Seaweed Phenolic Compounds: Comparison
Please note this is a comparison between Version 2 by Karina Chen and Version 1 by Ana Marta Gonçalves.

Seaweeds are a potential source of bioactive compounds that are useful for biotechnological applications and can be employed in different industrial areas in order to replace synthetic compounds with components of natural origin. Diverse studies demonstrate that there is a solid ground for the exploitation of seaweed bioactive compounds in order to prevent illness and to ensure a better and healthier lifestyle. Among the bioactive algal molecules, phenolic compounds are produced as secondary metabolites with beneficial effects on plants, and also on human beings and animals, due to their inherent bioactive properties, which exert antioxidant, antiviral, and antimicrobial activities. 

  • seaweeds
  • phenolic compounds
  • bioactive compounds
  • pharmaceutical application
  • nutraceutical application

1. Seaweed Phenolic Compounds

Seaweed phenolic compounds are attracting the attention of the scientific community, as well as several industries, due to their high variety and potential uses [14,37,38]. For instance, the occurrence of phlorotannins (in brown seaweeds) and bromophenols, flavonoids, phenolic terpenoids, and mycosporine-like amino acids (MAAs) in green and red seaweeds has been recorded (

Seaweed phenolic compounds are attracting the attention of the scientific community, as well as several industries, due to their high variety and potential uses [1][2][3]. For instance, the occurrence of phlorotannins (in brown seaweeds) and bromophenols, flavonoids, phenolic terpenoids, and mycosporine-like amino acids (MAAs) in green and red seaweeds has been recorded (

) [4][5].

Phenolic acids consist of a single phenol ring and at least a group of functional carboxylic acids and are typically graded according to the number or the amount of carbon in the chain bound to the phenolic ring. These phenolic acids are also categorized as C6-C1 for hydroxybenzoic acid (HBA; one carbon chain linked to the phenolic ring), C6-C2 for acetophenones and phenylacetic acids (two carbon chains linked to the phenolic ring) and C6-C3 (3 carbon chains attached to the phenol ring) for hydroxycinnamic acid (HCA) [41,42]. HBAs include, among others, gallic acid, p-hydroxybenzoic acid, vanillic acid, syringic acid, and protocatechins, in which there are differences in the basic structure of the HBA, including an aromatic ring hydroxylation and methoxylation [41,42].

Phenolic acids consist of a single phenol ring and at least a group of functional carboxylic acids and are typically graded according to the number or the amount of carbon in the chain bound to the phenolic ring. These phenolic acids are also categorized as C6-C1 for hydroxybenzoic acid (HBA; one carbon chain linked to the phenolic ring), C6-C2 for acetophenones and phenylacetic acids (two carbon chains linked to the phenolic ring) and C6-C3 (3 carbon chains attached to the phenol ring) for hydroxycinnamic acid (HCA) [6][7]. HBAs include, among others, gallic acid, p-hydroxybenzoic acid, vanillic acid, syringic acid, and protocatechins, in which there are differences in the basic structure of the HBA, including an aromatic ring hydroxylation and methoxylation [6][7].

Trans-phenyl-3-propenoic acids are hydroxycinnamic acids (HCA), which vary in their ring constitution [41]. These HCA derivatives include caffeic (3,4-dihydroxycinnamic), ferulic (3-methoxy-4-hydroxy), sinapic (3,5-dimethoxy-4-hydroxy), and p-coumaric (4-hydroxy) acids, all of which are commonly distributed as conjugates, primarily as quinic acid esters (chlorogenic acids) [41,42]. In addition, these acids can be subcategorized up into different groups based on the identity, location, and number of the acyl residue: (1) mono-esters of caffeic, ferulic, and p-coumaric acids; (2) bi-, tri-, and tetra-esters of caffeic acids; (3) mixed di-esters of caffeic-ferulic acid or caffeic-sinapic acids; and (4) mixed caffeic acid esters with aliphatic dibasic acids, such as oxalic or succinic acid [41,42].

Trans-phenyl-3-propenoic acids are hydroxycinnamic acids (HCA), which vary in their ring constitution [6]. These HCA derivatives include caffeic (3,4-dihydroxycinnamic), ferulic (3-methoxy-4-hydroxy), sinapic (3,5-dimethoxy-4-hydroxy), and p-coumaric (4-hydroxy) acids, all of which are commonly distributed as conjugates, primarily as quinic acid esters (chlorogenic acids) [6][7]. In addition, these acids can be subcategorized up into different groups based on the identity, location, and number of the acyl residue: (1) mono-esters of caffeic, ferulic, and p-coumaric acids; (2) bi-, tri-, and tetra-esters of caffeic acids; (3) mixed di-esters of caffeic-ferulic acid or caffeic-sinapic acids; and (4) mixed caffeic acid esters with aliphatic dibasic acids, such as oxalic or succinic acid [6][7].

Some experiments have shown the presence of phenolic acids in marine algae [41,42,43]. For instance, coumarins have been found in green seaweed species such as

Some experiments have shown the presence of phenolic acids in marine algae [6][7][8]. For instance, coumarins have been found in green seaweed species such as

Dasycladus vermicularis

, as well as some vanillic acid derivatives in the

Cladophora socialis (Chlorophyta, green algae) [44].

(Chlorophyta, green algae) [9].

Ascophyllum nodosum

(

A),

Bifurcaria bifurcata

(

B), and

Fucus vesiculosus

(

Figure 1C) (Phaeophyceae, brown algae) have been distinguished by the presence of HBAs, rosmarinic acid, and quinic acid [45]. In addition, in the genus

C) (Phaeophyceae, brown algae) have been distinguished by the presence of HBAs, rosmarinic acid, and quinic acid [10]. In addition, in the genus

Gracilaria

(

Figure 1I) (Rhodophyta, red alga), phenolic acids have been detected, such as benzoic acid, p-hydroxybenzoic acid, salicylic acid, gentisic acid, protocatechuic acid, vanillic acid, gallic acid, and syringic acid [46,47,48].

I) (Rhodophyta, red alga), phenolic acids have been detected, such as benzoic acid, p-hydroxybenzoic acid, salicylic acid, gentisic acid, protocatechuic acid, vanillic acid, gallic acid, and syringic acid [11][12][13].

Figure 1.

Some seaweeds producing phenolic compounds: (

A

)—

Ascophyllum nodosum

(P); (

B

)—

Bifurcaria bifurcata

(P); (

C

)—

Fucus vesiculosus

(P); (

D

)—

Leathesia marina

(P); (

E

)—

Lobophora variegata

(P); (

F

)—

Macrocystis pyrifera

(P); (

G

)—

Asparagopsis armata

(R); (

H

)—

Chondrus crispus

(R); (

I

)—

Gracilaria

sp. (R); (

J

)—

Kappaphycus alvarezii

(R); (

K

)—

Neopyropia

sp. (R); (

L

)—

Palmaria palmata

(R); (

M

)—

Dasycladus vermicularis

(Chl); (

N

)—

Derbesia tenuissima

(Chl); (

O

)—

Ulva intestinalis

(Chl);

P—Phaeophyceae

, R—Rhodophyta; Chl—Chlorophyta.

Phlorotannins are well-known phenolic compounds synthesized by brown seaweeds. These compounds are constituted by oligomeric units of phloroglucinol [49,50]. Commonly, these secondary metabolites have a molecular weight ranging from 10 to 100 kDa, due to the high variability that these molecules can present in the structural bonds between phloroglucinol and the hydroxyl groups [51,52]. In this context, phlorotannins can be categorized into six categories: (1) fucols (aryl–aryl bonds), (2) phloretols (aryl–ether bonds), (3) eckols (dibenzo-1,4-dioxin bonds), (4) fucophloretols (ether or phenyl linage), (5) carmalols (dibenzodioxin moiety), and (6) fuhalols (ortho-/para- arranged ether bridges containing an additional hydroxyl group on one unit) [49,51,52]. Moreover, the complexity of these molecules classify them, by each category, into linear or branched phlorotannins [51,52]. Due to its biotechnological properties, dieckol is the most exploited phlorotannin, and it can be found in the species

Phlorotannins are well-known phenolic compounds synthesized by brown seaweeds. These compounds are constituted by oligomeric units of phloroglucinol [14][15]. Commonly, these secondary metabolites have a molecular weight ranging from 10 to 100 kDa, due to the high variability that these molecules can present in the structural bonds between phloroglucinol and the hydroxyl groups [16][17]. In this context, phlorotannins can be categorized into six categories: (1) fucols (aryl–aryl bonds), (2) phloretols (aryl–ether bonds), (3) eckols (dibenzo-1,4-dioxin bonds), (4) fucophloretols (ether or phenyl linage), (5) carmalols (dibenzodioxin moiety), and (6) fuhalols (ortho-/para- arranged ether bridges containing an additional hydroxyl group on one unit) [14][16][17]. Moreover, the complexity of these molecules classify them, by each category, into linear or branched phlorotannins [16][17]. Due to its biotechnological properties, dieckol is the most exploited phlorotannin, and it can be found in the species

Ecklonia cava (Phaeophyceae) [53].

(Phaeophyceae) [18].

Flavonoids are structurally characterized as phenolic compounds with a heterocyclic oxygen bound to two aromatic rings, which can then differ according to the degree of hydrogenation [54,55]. However, there is a generalized lack of studies regarding algal flavonoids’ isolation and characterization. Nevertheless, some research has shown that seaweeds are a rich source of flavonoids. Several species of the Chlorophyta, Rhodophyta phyla, and Phaeophyceae class were found to have flavonoids such as rutin, quercitin, and hesperidin [49,56]. For instance,

Flavonoids are structurally characterized as phenolic compounds with a heterocyclic oxygen bound to two aromatic rings, which can then differ according to the degree of hydrogenation [19][20]. However, there is a generalized lack of studies regarding algal flavonoids’ isolation and characterization. Nevertheless, some research has shown that seaweeds are a rich source of flavonoids. Several species of the Chlorophyta, Rhodophyta phyla, and Phaeophyceae class were found to have flavonoids such as rutin, quercitin, and hesperidin [14][21]. For instance,

Chondrus crispus

(

H) and

Porphyra

/

Pyropia

spp. (Rhodophyta) and

Sargassum muticum

and

Sargassum vulgare (Phaeophyceae) can synthesize isoflavones, likewise daidzein or genistein [57]. Moreover, many flavonoid glycosides have also been recorded in the brown seaweeds

(Phaeophyceae) can synthesize isoflavones, likewise daidzein or genistein [22]. Moreover, many flavonoid glycosides have also been recorded in the brown seaweeds

Durvillaea antarctica

,

Lessonia spicata

, and

Macrocystis pyrifera

(also known as

Macrocystis integrifolia

) (

Figure 1F) [49]. Furthermore, green (

F) [14]. Furthermore, green (

Acetabularia ryukyuensis

), brown (

Eisenia bicyclis

—as

Ecklonia bicyclis

,

Padina arborescens

,

Padina minor

), and red seaweeds (

Neopyropia yezoensis

—also known as

Porphyra yezoensis

K,

Gelidium elegans

, and

Portieria hornemannii

—also known as

Chondrococcus hornemannii) proved to be a valuable source of catechin, epicatechin, epigallocatechin, catechin gallate, epicatechin gallate, or epigallocatechin gallate [58].

) proved to be a valuable source of catechin, epicatechin, epigallocatechin, catechin gallate, epicatechin gallate, or epigallocatechin gallate [23].

Bromophenols are brominated phenolic compounds characterized by the presence of one or more benzene rings and hydroxyl substituents [59,60]. These compounds can be found in green [61,62,63,64], red [65,66,67] and brown seaweeds [68,69]. Nevertheless, red seaweeds often exhibit a higher content of these molecules [70]. However, due to the low content of bromophenols in seaweeds, there are just a few studies regarding the isolation and characterization of these compounds.

Bromophenols are brominated phenolic compounds characterized by the presence of one or more benzene rings and hydroxyl substituents [24][25]. These compounds can be found in green [26][27][28][29], red [30][31][32] and brown seaweeds [33][34]. Nevertheless, red seaweeds often exhibit a higher content of these molecules [35]. However, due to the low content of bromophenols in seaweeds, there are just a few studies regarding the isolation and characterization of these compounds.

Phenolic terpenoids are secondary metabolites that have already been identified in seaweeds [50]. For instance, meroditerpenoids (such as plastoquinones, chromanols, and chromenes) were found in brown seaweeds, mainly from the family Sargassaceae (Phaeophyceae). These compounds are partially derived from terpenoids and are characterized for having a polyprenyl chain linked to a hydroquinone ring moiety [71]. Red seaweeds also synthesize phenolic terpenoids, such as diterpenes and sesquiterpenes in Rhodomelaceae. For example, the species

Phenolic terpenoids are secondary metabolites that have already been identified in seaweeds [15]. For instance, meroditerpenoids (such as plastoquinones, chromanols, and chromenes) were found in brown seaweeds, mainly from the family Sargassaceae (Phaeophyceae). These compounds are partially derived from terpenoids and are characterized for having a polyprenyl chain linked to a hydroquinone ring moiety [36]. Red seaweeds also synthesize phenolic terpenoids, such as diterpenes and sesquiterpenes in Rhodomelaceae. For example, the species

Callophycus serratus synthesizes a specific diterpene, bromophycolide [72].

synthesizes a specific diterpene, bromophycolide [37].

Table 1.

Seaweed phenolic compounds recorded, according to phyla and phenolic compound group.

Mycosporine-like amino acids (MAAs) are secondary metabolites that, despite being synthesized by several organisms, were found to be more often produced by marine organisms [80,81,82]. Such compounds present a low molecular weight (<400 kDa) and are soluble in water. Moreover, they present a cyclohexanone or cyclohexenine ring, with amino acid moieties in their chemical structure [80,83]. Thus, these compounds can be found mainly in red seaweeds. For example, it was found that the edible red seaweed

Mycosporine-like amino acids (MAAs) are secondary metabolites that, despite being synthesized by several organisms, were found to be more often produced by marine organisms [45][46][47]. Such compounds present a low molecular weight (<400 kDa) and are soluble in water. Moreover, they present a cyclohexanone or cyclohexenine ring, with amino acid moieties in their chemical structure [45][48]. Thus, these compounds can be found mainly in red seaweeds. For example, it was found that the edible red seaweed

Palmaria palmata

(

Figure 1L) biosynthesizes the MAA palythine, shinorine, asterina-330, palythinol, and porphyra-334 [73]. In addition, the tetrasporophyte phase of

L) biosynthesizes the MAA palythine, shinorine, asterina-330, palythinol, and porphyra-334 [38]. In addition, the tetrasporophyte phase of

Asparagopsis armata

(

Figure 1G) was found to produce palythine and shinorine [74].

G) was found to produce palythine and shinorine [39].

 

2. Phenolic Compounds Application in Biotechnology

Biological compounds extracted from seaweeds exert several activities that can be exploited for the production of food, animal feed, and new drugs, substituting synthetic compounds with natural-origin compounds.

The most exploited phenolic compounds are phlorotannins, which are exclusively present in high concentration in brown seaweeds [17,124] and are involved in defense activities [150,151,152], showing strong antioxidant properties and antimicrobial activity, which help to inhibit bacterial growth [17]. Phlorotannins can be exploited in different biotechnological sectors. They exert a powerful antioxidant activity, as in the case of phlorotannins extracted from

The most exploited phenolic compounds are phlorotannins, which are exclusively present in high concentration in brown seaweeds [49][50] and are involved in defense activities [51][52][53], showing strong antioxidant properties and antimicrobial activity, which help to inhibit bacterial growth [49]. Phlorotannins can be exploited in different biotechnological sectors. They exert a powerful antioxidant activity, as in the case of phlorotannins extracted from

Sirophysalis trinodis

(formerly known as

Cystoseira trinodis, Phaeophyceae), which makes considering this species a potential source of phenolic compounds for diverse applications [153].

, Phaeophyceae), which makes considering this species a potential source of phenolic compounds for diverse applications [54].

The synthesis of these compounds is driven by different factors. For example, seaweeds are particularly sensitive to external stressors; consequentially, they produce phenolic compounds, which develop multiple activities in order to protect seaweeds [153,154,155,156,157,158]. Due to several biological activities that involve phenolic compounds, they have been found interesting to be applied in the nutraceutical, pharmaceutical, medical, and industrial areas [34,75,159].

The synthesis of these compounds is driven by different factors. For example, seaweeds are particularly sensitive to external stressors; consequentially, they produce phenolic compounds, which develop multiple activities in order to protect seaweeds [54][55][56][57][58][59]. Due to several biological activities that involve phenolic compounds, they have been found interesting to be applied in the nutraceutical, pharmaceutical, medical, and industrial areas [60][40][61].

2.1. Medical and Pharmaceutical Applications

The consumption of seaweeds can prevent diseases or help the healing. Their bioactive compounds have positive effects on human health. For example, Tanniou et al. [149] identified the brown alga

The consumption of seaweeds can prevent diseases or help the healing. Their bioactive compounds have positive effects on human health. For example, Tanniou et al. [62] identified the brown alga

Sargassum muticum as a potential source of bioactive phenolic compounds: this species showed a strong antioxidant activity [150] and anti-proliferative activity in breast cancer cells [160] that may suggest the involvement of

as a potential source of bioactive phenolic compounds: this species showed a strong antioxidant activity [51] and anti-proliferative activity in breast cancer cells [63] that may suggest the involvement of

S. muticum

in biotechnological applications.

Shibata et al. [161] compared the antioxidant activity of phlorotannins extracted from

Shibata et al. [64] compared the antioxidant activity of phlorotannins extracted from

Eisenia bicyclis

(Phaeophyceae) in vitro to available and active compounds such as vitamin C (ascorbic acid) and vitamin E (α-tocopherol). This study demonstrates that the antioxidant activity of phlorotannin was 10 times higher than that of other biological compounds.

The isolation and studies on phlorotannin derivates demonstrate that their high anti-proliferation activity is able to induce growth inhibition and apoptosis in human breast cancer cells [162,163]. For example, the red seaweed

The isolation and studies on phlorotannin derivates demonstrate that their high anti-proliferation activity is able to induce growth inhibition and apoptosis in human breast cancer cells [65][66]. For example, the red seaweed

Kappaphycus alvarezii

(also known as

Eucheuma cottonii

) (

Figure 1J) polyphenol in vitro extracts were analyzed to evaluate antiproliferative, apoptotic, and cell cycle effects. Results showed an effect of these compounds against cancer cells [164].

J) polyphenol in vitro extracts were analyzed to evaluate antiproliferative, apoptotic, and cell cycle effects. Results showed an effect of these compounds against cancer cells [67].

The uptake of phlorotannins has also been related to the reduction in cardiovascular diseases and hypercholesterolemia [165,166].

The uptake of phlorotannins has also been related to the reduction in cardiovascular diseases and hypercholesterolemia [68][69].

Phlorotannins are responsible for the absorption of UV-B radiation [167,168,169,170], acting as photoprotective agent for algal cells [38,171], to avoid DNA damage [172,173,174]. This property is also effective for human and animal skin, reducing the probability of skin cancer due to UV-B radiation [168]. Additionally, phlorotannins prevent the production of matrix metalloproteinases (MMPs), enzymes that encourage the presence of wrinkles by degrading the extracellular matrix. For this purpose, seaweed phenolic compounds may be involved in the production of anti-aging creams and skin products [175].

Phlorotannins are responsible for the absorption of UV-B radiation [70][71][72][73], acting as photoprotective agent for algal cells [3][74], to avoid DNA damage [75][76][77]. This property is also effective for human and animal skin, reducing the probability of skin cancer due to UV-B radiation [71]. Additionally, phlorotannins prevent the production of matrix metalloproteinases (MMPs), enzymes that encourage the presence of wrinkles by degrading the extracellular matrix. For this purpose, seaweed phenolic compounds may be involved in the production of anti-aging creams and skin products [78].

Phlorotannins are also involved in the development of therapies to treat diverse allergic diseases. In Korean traditional medicine, phlorotannin extracts from the brown alga

Sargassum hemiphyllum

and the red alga

Polyopes affinis

(formerly known as

Carpopeltis affinis) have been confirmed to have effective antiallergic properties in vitro [176]. The Japanese brown alga

) have been confirmed to have effective antiallergic properties in vitro [79]. The Japanese brown alga

Ecklonia arborea

(formerly known as

Eisenia arborea

) has been found to contain effective inhibitors of histamine; the presence of phlorofucofuroeckol B (phlorotannin) may be the reason for the anti-allergic activity shown in rats.

Ecklonia arborea is popular in Japan since it has been consumed for years as healthy food and folkloristic therapies [177].

is popular in Japan since it has been consumed for years as healthy food and folkloristic therapies [80].

Among phenolic compounds, bromophenol and its derivates are widely investigated due to their potential activities. Studies conducted with

Leathesia marina

(formerly known as

Leathesia nana

) (

Figure 1D) (Phaeophyceae) indicate that bromophenol derivatives respond positively to the inhibition of human cancer cells proliferation in vitro [178]. Alongside the ideal exploitation of bromophenol derivates for the development of new therapies for tumor treatment, these biological compounds reported antiviral activity against Herpes Simplex Viruses-1. For instance, extracts from the red alga

D) (Phaeophyceae) indicate that bromophenol derivatives respond positively to the inhibition of human cancer cells proliferation in vitro [81]. Alongside the ideal exploitation of bromophenol derivates for the development of new therapies for tumor treatment, these biological compounds reported antiviral activity against Herpes Simplex Viruses-1. For instance, extracts from the red alga

Symphyocladia latiuscula

(

Figure 1N), which is abundant in Korea, demonstrate antiviral activity against HSV-1, likely due to the presence of its bromophenols, the major compounds [179].

N), which is abundant in Korea, demonstrate antiviral activity against HSV-1, likely due to the presence of its bromophenols, the major compounds [82].

Moreover, researchers have proven the antimicrobial effect of bromophenols extracted from the red alga

Rhodomela confervoides

, which act against some

Staphylococcus

and

Pseudomonas aeruginosa strains [180].

strains [83].

Advantages of Phenolic Compounds Consumption for Human Health

Benefits of phenolic compounds are very common in human diet, since they can be ingested as food or food supplements and provide the human organisms with multiple positive effects [181]. They can be found in food and beverages from natural origin such as plants, seaweeds, fruits, coffee, black tea, and chocolate [182,183], but they can be also added to our daily diet as colorants or as antioxidants [184].

Benefits of phenolic compounds are very common in human diet, since they can be ingested as food or food supplements and provide the human organisms with multiple positive effects [84]. They can be found in food and beverages from natural origin such as plants, seaweeds, fruits, coffee, black tea, and chocolate [85][86], but they can be also added to our daily diet as colorants or as antioxidants [87].

Many synthetic antioxidants have been developed to retard the oxidation in foods. However, synthetic compounds may have collateral effects [185] that could be avoided by the intake of natural antioxidant compounds, such as phenolic compounds extracted from seaweeds [186]. Phenolic acids present in food are also responsible for organoleptic properties, influencing color, flavor, and nutritional values [187].

Many synthetic antioxidants have been developed to retard the oxidation in foods. However, synthetic compounds may have collateral effects [88] that could be avoided by the intake of natural antioxidant compounds, such as phenolic compounds extracted from seaweeds [89]. Phenolic acids present in food are also responsible for organoleptic properties, influencing color, flavor, and nutritional values [90].

Brown algae have already been exploited as food in Asia in the past 15 centuries; phlorotannin extracts from

Ecklonia cava

are already available in the market since 2018, when the European Food Safety Authority (EFSA) Panel on Dietetic Products, Nutrition and Allergies (NDA) attested that these extracts are indicated for diet due to their nutritional properties.

Ecklonia cava

thallus is consumed as salad and as a component of soups, while

E. cava powder is also used to dye food, especially sweets, such as candies or rice cakes [188].

powder is also used to dye food, especially sweets, such as candies or rice cakes [91].

Phlorotannins have anti-diabetic effects: Roy et al. [189] assessed the in vitro inhibitory activity of phlorotannins extracted from

Phlorotannins have anti-diabetic effects: Roy et al. [92] assessed the in vitro inhibitory activity of phlorotannins extracted from

Ascophyllum nodosum

and

Fucus vesiculosus, and their effect on rat blood glucose and insulin levels. It has been noticed that, 20 min after the consumption of animal feed enriched in phlorotannins, the normal increase in postprandial blood glucose was reduced by 90%, with a consequential reduction by 40% of insulin secretion [189].

, and their effect on rat blood glucose and insulin levels. It has been noticed that, 20 min after the consumption of animal feed enriched in phlorotannins, the normal increase in postprandial blood glucose was reduced by 90%, with a consequential reduction by 40% of insulin secretion [92].

As different classes of polyphenols from seaweeds can assure health benefits, it is suggested to consume the whole algae in order to uptake a higher quantity of bioactive compounds, instead of consuming only algae extracts as food supplements [190].

As different classes of polyphenols from seaweeds can assure health benefits, it is suggested to consume the whole algae in order to uptake a higher quantity of bioactive compounds, instead of consuming only algae extracts as food supplements [93].

Flavonoids have been investigated for a long time for their powerful antioxidant activities. Their uptake has been linked with a reduced risk of lung cancer [181].

Flavonoids have been investigated for a long time for their powerful antioxidant activities. Their uptake has been linked with a reduced risk of lung cancer [84].

2.2. Aquaculture and Industrial Applications

Bromophenols are also investigated for the flavor they give to seafood [113,191,192]. Studies attested that bromophenols are responsible for the typical iodine-like flavor of marine fish [192], prawns, and marine algae [113]. It is quite likely that bromophenols detected in marine fish and prawns derived from their diet based on seaweeds that can synthesize these compounds [113,193].

Bromophenols are also investigated for the flavor they give to seafood [94][95][96]. Studies attested that bromophenols are responsible for the typical iodine-like flavor of marine fish [96], prawns, and marine algae [94]. It is quite likely that bromophenols detected in marine fish and prawns derived from their diet based on seaweeds that can synthesize these compounds [94][97].

The Japanese brown algae

Padina

spp.,

Sargassum

spp., and

Lobophora

spp. (

Figure 1E) have been detected as sources of bromophenols for local fish. It is likely that fish assimilate the typical marine flavor after the ingestion of these algae [69].

E) have been detected as sources of bromophenols for local fish. It is likely that fish assimilate the typical marine flavor after the ingestion of these algae [34].

The presence of bromophenols in the diet of prawns may be useful for aquaculture [113,192]: crustaceans used as fish feed in aquaculture systems have low amounts of bromophenols due to their diet, with a consequential absence of iodine-like flavor in farmed fish [192]. The inclusion of seaweeds in prawns feed may thus increase the sea-like flavor of aquaculture seafood, enhancing their taste [69]

The presence of bromophenols in the diet of prawns may be useful for aquaculture [94][96]: crustaceans used as fish feed in aquaculture systems have low amounts of bromophenols due to their diet, with a consequential absence of iodine-like flavor in farmed fish [96]. The inclusion of seaweeds in prawns feed may thus increase the sea-like flavor of aquaculture seafood, enhancing their taste [34]

Moreover, other compounds, such as flavonoids, play an important role in retarding lipid oxidation that occurs in muscle, especially in fish, in order to delay the deterioration of seafood [186,194].

Moreover, other compounds, such as flavonoids, play an important role in retarding lipid oxidation that occurs in muscle, especially in fish, in order to delay the deterioration of seafood [89][98].

Over the last years, textile industries dedicated more attention towards medical textiles since their usage is not restricted to medical centers and care facilities: it is also present in other fields where hygienic conditions are required, e.g., hotels or restaurants [195]. Natural fibers such as cotton or silk are limited; therefore, medical textile industries started to use synthetic fibers, such as polyester, viscose, polyamides, and polypropylene [195]. A critical problem with synthetic fibers is the risk of spreading infections. To overcome this problem, seaweeds’ bio-compounds may be utilized for textile production. Due to the properties of phenolic compounds, new biological textiles may be developed. The new textiles could have antioxidant and antimicrobial properties [196] with the advantage of being natural and not irritating to the skin and being biodegradable and biocompatible [197,198]. The natural bioactive agents are non-toxic and skin and eco-friendly. From the extraction and treatment of cellulose-based polyphenols, these textiles can be brought into contact with the human skin and tissues and body fluids [195].

Over the last years, textile industries dedicated more attention towards medical textiles since their usage is not restricted to medical centers and care facilities: it is also present in other fields where hygienic conditions are required, e.g., hotels or restaurants [99]. Natural fibers such as cotton or silk are limited; therefore, medical textile industries started to use synthetic fibers, such as polyester, viscose, polyamides, and polypropylene [99]. A critical problem with synthetic fibers is the risk of spreading infections. To overcome this problem, seaweeds’ bio-compounds may be utilized for textile production. Due to the properties of phenolic compounds, new biological textiles may be developed. The new textiles could have antioxidant and antimicrobial properties [100] with the advantage of being natural and not irritating to the skin and being biodegradable and biocompatible [101][102]. The natural bioactive agents are non-toxic and skin and eco-friendly. From the extraction and treatment of cellulose-based polyphenols, these textiles can be brought into contact with the human skin and tissues and body fluids [99].

Moreover, the use of flavonoids to obtain UV-protective clothing has been suggested, since they show UV protection ability linked with antibacterial and anti-inflammatory properties [195].

Moreover, the use of flavonoids to obtain UV-protective clothing has been suggested, since they show UV protection ability linked with antibacterial and anti-inflammatory properties [99].

References

  1. Mekinić, I.G.; Skroza, D.; Šimat, V.; Hamed, I.; Čagalj, M.; Perković, Z.P. Phenolic content of brown algae (Phaeophyceae) species: Extraction, identification, and quantification. Biomolecules 2019, 9, 244.
  2. Lopes, G.; Pinto, E.; Andrade, P.B.; Valentão, P. Antifungal Activity of Phlorotannins against Dermatophytes and Yeasts: Approaches to the Mechanism of Action and Influence on Candida albicans Virulence Factor. PLoS ONE 2013, 8.
  3. Wijesekara, I.; Kim, S.K.; Li, Y.X.; Li, Y.X. Phlorotannins as bioactive agents from brown algae. Process Biochem. 2011, 46, 2219–2224.
  4. 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.
  5. Gómez-Guzmán, M.; Rodríguez-Nogales, A.; Algieri, F.; Gálvez, J. Potential role of seaweed polyphenols in cardiovascular-associated disorders. Mar. Drugs 2018, 16, 250.
  6. Pietta, P.; Minoggio, M.; Bramati, L. Plant Polyphenols: Structure, Occurrence and Bioactivity. In Studies in Natural Products Chemistry; Rahman, A., Ed.; Elsevier: Amsterdam, The Netherlands, 2003; pp. 257–312.
  7. Luna-Guevara, M.L.; Luna-Guevara, J.J.; Hernández-Carranza, P.; Ruíz-Espinosa, H.; Ochoa-Velasco, C.E. Phenolic Compounds: A Good Choice Against Chronic Degenerative Diseases. In Studies in Natural Products Chemistry; Rahman, A., Ed.; Elsevier: Amsterdam, The Netherlands, 2018; pp. 79–108.
  8. Mancini-Filho, J.; Novoa, A.V.; González, A.E.B.; de Andrade-Wartha, E.R.S.; Mancini, D.A.P. Free Phenolic Acids from the Seaweed Halimeda monile with Antioxidant Effect Protecting against Liver Injury. Zeitschrift für Naturforsch. C 2009, 64, 657–663.
  9. Feng, Y.; Carroll, A.R.; Addepalli, R.; Fechner, G.A.; Avery, V.M.; Quinn, R.J. Vanillic acid derivatives from the green algae Cladophora socialis as potent protein tyrosine phosphatase 1B inhibitors. J. Nat. Prod. 2007, 70, 1790–1792.
  10. Agregán, R.; Munekata, P.E.S.; Franco, D.; Dominguez, R.; Carballo, J.; Lorenzo, J.M. Phenolic compounds from three brown seaweed species using LC-DAD–ESI-MS/MS. Food Res. Int. 2017, 99, 979–985.
  11. Farvin, K.H.S.; Jacobsen, C.; Sabeena Farvin, K.H.; Jacobsen, C. Phenolic compounds and antioxidant activities of selected species of seaweeds from Danish coast. Food Chem. 2013, 138, 1670–1681.
  12. Xu, T.; Sutour, S.; Casabianca, H.; Tomi, F.; Paoli, M.; Garrido, M.; Pasqualini, V.; Aiello, A.; Castola, V.; Bighelli, A. Rapid Screening of Chemical Compositions of Gracilaria dura and Hypnea mucisformis (Rhodophyta) from Corsican Lagoon. Int. J. Phytocosmetics Nat. Ingred. 2015, 2, 8.
  13. Souza, B.W.S.; Cerqueira, M.A.; Martins, J.T.; Quintas, M.A.C.; Ferreira, A.C.S.; Teixeira, J.A.; Vicente, A.A. Antioxidant Potential of Two Red Seaweeds from the Brazilian Coasts. J. Agric. Food Chem. 2011, 59, 5589–5594.
  14. Santos, S.A.O.; Félix, R.; Pais, A.C.S.; Rocha, S.M.; Silvestre, A.J.D. The quest for phenolic compounds from macroalgae: A review of extraction and identification methodologies. Biomolecules 2019, 9, 847.
  15. 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.
  16. Imbs, T.I.; Zvyagintseva, T.N. Phlorotannins are Polyphenolic Metabolites of Brown Algae. Russ. J. Mar. Biol. 2018, 44, 263–273.
  17. Achkar, J.; Xian, M.; Zhao, H.; Frost, J.W. Biosynthesis of Phloroglucinol. J. Am. Chem. Soc. 2005, 127, 5332–5333.
  18. Yoon, M.; Kim, J.-S.; Um, M.Y.; Yang, H.; Kim, J.; Kim, Y.T.; Lee, C.; Kim, S.-B.; Kwon, S.; Cho, S. Extraction Optimization for Phlorotannin Recovery from the Edible Brown Seaweed Ecklonia Cava. J. Aquat. Food Prod. Technol. 2017, 26, 801–810.
  19. Mukherjee, P.K. Bioactive phytocomponents and their analysis. In Quality Control and Evaluation of Herbal Drugs; Mukherjee, P.K., Ed.; Elsevier: Amsterdam, The Netherlands, 2019; pp. 237–328.
  20. Bilal Hussain, M.; Hassan, S.; Waheed, M.; Javed, A.; Adil Farooq, M.; Tahir, A. Bioavailability and Metabolic Pathway of Phenolic Compounds. In Plant Physiological Aspects of Phenolic Compounds; IntechOpen: London, UK, 2019; pp. 1–18.
  21. Yoshie-Stark, Y.; Hsieh, Y. Distribution of flavonoids and related compounds from seaweeds in Japan. TOKYO Univ. Fish. 2003, 89, 1–6.
  22. Klejdus, B.; Lojková, L.; Plaza, M.; Šnóblová, M.; Štěrbová, D. Hyphenated technique for the extraction and determination of isoflavones in algae: Ultrasound-assisted supercritical fluid extraction followed by fast chromatography with tandem mass spectrometry. J. Chromatogr. A 2010, 1217, 7956–7965.
  23. Yoshie, Y.; Wang, W.; Petillo, D.; Suzuki, T. Distribution of catechins in Japanese seaweeds. Fish. Sci. 2000, 66, 998–1000.
  24. Güven, K.C.; Percot, A.; Sezik, E. Alkaloids in Marine Algae. Mar. Drugs 2010, 8, 269.
  25. Biris-Dorhoi, E.S.; Michiu, D.; Pop, C.R.; Rotar, A.M.; Tofana, M.; Pop, O.L.; Socaci, S.A.; Farcas, A.C. Macroalgae—A sustainable source of chemical compounds with biological activities. Nutrients 2020, 12, 3085.
  26. Flodin, C.; Helidoniotis, F.; Whitfield, F.B. Seasonal variation in bromophenol content and bromoperoxidase activity in Ulva lactuca. Phytochemistry 1999, 51, 135–138.
  27. Colon, M.; Guevara, P.; Gerwick, W.H.; Ballantine, D. 5′-hydroxyisoavrainvilleol, a new Diphenylmethane Derivative from the Tropical Green Alga Avrainvillea nigricans. J. Nat. Prod. 1987, 50, 368–374.
  28. Flodin, C.; Whitfield, F.B. 4-hydroxybenzoic acid: A likely precursor of 2,4,6-tribromophenol in Ulva lactuca. Phytochemistry 1999, 51, 249–255.
  29. Wall, M.E.; Wani, M.C.; Manikumar, G.; Taylor, H.; Hughes, T.J.; Gaetano, K.; Gerwick, W.H.; McPhail, A.T.; McPhail, D.R. Plant antimutagenic agents, 7. Structure and antimutagenic properties of cymobarbatol and 4-isocymobarbatol, new cymopols from green alga (Cymopolia barbata). J. Nat. Prod. 1989, 52, 1092–1099.
  30. Katsui, N.; Suzuki, Y.; Kitamura, S.; Irie, T. 5,6-dibromoprotocatechualdehyde and 2,3-dibromo-4,5-dihydroxybenzyl methyl ether. New dibromophenols from Rhodomela larix. Tetrahedron 1967, 23, 1185–1188.
  31. Fan, X.; Xu, N.J.; Shi, J.G. Bromophenols from the red alga Rhodomela confervoides. J. Nat. Prod. 2003, 66, 455–458.
  32. Ko, S.C.; Ding, Y.; Kim, J.; Ye, B.R.; Kim, E.A.; Jung, W.K.; Heo, S.J.; Lee, S.H. Bromophenol (5-bromo-3,4-dihydroxybenzaldehyde) isolated from red alga Polysiphonia morrowii inhibits adipogenesis by regulating expression of adipogenic transcription factors and AMP-activated protein kinase activation in 3T3-L1 adipocytes. Phyther. Res. 2019, 33, 737–744.
  33. Xu, X.-L.; Fan, X.; Song, F.-H.; Zhao, J.-L.; Han, L.-J.; Yang, Y.-C.; Shi, J.-G. Bromophenols from the brown alga Leathesia nana. J. Asian Nat. Prod. Res. 2004, 6, 217–221.
  34. Chung, H.Y.; Ma, W.C.J.; Ang, P.O.; Kim, J.S.; Chen, F. Seasonal variations of bromophenols in brown algae (Padina arboroscens, Sargassum siliquastrum, and Lobophora variegata) collected in Hong Kong. J. Agric. Food Chem. 2003, 51, 2619–2624.
  35. Liu, M.; Hansen, P.E.; Lin, X. Bromophenols in Marine Algae and Their Bioactivities. Mar. Drugs 2011, 9, 1273.
  36. Reddy, P.; Urban, S. Meroditerpenoids from the southern Australian marine brown alga Sargassum fallax. Phytochemistry 2009, 70, 250–255.
  37. Stout, E.P.; Prudhomme, J.; Le Roch, K.; Fairchild, C.R.; Franzblau, S.G.; Aalbersberg, W.; Hay, M.E.; Kubanek, J. Unusual antimalarial meroditerpenes from tropical red macroalgae. Bioorganic Med. Chem. Lett. 2010, 20, 5662–5665.
  38. Yuan, Y.V.; Westcott, N.D.; Hu, C.; Kitts, D.D. Mycosporine-like amino acid composition of the edible red alga, Palmaria palmata (dulse) harvested from the west and east coasts of Grand Manan Island, New Brunswick. Food Chem. 2009, 112, 321–328.
  39. Figueroa, F.L.; Bueno, A.; Korbee, N.; Santos, R.; Mata, L.; Schuenhoff, A. Accumulation of Mycosporine-like Amino Acids in Asparagopsis armata Grown in Tanks with Fishpond Effluents of Gilthead Sea Bream, Sparus aurata. J. World Aquac. Soc. 2008, 39, 692–699.
  40. Ferreres, F.; Lopes, G.; Gil-Izquierdo, A.; Andrade, P.B.; Sousa, C.; Mouga, T.; Valentão, P. Phlorotannin extracts from fucales characterized by HPLC-DAD-ESI-MS n: Approaches to hyaluronidase inhibitory capacity and antioxidant properties. Mar. Drugs 2012, 10, 2766.
  41. Kim, S.-Y.; Ahn, G.; Kim, H.-S.; Je, J.-G.; Kim, K.-N.; Jeon, Y.-J. Diphlorethohydroxycarmalol (DPHC) Isolated from the Brown Alga Ishige okamurae Acts on Inflammatory Myopathy as an Inhibitory Agent of TNF-α. Mar. Drugs 2020, 18, 529.
  42. Allwood, J.W.; Evans, H.; Austin, C.; McDougall, G.J. Extraction, Enrichment, and LC-MSn-Based Characterization of Phlorotannins and Related Phenolics from the Brown Seaweed, Ascophyllum nodosum. Mar. Drugs 2020, 18, 448.
  43. Gonçalves-Fernández, C.; Sineiro, J.; Moreira, R.; Gualillo, O. Extraction and characterization of phlorotannin-enriched fractions from the Atlantic seaweed Bifurcaria bifurcata and evaluation of their cytotoxic activity in murine cell line. J. Appl. Phycol. 2019, 31, 2573–2583.
  44. Soares, A.R.; Duarte, H.M.; Tinnoco, L.W.; Pereira, R.C.; Teixeira, V.L. Intraspecific variation of meroditerpenoids in the brown alga Stypopodium zonale guiding the isolation of new compounds. Rev. Bras. Farmacogn. 2015, 25, 627–633.
  45. Llewellyn, C.A.; Airs, R.L. Distribution and abundance of MAAs in 33 species of microalgae across 13 classes. Mar. Drugs 2010, 8, 1273.
  46. Urquiaga, I.; Leighton, F. Plant Polyphenol Antioxidants and Oxidative Stress. Biol. Res. 2000, 33, 55–64.
  47. Bedoux, G.; Pliego-Cortés, H.; Dufau, C.; Hardouin, K.; Boulho, R.; Freile-Pelegrín, Y.; Robledo, D.; Bourgougnon, N. Production and properties of mycosporine-like amino acids isolated from seaweeds. In Advances in Botanical Research; Bourgougnon, N., Ed.; Elsevier: Amsterdam, The Netherlands, 2020; pp. 213–245.
  48. Carreto, J.I.; Carignan, M.O. Mycosporine-Like Amino Acids: Relevant Secondary Metabolites. Chemical and Ecological Aspects. Mar. Drugs 2011, 9, 387.
  49. Gupta, S.; Abu-Ghannam, N. Recent developments in the application of seaweeds or seaweed extracts as a means for enhancing the safety and quality attributes of foods. Innov. Food Sci. Emerg. Technol. 2011, 12, 600–609.
  50. Targett, N.M.; Arnold, T.M. Predicting the effects of brown algal phlorotannins on marine herbivores in tropical and temperate oceans. J. Phycol. 1998, 34, 195–205.
  51. Le Lann, K.; Jégou, C.; Stiger-Pouvreau, V. Effect of different conditioning treatments on total phenolic content and antioxidant activities in two Sargassacean species: Comparison of the frondose Sargassum muticum (Yendo) Fensholt and the cylindrical Bifurcaria bifurcata R. Ross. Phycol. Res. 2008, 56, 238–245.
  52. Plouguerné, E.; Le Lann, K.; Connan, S.; Jechoux, G.; Deslandes, E.; Stiger-Pouvreau, V. Spatial and seasonal variation in density, reproductive status, length and phenolic content of the invasive brown macroalga Sargassum muticum (Yendo) Fensholt along the coast of Western Brittany (France). Aquat. Bot. 2006, 85, 337–344.
  53. Swanson, A.K.; Druehl, L.D. Induction, exudation and the UV protective role of kelp phlorotannins. Aquat. Bot. 2002, 73, 241–253.
  54. Sathya, R.; Kanaga, N.; Sankar, P.; Jeeva, S. Antioxidant properties of phlorotannins from brown seaweed Cystoseira trinodis (Forsskål) C. Agardh. Arab. J. Chem. 2017, 10, S2608–S2614.
  55. Rotini, A.; Belmonte, A.; Barrote, I.; Micheli, C.; Peirano, A.; Santos, R.O.; Silva, J.; Migliore, L. Effectiveness and consistency of a suite of descriptors for assessing the ecological status of seagrass meadows (Posidonia oceanica L. Delile). Estuar. Coast. Shelf Sci. 2013, 130, 252–259.
  56. Stiger, V.; Deslandes, E.; Payri, C.E. Phenolic contents of two brown algae, Turbinaria ornata and Sargassum mangarevense on Tahiti (French Polynesia): Interspecific, ontogenic and spatio-temporal variations. Bot. Mar. 2004, 47, 402–409.
  57. Migliore, L.; Rotini, A.; Randazzo, D.; Albanese, N.N.; Giallongo, A. Phenols content and 2-D electrophoresis protein pattern: A promising tool to monitor Posidonia meadows health state. BMC Ecol. 2007, 7, 1–8.
  58. Dang, T.T.; Bowyer, M.C.; Van Altena, I.A.; Scarlett, C.J. Comparison of chemical profile and antioxidant properties of the brown algae. Int. J. Food Sci. Technol. 2018, 53, 174–181.
  59. Lemesheva, V.; Tarakhovskaya, E. Physiological functions of phlorotannins. Biol. Commun. 2018, 63, 70–76.
  60. Messina, C.M.; Renda, G.; Laudicella, V.A.; Trepos, R.; Fauchon, M.; Hellio, C.; Santulli, A. From ecology to biotechnology, study of the defense strategies of algae and halophytes (from trapani saltworks, NW sicily) with a focus on antioxidants and antimicrobial properties. Int. J. Mol. Sci. 2019, 20, 881.
  61. Farvin, K.H.S.; Surendraraj, A.; Al-Ghunaim, A.; Al-Yamani, F. Chemical profile and antioxidant activities of 26 selected species of seaweeds from Kuwait coast. J. Appl. Phycol. 2019, 31, 2653–2668.
  62. Tanniou, A.; Vandanjon, L.; Incera, M.; Serrano Leon, E.; Husa, V.; Le Grand, J.; Nicolas, J.-L.; Poupart, N.; Kervarec, N.; Engelen, A.; et al. Assessment of the spatial variability of phenolic contents and associated bioactivities in the invasive alga Sargassum muticum sampled along its European range from Norway to Portugal. J. Appl. Phycol. 2013.
  63. Namvar, F.; Mohamad, R.; Baharara, J.; Zafar-Balanejad, S.; Fargahi, F.; Rahman, H.S. Antioxidant, antiproliferative, and antiangiogenesis effects of polyphenol-rich seaweed (Sargassum muticum). BioMed Res. Int. 2013, 2013.
  64. Shibata, T.; Ishimaru, K.; Kawaguchi, S.; Yoshikawa, H.; Hama, Y. Antioxidant activities of phlorotannins isolated from Japanese Laminariaceae. J. Appl. Phycol. 2008, 20, 705–711.
  65. Nwosu, F.; Morris, J.; Lund, V.A.; Stewart, D.; Ross, H.A.; McDougall, G.J. Anti-proliferative and potential anti-diabetic effects of phenolic-rich extracts from edible marine algae. Food Chem. 2011, 126, 1006–1012.
  66. Gupta, S.; Abu-Ghannam, N. Bioactive potential and possible health effects of edible brown seaweeds. Trends Food Sci. Technol. 2011, 22, 315–326.
  67. Namvar, F.; Mohamed, S.; Fard, S.G.; Behravan, J.; Mustapha, N.M.; Alitheen, N.B.M.; Othman, F. Polyphenol-rich seaweed (Eucheuma cottonii) extract suppresses breast tumour via hormone modulation and apoptosis induction. Food Chem. 2012, 130, 376–382.
  68. Iso, H. Lifestyle and cardiovascular disease in Japan. J. Atheroscler. Thromb. 2011, 18, 83–88.
  69. Kim, J.; Shin, A.; Lee, J.S.; Youn, S.; Yoo, K.Y. Dietary factors and breast cancer in Korea: An ecological study. Breast J. 2009, 15, 683–686.
  70. Mao, S.C.; Guo, Y.W. Sesquiterpenes from Chinese Red Alga Laurencia okamurai. Chin. J. Nat. Med. 2010, 8, 321–325.
  71. Ryu, B.M.; Qian, Z.J.; Kim, M.M.; Nam, K.W.; Kim, S.K. Anti-photoaging activity and inhibition of matrix metalloproteinase (MMP) by marine red alga, Corallina pilulifera methanol extract. Radiat. Phys. Chem. 2009, 78, 98–105.
  72. Guinea, M.; Franco, V.; Araujo-Bazán, L.; Rodríguez-Martín, I.; González, S. In vivo UVB-photoprotective activity of extracts from commercial marine macroalgae. Food Chem. Toxicol. 2012, 50, 1109–1117.
  73. Kumar, K.S.; Ganesan, K.; Rao, P.V.S. Antioxidant potential of solvent extracts of Kappaphycus alvarezii (Doty) Doty—An edible seaweed. Food Chem. 2008, 107, 289–295.
  74. Kang, C.; Jin, Y.B.; Lee, H.; Cha, M.; Sohn, E.T.; Moon, J.; Park, C.; Chun, S.; Jung, E.S.; Hong, J.S.; et al. Brown alga Ecklonia cava attenuates type 1 diabetes by activating AMPK and Akt signaling pathways. Food Chem. Toxicol. 2010, 48, 509–516.
  75. O’Sullivan, A.M.; O’Callaghan, Y.C.; O’Grady, M.N.; Queguineur, B.; Hanniffy, D.; Troy, D.J.; Kerry, J.P.; O’Brien, N.M. In vitro and cellular antioxidant activities of seaweed extracts prepared from five brown seaweeds harvested in spring from the west coast of Ireland. Food Chem. 2011, 126, 1064–1070.
  76. Heo, S.J.; Park, E.J.; Lee, K.W.; Jeon, Y.J. Antioxidant activities of enzymatic extracts from brown seaweeds. Bioresour. Technol. 2005, 96, 1613–1623.
  77. López, A.; Rico, M.; Rivero, A.; Suárez de Tangil, M. The effects of solvents on the phenolic contents and antioxidant activity of Stypocaulon scoparium algae extracts. Food Chem. 2011, 125, 1104–1109.
  78. Bedoux, G.; Hardouin, K.; Burlot, A.S.; Bourgougnon, N. Bioactive components from seaweeds: Cosmetic applications and future development; Elsevier: Amsterdam, The Netherlands, 2014; Volume 71, ISBN 9780124080621.
  79. Vo, T.S.; Ngo, D.H.; Kim, S.K. Marine algae as a potential pharmaceutical source for anti-allergic therapeutics. Process Biochem. 2012, 47, 386–394.
  80. Sugiura, Y.; Matsuda, K.; Yamada, Y.; Nishikawa, M.; Shioya, K.; Katsuzaki, H.; Imai, K.; Amano, H. Isolation of a new anti-allergic phlorotannin, phlorofucofuroeckol-B, from an edible brown alga, Eisenia arborea. Biosci. Biotechnol. Biochem. 2006, 70, 2807–2811.
  81. Shi, D.; Li, J.; Guo, S.; Su, H.; Fan, X. The antitumor effect of bromophenol derivatives in vitro and Leathesia nana extract in vivo. Chinese J. Oceanol. Limnol. 2009, 27, 277–282.
  82. Park, H.-J.; Kurokawa, M.; Shiraki, K.; Nakamura, N.; Choi, J.-S.; Hattori, M. Antiviral Activity of the Marine Alga Symphyocladia latiuscula against Herpes Simplex Virus (HSV-1) in Vitro and Its Therapeutic Efficacy against HSV-1 Infection in Mice. Biol. Pharm. Bull. 2005, 28, 2258–2262.
  83. Xu, N.; Fan, X.; Yan, X.; Li, X.; Niu, R.; Tseng, C.K. Antibacterial bromophenols from the marine red alga Rhodomela confervoides. Phytochemistry 2003, 62, 1221–1224.
  84. Manach, C.; Scalbert, A.; Morand, C.; Rémésy, C.; Jiménez, L. Polyphenols: Food sources and bioavailability. Am. J. Clin. Nutr. 2004, 79, 727–747.
  85. Clifford, M.N. Chlorogenic acids and other cinnamates—Nature, occurrence and dietary burden. J. Sci. Food Agric. 1999, 79, 362–372.
  86. Arts, I.C.W.; Hollman, P.C.H.; Kromhout, D. Chocolate as a source of tea flavonoids. Lancet 1999, 354, 488.
  87. O’Connell, J.E.; Fox, P.F. Significance and applications of phenolic compounds in the production and quality of milk and dairy products: A review. Int. Dairy J. 2001, 11, 103–120.
  88. Park, P.J.; Jung, W.K.; Nam, K.S.; Shahidi, F.; Kim, S.K. Purification and characterization of antioxidative peptides from protein hydrolysate of lecithin-free egg yolk. JAOCS J. Am. Oil Chem. Soc. 2001, 78, 651–656.
  89. 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.
  90. Craft, B.D.; Kerrihard, A.L.; Amarowicz, R.; Pegg, R.B. Phenol-Based Antioxidants and the In Vitro Methods Used for Their Assessment. Compr. Rev. Food Sci. Food Saf. 2012, 11, 148–173.
  91. Turck, D.; Bresson, J.; Burlingame, B.; Dean, T.; Fairweather-Tait, S.; Heinonen, M.; Hirsch-Ernst, K.I.; Mangelsdorf, I.; McArdle, H.J.; Naska, A.; et al. Safety of Ecklonia cava phlorotannins as a novel food pursuant to Regulation (EC) No 258/97. EFSA J. 2017, 15.
  92. Roy, M.C.; Anguenot, R.; Fillion, C.; Beaulieu, M.; Bérubé, J.; Richard, D. Effect of a commercially-available algal phlorotannins extract on digestive enzymes and carbohydrate absorption in vivo. Food Res. Int. 2011, 44, 3026–3029.
  93. Machu, L.; Misurcova, L.; Ambrozova, J.V.; Orsavova, J.; Mlcek, J.; Sochor, J.; Jurikova, T. Phenolic content and antioxidant capacity in algal food products. Molecules 2015, 20, 1118.
  94. Whitfield, F.B.; Helidoniotis, F.; Shaw, K.J.; Svoronos, D. Distribution of Bromophenols in Species of Marine Algae from Eastern Australia. J. Agric. Food Chem. 1999, 47, 2367–2373.
  95. Boyle, J.L.; Lindsay, R.C.; Stuiber, D.A. Bromophenol Distribution in Salmon and Selected Seafoods of Fresh- and Saltwater Origin. J. Food Sci. 1992, 57, 918–922.
  96. Whitfield, F.B.; Helidoniotis, F.; Shaw, K.J.; Svoronos, D. Distribution of Bromophenols in Australian Wild-Harvested and Cultivated Prawns (Shrimp). J. Agric. Food Chem. 1997, 45, 4398–4405.
  97. Whitfield, F.B.; Helidoniotis, F.; Svoronos, D.; Shaw, K.J.; Ford, G.L. The source of bromophenols in some species of australian ocean fish. Water Sci. Technol. 1995, 31, 113–120.
  98. Jittrepotch, N.; Ushio, H.; Ohshima, T. Effects of EDTA and a combined use of nitrite and ascorbate on lipid oxidation in cooked Japanese sardine (Sardinops melanostictus) during refrigerated storage. Food Chem. 2006, 99, 70–82.
  99. Janarthanan, M.; Senthil Kumar, M. The properties of bioactive substances obtained from seaweeds and their applications in textile industries. J. Ind. Text. 2018, 48, 361–401.
  100. Sahu, S.C.; Toxicology, B.; Assessment, S.; Nutrition, A. Food, U.S. Dual role of organosulfur compounds in foods. J. Environ. Sci. Health Part C Environ. Carcinog. Ecotoxicol. Rev. 2006, 20, 37–41.
  101. Ye, W.; Xin, J.H.; Li, P.; Lee, K.L.D.; Kwong, T.L. Durable antibacterial finish on cotton fabric by using chitosan-based polymeric core-shell particles. J. Appl. Polym. Sci. 2006, 102, 1787–1793.
  102. Wang, X.; Du, Y.; Fan, L.; Liu, H.; Hu, Y. Chitosan- metal complexes as antimicrobial agent: Synthesis, characterization and Structure-activity study. Polym. Bull. 2005, 55, 105–113.
More