 (MKTPITE) = 0.007 mg/mL[392]
Sargassum maclureiPepsinIn vivo and in vitro, antihypertensiveSystolic blood pressure reduction from 170 to 150 mm Hg at 100 mg/kg bw; 25% endothelin-1 inhibition at 1.5 mg/mL[393]
Caulerpa lentilliferaThermolysinIn vitro, antihypertensiveACE inhibition. IC50 (FDGIP) = 0.03 mg/mL, IC50 (AIDPVRA) = 0.04 mg/mL[394]
Gracilariopsis lemaneiformisα-ChymotrypsinIn vitro, antioxidantDPPH radical scavenging, EC50 = 1.51 mg/mL[395]
Porphyra dioicaProlyveIn vitro, antioxidantORAC (IC50 = 2.7 mmol TE/g), DPPH (IC50 = 0.2 mmol TE/g), FRAP (IC50 = 0.4 mmol TE/g)[396]
Porphyra yezoensisChemical synthesisIn vitro, antitumorDoses ≥ 125 ng/mL induced autophagy and apoptosis in MCF-7 cells via the mTOR pathway[397]
Pyropia columbinaTrypsinIn vitro, antioxidant and anti-inflammatoryDPPH (IC50 = 2.8 mg/mL), ABTS (IC50 = 2.4 mg/mL); upregulation of IL10 at 0.1 mg/mL[398]
Palmaria palmataChymotrypsinIn vitro, antioxidant and antihypertensiveA < 10 kDa fraction was the most bioactive at ≥0.75 mg/mL[399]
Pyropia columbinaFungal proteaseIn vitro, anti-inflammatoryUpregulation of IL-10 in murine spenocytes, macrophages, and lymphocytes at ≥0.01 mg/mL[400]
Neopyropia yezoensisChemical synthesisIn vitro, anti-inflammatoryDoses ≥ 250 ng/mL inhibited the expression of inflammatory cytokines in murine macrophages[401]
Bangia fusco-purpureaPepsinIn vitro, antihypertensiveACE inhibition. IC50(ALLAGDPSVLEDR) = 57.2 mg/mL, IC50(VVGGTGPVDEWGIAGAR) = 66.2 μg/mL[402]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kalasariya, H.S.; Maya-Ramírez, C.E.; Cotas, J.; Pereira, L. Cosmeceutical Significance of Seaweed: A Focus on Carbohydrates and Peptides in Skin Applications. Phycology 2024, 4, 276-313. https://doi.org/10.3390/phycology4020015

AMA Style

Kalasariya HS, Maya-Ramírez CE, Cotas J, Pereira L. Cosmeceutical Significance of Seaweed: A Focus on Carbohydrates and Peptides in Skin Applications. Phycology. 2024; 4(2):276-313. https://doi.org/10.3390/phycology4020015

Chicago/Turabian Style

Kalasariya, Haresh S., Carlos Eliel Maya-Ramírez, João Cotas, and Leonel Pereira. 2024. "Cosmeceutical Significance of Seaweed: A Focus on Carbohydrates and Peptides in Skin Applications" Phycology 4, no. 2: 276-313. https://doi.org/10.3390/phycology4020015

Article Metrics

Back to TopTop