Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 + 3582 word(s) 3582 2021-04-21 10:34:12 |
2 format correct Meta information modification 3582 2021-04-21 10:49:35 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Prieto Lage, M. Invasive Algae Species. Encyclopedia. Available online: https://encyclopedia.pub/entry/8862 (accessed on 09 May 2024).
Prieto Lage M. Invasive Algae Species. Encyclopedia. Available at: https://encyclopedia.pub/entry/8862. Accessed May 09, 2024.
Prieto Lage, Miguel. "Invasive Algae Species" Encyclopedia, https://encyclopedia.pub/entry/8862 (accessed May 09, 2024).
Prieto Lage, M. (2021, April 21). Invasive Algae Species. In Encyclopedia. https://encyclopedia.pub/entry/8862
Prieto Lage, Miguel. "Invasive Algae Species." Encyclopedia. Web. 21 April, 2021.
Invasive Algae Species
Edit

Invasive alien species (IAS), also known as exotic or non-native species, are plants or animals that have been introduced, intentionally or not, into regions where it is not usual to find them.

seaweeds aquaculture feed invasive macroalgae species

1. Introduction

Invasive alien species (IAS), also known as exotic or non-native species, are plants or animals that have been introduced, intentionally or not, into regions where it is not usual to find them [1][2]. This situation often leads to negative consequences for the new host ecosystem, generally related to the community biodiversity reduction, changes in the abundance of the species and in the population’s configuration across the habitats, as well as trophic displacements that can trigger other cascade effects [3]. Spanish law 42/2007, of 13 December, on Natural Heritage and Biodiversity, defines IAS as “species that are introduced and established in an ecosystem or natural habitat, which are an agent of change and a threat to native biological diversity, either by their invasive behavior, or by the risk of genetic contamination”. IAS usually present high growth and reproduction rates, the ability to prosper in different environments, the capacity to use several food sources, and the ability to tolerate a wide range of environmental conditions. All these factors, along with the lack of natural predators, make these organisms more difficult to control and allow them to succeed in colonizing new ecosystems [3][4]. In addition, these species may feed on natural species or may carry pathogens for native organisms and even humans [5]. The invasion of non-native species also entails economic cost, which have been estimated at $1.4 trillion in the last decade [6].

Among marine IAS declared in Europe, around 20–40% are macroalgae (seaweeds) [7], a term that refers to several species of multicellular and macroscopic marine algae, including different types of Chlorophyta (green), Phaeophyta (brown), and Rhodophyta (red) macroalgae. Non-native seaweeds are particularly prone to become invasive due to their high reproductive rates, the production of toxic metabolites, and their perennial status that makes them more competitive than native species [1]. Several species periodically become a major problem, causing red tides, fouling nets, clogging waterways, and changing nutrient regimes in areas near to fisheries, aquaculture systems, and desalination facilities [1][4]. In the last years, the presence of invasive macroalgae in the northwestern marine areas of Spain has become a common problem due to growing globalization, climate change, aquaculture, fisheries, and marine tourism [8]. However, their proliferation could also offer new opportunities since the recovery of the algal biomass and their novel applications in different economic sectors could increase their added value. Obtaining natural compounds with biological properties of interest for both the food and the pharmaceutical industries is one of these possible applications.

2. Possible Exploitation of the Invasive Species

The exploitation of macroalgae is a growing industry with several applications, including human food and animal feed, biorefinery, fertilizers, production of phycocolloids, and obtaining compounds with biological properties [6][9]. Several applications are briefly discussed below.

2.1. Food Industry

Macroalgae have been consumed since ancient times in many countries around the world, mainly in the Asian regions. Nevertheless, their consumption has increased in the last decades in western countries, which has been attributed to the high nutritional values of macroalgae and their health benefits [10][11]. Some of the most consumed macroalgae are nori or purple laver (Porphyra spp.), kombu (Laminaria japonica), wakame (Undaria pinnatifida), Hiziki (Hizikia fusiforme), or Irish moss (Chondrus crispus), which can be consumed in different food formats (salads, soups, snacks, pasta, etc.) [11][12]. Still, most of them are considered an innovative niche product. Macroalgae are also widely used in the food industry to produce phycocolloids (polysaccharides of high molecular weight composed mostly of simple sugars), mainly alginates, agars, and carrageenans, which are frequently used as thickeners, stabilizers, as well as for probiotics encapsulation, gels, and water-soluble films formation [6][13]. Furthermore, diverse molecules present in algae have been shown to exert several bioactivities, such as antioxidant, anti-inflammatory, antimicrobial, and antiviral effects. These bioactive compounds (mainly proteins, polyunsaturated fatty acids, carotenoids, vitamins, and minerals) may play important roles in functional foods (e.g., dairy products, desserts, pastas, oil derivatives, or supplements) with favorable outcomes on human health [14]. Other applications of algae in the food industry include their use as colorant agents and the extraction of valuable oils (such as eicosapentaenoic acid, docosahexaenoic acid, and arachidonic acid) [15].

2.2. Biofuel

The development of algal biofuels (“third-generation biofuels”) has been considered an option to reduce the use of petroleum-based fuels and avoid competition between food and energy production for arable soil, since macroalgae grow in water. These organisms do not contain lignin, thus they are good substrates for biogas production in anaerobic digesters, while fermentable carbohydrates are fit for bioethanol production. Although the production of bioenergy from macroalgae is not economically feasible nowadays, several measures have been proposed to achieve a rational production cost in the future [16]. On the other hand, microalgae are considered a more suitable source to produce biodiesel due to the greater ease of controlling the life cycle and increasing the reproduction rate [17]. Microalgae biomass can be used for electricity generation or biofuel production after the lipid extraction. It has shown 80% of the average energy content of petroleum. The lipid content is highly dependent on the microalgae species and the cultivation conditions, thus not all species will be profitable, and choosing appropriate microalga strain is crucial [18]. Some microalgae used to produce biofuel are Chlorella spp., Dunaliella salina, Haematococcus pluvialis, Spirulina platensis, Porphyridium cruentum, Microcystis aeruginosa, and Scenedesmus obliquus [19].

2.3. Therapeutic and Cosmetic Products

The use of macroalgae for therapeutic purposes has a long history, but the search for biologically active substances from these organisms is quite recent. Numerous studies have demonstrated the biological properties of macroalgae extracts and compounds, including antioxidant, anti-inflammatory [20], antithrombotic, anticoagulant and coagulant [21], antimicrobial [22], and anticancer [23]. In addition, macroalgae have been demonstrated to exert biological properties applicable to cosmetic products, such as photo-protection, anti-aging, or anti-cellulite (Table 1). Considering this range of activities, macroalgae extracts and compounds have been considered for different pharmacologic and cosmetic products [24]. Regarding cosmetics, brown and red seaweeds are usually employed. The interest of these species lies in their content in cosmeceuticals ingredients, such as phlorotannins, polysaccharides, and carotenoid pigments [25]. These compounds are incorporated into cosmetics due to their bioactivities, their capacity to improve organoleptic properties, and their capacity to stabilize and preserve the products [26].

Table 1. Properties and applications of extracts and compounds isolated from algae in the cosmetic field.

2.4. Fertilizer and Animal Feed

Currently, the negative environmental impacts of synthetic fertilizers have been identified. Thus, the use of organic fertilizers, including macroalgae, has been proposed as a suitable alternative to reduce the impact on the environment [45][46]. In fact, macroalgae have been used since ancient times as fertilizers, and several beneficial effects have been described, such as enhancement of crops growth and yield, increased resistance against abiotic and biotic stresses, or nutrient intake [46][47][48]. The biostimulant effects of macroalgae have been attributed to diverse biological compounds such as plant hormones, phlorotannins, and oligosaccharides [48].

Regarding animal feed, macroalgae have been employed for this purpose since ancient times as feed but also as nutritious supplements [49]. Several studies have evaluated the positive effects of macroalgae-enriched food, both for terrestrial animals [50] and specially in aquaculture animals [51][52][53][54].

3. Main Invasive Species of Northwest Spain and Their Bioactive Compounds

According to the Spanish Catalogue of IAS of Algae [55], there are 14 species of invasive seaweeds in Spain which can be divided into: (i) red species: Acrothamnion preissii, Asparagopsis armata, Asparagopsis taxiformis, Grateloupia turuturu, Lophocladia lallemandii, and Womersleyella setacea; (ii) brown species: Gracilaria vermiculophylla, Sargassum muticum, Stypopodium schimper, and Undaria pinnatifida; and (iii) green species: Caulerpa taxifolia, Codium fragile, and Caulerpa racemosa. In addition, there are also invasive diatoms, such as the Didymosphenia geminata, also known as rock snot or didymo (Table 2). However, it should be noted that this catalogue is a dynamic instrument subjected to continuous changes and updating. Most of these invasive species are originally from the Indo-Pacific Ocean (Western Australia, New Zealand, and Japan), and it is thought that they have been introduced into the Spanish coasts through the Suez Canal. Maritime traffic, ballast water, fishing nets, trade of oysters, aquaculture, and fouling are considered the main routes of dispersion [8][56][57][58].

Table 2. Invasive algae species in Spain: taxonomy, origin, geographical distribution, and principal uses.

The use of some algae (e.g., Caulerpa racemosa) as ornamental species in aquariums has also contributed to their proliferation [59][60]. Among these species, only five are considered invasive (*) or potentially invasive (**) in Galicia (northwest Spain): Asparagopsis armata**, Codium fragile subs. tomentosoides*, Grateloupia turuturu**, Sargassum muticum*, and Gracilaria vermiculophylla*. Galician waters also feature the presence of two other exotic invasive species, though they do not appear in the regulation of Real Decreto (RD) 1628/201; these are Gymnodinium catenatum and Bonamia ostreae [61].

For many years, non-native species of algae have been considered threats, thus a series of methods to eradicate them from non-endemic areas have been developed and optimized. However, the marine biomass, including invasive macroalgae, is currently the focus of several industries, such as pharmaceutical, food, cosmetic, and biotechnological industries, due their biological activities, e.g., antioxidant, antimicrobial, anti-inflammatory, anticancer. The aim of these industries is to revalorize invasive macroalgae as a source of extracts and compounds with industrial interest [8]. Although many studies have evaluated the biological properties of various extracts of A. armata, C. fragile, G. turuturu, S. muticum, and G. verniculophylla, in some cases, the bioactive compounds responsible for this activity have not yet been identified. In the following paragraphs, the current knowledge about target compounds for industrial applications and the bioactive compounds identified in the macroalgae species considered invasive in Galicia are compiled. They are also summarized in Table 3.

Table 3. Main compounds and bioactive compounds reported for the invasive macroalgae in northwest Spain.

3.1. Polysaccharides

In the case of A. armata, the polysaccharides derived from sulfated galactans have shown strong antiviral effects against human immunodeficiency virus (HIV), inhibiting its reproduction [71]. A study confirmed the inhibition of herpes simplex virus type 1 by different extracts of numerous red algae, including A. armata. Although the authors did not identify the compounds involved in the activity, the good results of the water extract were attributed to water-soluble polysaccharides [72]. Mannitol has been also identified in the ethanolic extract of A. armata, in a concentration of 34.70 mg/100 g of dry macroalgae [73].

In the case of C. fragile, several bioactivities have been attributed to its sulfated polysaccharides (SPs). The administration of this type of compounds reduced the oxidative damage associated with diabetes mellitus and obesity in several animal models without any cytotoxic effect [74][75]. Recently, a study stated that SPs from C. fragile scavenge effectively freed radicals in vitro and suppressed the oxidative damage caused by H2O2 in Vero cell cultures and in zebrafish [76]. It has also been reported that SPs from C. fragile increased the coagulation time of human blood in a dose-dependent manner according to the methods activated partial thromboplastin time (APTT) [77][78], thrombin time (TT), and prothrombin time (PT) [78]. SPs from C. fragile inhibited HeLa cells proliferation [79] by stimulating tumor necrosis factor (TNF)-related apoptosis-inducing ligand, a promising anticancer target [80].

Finally, these compounds also show immune-stimulating properties in both in vitro and in vivo models. Sulfated galactan obtained from C. fragile stimulated murine macrophages RAW264.7 cell line, increasing the levels of nitric oxide and both pro-inflammatory and anti-inflammatory cytokines, which are fundamental for the host immune response [81][82][83]. In head kidney cells, SPs had a stimulatory effect on immune genes, including interleukin (IL)-1β, IL-8, TNF-α, interferon (IFN)-γ, and lysozyme [84]. Immuno-stimulant properties have been also observed in human peripheral blood dendritic cells and T cells, which were activated by SPs. This suggests that these compounds could be candidates for products aimed to enhance human immune system [85].

S. muticum is a source of several valuable polysaccharides, such as fucoidans, alginate, guluronic and mannuronic acids, laminarin, and their derivatives [86]. Alginate obtained from S. muticum has been demonstrated to possess anticancer properties, stimulating cell death in A549 cells (epithelial lung adenocarcinoma), PSN1 cells (pancreatic adenocarcinoma), HCT- 116 cells (colon carcinoma), and T98G cells (glioblastoma) [87].

Finally, G. vermiculophylla and G. turuturu are being used in the phycocolloid industry for obtaining agar and carrageenan, respectively, turning them into valuable matrixes [88][89]. Recently, polysaccharide extracts from G. turuturu have shown antimicrobial properties against Escherichia coli and Staphylococcus aureus [90].

3.2. Lipids

Starting with A. armata, it has been reported that these macroalgae contain some sterols such as cholesta-5,25-diene-3,24-diol, (3β,24S)-form [91], palmitic and stearic fatty acids, and cholestanol [73]. Recently, different crude extracts and fractions of this species were demonstrated to present antibacterial and antifouling properties. In the crude extract and most active fractions, several compounds were identified, including hexadecanoic, dodecanoic, octadecanoic, and tetradecanoic acids, which may be involved in this activity [92][93].

Regarding C. fragile, clerosterol (a derivative of cholesterol) was found in several extracts. This compound shows antioxidant properties, since it attenuated UVB-induced oxidative damage in human immortalized keratinocyte HaCaT cells and BALB/c mice models, reducing lipid and protein oxidation [94]. In addition, clerosterol stimulated apoptosis in A2058 human melanoma cells [95] and modulated several apoptotic factors in human leukemia cells [96]. Recently, a study observed that C. fragile displayed neuroprotective effects on neuroblastoma cell line SH-SY5Y. In the most bioactive fractions, several lipid compounds, among others, were identified. Although more research is needed, the authors considered that lipids are involved in the neuroprotective effect [97].

G. vermiculophylla contains high quantity of cholesterol (473.2 mg/kg dry weight), cholesterol derivatives, long-chain aliphatic alcohols, and monoglycerides, including 1-tetradecanol, 1-hexadecanol, 1-octadecanol, 1-eicosanol, and 1-docosanol [67]. Other lipids of great interest for nutraceutical and biotechnological industries include phospholipids, glycolipids, and eicosapentaenoic acid, present in high levels in this alga [69]. For example, three sphingolipids (gracilarioside, and gracilamides A and B) isolated from G. vermiculophylla (accepted name of G. asiatica) showed moderate cytotoxic effects against human A375-S2 melanoma cell line [98].

3.3. Proteins

To our knowledge, only G. vermiculophylla presents bioactive compounds of protein nature. This alga can absorb UV-A and UV-B radiations and decrease free radicals-induced effects, resulting from its high content in mycosporine-like amino acids [99].

3.4. Pigments

Siphonaxanthin from C. fragile has shown anticancer properties, stimulating the apoptosis of A549 lung cancer cells and modulating apoptotic factors in human leukemia cells [95][96]. Moreover, the anti-angiogenic effect of siphonaxanthin has been described in human umbilical vein endothelial cells as well as in a rat aortic ring angiogenic model [100], which suggests that this biomolecule could be an alternative to prevent pro-angiogenic diseases such as cancer. In addition, this alga also contains β-carotene [66].

In recent years, fucoxanthin has received a great deal of interest from the scientific community and industry due to the many beneficial health properties attributed to it, including anti-inflammatory [101]. Fucoxanthin extracted from S. muticum inhibited the lipopolysaccharide-induced nitric oxide production in RAW 264.7 macrophages and inhibited the expression of pro-inflammatory cytokines [102][103].

At industrial scale, G. turuturu is also used to produce R-phycoerythrin, a pink-purple pigment soluble in water present in large quantities, which presents diverse biological properties and potential industrial applications [89][104].

3.5. Vitamins

Different vitamins have been identified in the selected macroalgae, except in A. armata. In C. fragile, high levels of tocopherols have been reported (1617.6 µg/g lipid), including α, β, γ, and δ tocopherol and γ-tocotrienol [66]. G. vermiculophylla showed a considerable α-tocopherol content (28.4 μg/g of extract) [105]. Regarding G. turuturu, a chemical analysis revealed the presence of α-tocopherol and phytonadione (vitamin K1) [70]. Finally, S. muticum contains high amounts of α- and γ- tocopherol, 218 and 20.8 μg/g of extract, respectively [105].

3.6. Phenolic Compounds

Phenolic content has been evaluated in several species, although not all the studies have identified the target compounds. In the case of A. armata, phenolic content was determined by the Folin–Ciocalteu spectrophotometry method, which showed that it represented 1.13 ± 0.05% of dry weight [106]. Different extracts of C. fragile also contain phenolic compounds, mainly flavonoids and, to a lesser extent, tannins. These compounds showed a correlation with the antioxidant activity of the macroalgae [65]. The previous study of Farvin and Jacobsen (2013) identified several phenolic acids in both G. vermiculophylla aqueous extracts (gallic, protocatechuic, hydroxybenzoic, vanillic, syringic, and salicylic acids) and ethanolic extracts (gallic, protocatechuic, and gentisic acids). In correspondence with its content in phenolic compounds, a high antioxidant capacity has been demonstrated for these macroalgae according to the 2,2-Diphenyl-1-picrylhydrazyl (DPPH) and the ferric antioxidant power (FRAP) methods. In addition, G. vermiculophylla extracts inhibited lipid peroxidation [105]. Finally, some authors have reported the presence of phenolic compounds in S. muticum, including (ordered from highest to lowest concentration): hydroxybenzoic and gallic acids, p-hydroxybenzaldehyde, vanillic acid, 3,4-dihydroxybenzaldehyde and protocatechuic, ferulic, p-coumaric, caffeic, syringic, and chlorogenic acids [107]. Several bioactivities of S. muticum, such as antioxidant, antimicrobial, anticancer, or anti-inflammatory, have been attributed to the presence of phenolic compounds with high antioxidant capacity, particularly to phlorotannins (e.g., phloroglucinol, diphlorethol, bifuhalol), which are exclusively found in marine seaweed [68][108][109][110][111].

3.7. Other Minor Compounds

The invasive species A. armata presents high levels of halogenated secondary metabolites with recognized antibiotic activity [112]. They act as chemical defense against grazers and epibiota [113] and may be suitable for a wide range of applications [114][115]. For instance, the major metabolites bromoform and dibromoacetic acid, along with dibromochloromethane, bromochloroacetic acid, and dibromoacrylic acid, have shown high antifouling potential [62][63][64]. They can decrease the density of six bacteria strains on the algae surface: two marine (Vibrio harveyii and V. alginolyticus) and four biomedical strains (Pseudomonas aeruginosa, Staphylococcus aureus, Staphylococcus epidermis, and Escherichia coli) [116]. Recently, several brominated compounds, such as tribromomethanol, were found in the crude extract and fractions of A. armata, which showed antimicrobial antifouling potential [92][93].

A serine protease extracted from C. fragile was demonstrated to exert in vitro and in vivo anticoagulant and fibrinogenolytic activity [117]. Finally, it was found that G. turuturu contains squalene, which was reported to exert several beneficial activities [70].

References

  1. Máximo, P.; Ferreira, L.M.; Branco, P.; Lima, P.; Lourenço, A. Secondary metabolites and biological activity of invasive macroalgae of southern Europe. Mar. Drugs 2018, 16, 265.
  2. Shackleton, R.T.; Shackleton, C.M.; Kull, C.A. The role of invasive alien species in shaping local livelihoods and human well-being: A review. J. Environ. Manag. 2019, 229, 145–157.
  3. Pyšek, P.; Hulme, P.E.; Simberloff, D.; Bacher, S.; Blackburn, T.M.; Carlton, J.T.; Dawson, W.; Essl, F.; Foxcroft, L.C.; Genovesi, P.; et al. Scientists’ warning on invasive alien species. Biol. Rev. 2020, 95, 1511–1534.
  4. Otero, M.; Cebrian, E.; Francour, P.; Galil, B.; Savini, D. Monitoring Marine Marine Protected in Mediterranean Invasive Species Areas (MPAs)—A Strategy and Practical Guide for Managers; IUCN: Malaga, Spain, 2013.
  5. Commision European. Invasive Alien Species of Union Concern; Commision European: Luxembourg, 2020.
  6. Milledge, J.J.; Nielsen, B.V.; Bailey, D. High-value products from macroalgae: The potential uses of the invasive brown seaweed, Sargassum muticum. Rev. Environ. Sci. Biotechnol. 2015, 15, 67–88.
  7. Davoult, D.; Surget, G.; Stiger-Pouvreau, V.; Noisette, F.; Riera, P.; Stagnol, D.; Androuin, T.; Poupart, N. Multiple effects of a Gracilaria vermiculophylla invasion on estuarine mudflat functioning and diversity. Mar. Environ. Res. 2017, 131, 227–235.
  8. Pinteus, S.; Lemos, M.F.L.; Alves, C.; Neugebauer, A.; Silva, J.; Thomas, O.P.; Botana, L.M.; Gaspar, H.; Pedrosa, R. Marine invasive macroalgae: Turning a real threat into a major opportunity—the biotechnological potential of Sargassum muticum and Asparagopsis armata. Algal Res. 2018, 34, 217–234.
  9. Machmudah, S.; Diono, W.; Kanda, H.; Goto, M. Supercritical fluids extraction of valuable compounds from algae: Future perspectives and challenges. Eng. J. 2018, 22, 13–30.
  10. Buschmann, A.H.; Camus, C.; Infante, J.; Neori, A.; Israel, Á.; Hernández-González, M.C.; Pereda, S.V.; Gomez-Pinchetti, J.L.; Golberg, A.; Tadmor-Shalev, N.; et al. Seaweed production: Overview of the global state of exploitation, farming and emerging research activity. Eur. J. Phycol. 2017, 52, 391–406.
  11. Fernández-Segovia, I.; Lerma-García, M.J.; Fuentes, A.; Barat, J.M. Characterization of Spanish powdered seaweeds: Composition, antioxidant capacity and technological properties. Food Res. Int. 2018, 111, 212–219.
  12. Gómez-Zavaglia, A.; Prieto Lage, M.A.; Jiménez-López, C.; Mejuto, J.C.; Simal-Gándara, J. The Potential of Seaweeds as a Source of Functional Ingredients of Prebiotic and Antioxidant Value. Antioxidants 2019, 8, 406.
  13. Gomez, L.P.; Alvarez, C.; Zhao, M.; Tiwari, U.; Curtin, J.; Garcia-Vaquero, M.; Tiwari, B.K. Innovative processing strategies and technologies to obtain hydrocolloids from macroalgae for food applications. Carbohydr. Polym. 2020, 248, 116784.
  14. Camacho, F.; Macedo, A.; Malcata, F. Potential industrial applications and commercialization of microalgae in the functional food and feed industries: A short review. Mar. Drugs 2019, 17, 312.
  15. Matos, Â.P. The Impact of Microalgae in Food Science and Technology. JAOCS J. Am. Oil Chem. Soc. 2017, 94, 1333–1350.
  16. Soleymani, M.; Rosentrater, K.A. Techno-economic analysis of biofuel production from macroalgae (Seaweed). Bioengineering 2017, 4, 92.
  17. Culaba, A.B.; Ubando, A.T.; Ching, P.M.L.; Chen, W.H.; Chang, J.S. Biofuel from microalgae: Sustainable pathways. Sustainability 2020, 12, 9.
  18. Milano, J.; Ong, H.C.; Masjuki, H.H.; Chong, W.T.; Lam, M.K.; Loh, P.K.; Vellayan, V. Microalgae biofuels as an alternative to fossil fuel for power generation. Renew. Sustain. Energy Rev. 2016, 58, 180–197.
  19. Shuba, E.S.; Kifle, D. Microalgae to biofuels: ‘Promising’ alternative and renewable energy, review. Renew. Sustain. Energy Rev. 2018, 81, 743–755.
  20. Blunt, J.W.; Carroll, A.R.; Copp, B.R.; Davis, R.A.; Keyzers, R.A.; Prinsep, M.R. Marine natural products. Nat. Prod. Rep. 2018, 35, 8–53.
  21. Holdt, S.L.; Kraan, S. Bioactive compounds in seaweed: Functional food applications and legislation. J. Appl. Phycol. 2011, 23, 543–597.
  22. Silva, A.; Silva, S.A.; Carpena, M.; Garcia-Oliveira, P.; Gullón, P.; Barroso, M.F.; Prieto, M.A.; Simal-Gandara, J. Macroalgae as a source of valuable antimicrobial compounds: Extraction and applications. Antibiotics 2020, 9, 642.
  23. Gopeechund, A.; Bhagooli, R.; Neergheen, V.S.; Bolton, J.J.; Bahorun, T. Anticancer activities of marine macroalgae: Status and future perspectives. In Biodiversity and Biomedicine; Ozturk, M., Egamberdieva, D., Pešić, M., Eds.; Academic Press: Cambridge, MA, USA, 2020; pp. 257–275. ISBN 9780128195413.
  24. Cikoš, A.-M.; Jerković, I.; Molnar, M.; Šubarić, D.; Jokić, S. New trends for macroalgal natural products applications. Nat. Prod. Res. 2019, 1–12.
  25. Kim, S.K. Marine Cosmeceuticals: Trends and Prospects, 1st ed.; Kim, S.K., Ed.; Tayor & Francis Group: Boca Ratón, FL, USA, 2012; ISBN 9781439860281.
  26. Bedoux, G.; Hardouin, K.; Burlot, A.S.; Bourgougnon, N. Bioactive components from seaweeds: Cosmetic applications and future development. In Advances in Botanical Research; Academic Press: Cambridge, MA, USA, 2014.
  27. 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.
  28. Nizard, C.; Friguet, B.; Moreau, M.; Bulteau, A.-L.; Saunois , A. Use of Phaeodactylum Algae Extract as Cosmetic Agent Promoting the Proteasome Activity of Skin Cells and Cosmetic Composition Comprising Same. U.S. Patent 7,220,417, 22 May 2007.
  29. Hwang, E.; Park, S.Y.; Sun, Z.; Shin, H.S.; Lee, D.G.; Yi, T.H. The Protective Effects of Fucosterol Against Skin Damage in UVB-Irradiated Human Dermal Fibroblasts. Mar. Biotechnol. 2014, 16, 361–370.
  30. Joe, M.J.; Kim, S.N.; Choi, H.Y.; Shin, W.S.; Park, G.M.; Kang, D.W.; Yong, K.K. The inhibitory effects of eckol and dieckol from Ecklonia stolonifera on the expression of matrix metalloproteinase-1 in human dermal fibroblasts. Biol. Pharm. Bull. 2006, 29, 1735–1739.
  31. Le Lann, K.; Surget, G.; Couteau, C.; Coiffard, L.; Cérantola, S.; Gaillard, F.; Larnicol, M.; Zubia, M.; Guérard, F.; Poupart, N.; et al. Sunscreen, antioxidant, and bactericide capacities of phlorotannins from the brown macroalga Halidrys siliquosa. J. Appl. Phycol. 2016, 28, 3547–3559.
  32. Sanjeewa, K.K.A.; Kim, E.A.; Son, K.T.; Jeon, Y.J. Bioactive properties and potentials cosmeceutical applications of phlorotannins isolated from brown seaweeds: A review. J. Photochem. Photobiol. B Biol. 2016, 162, 100–105.
  33. 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.
  34. Heo, S.-J.; Jeon, Y.-J. Protective effect of fucoxanthin isolated from Sargassum siliquastrum on UV-B induced cell damage. J. Photochem. Photobiol. B Biol. 2009, 95, 101–107.
  35. Fernando, I.P.S.; Dias, M.K.H.M.; Madusanka, D.M.D.; Han, E.J.; Kim, M.J.; Jeon, Y.J.; Ahn, G. Fucoidan refined by Sargassum confusum indicate protective effects suppressing photo-oxidative stress and skin barrier perturbation in UVB-induced human keratinocytes. Int. J. Biol. Macromol. 2020, 164, 149–161.
  36. 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.
  37. 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.
  38. Choi, J.S.; Moon, W.S.; Choi, J.N.; Do, K.H.; Moon, S.H.; Cho, K.K.; Han, C.J.; Choi, I.S. Effects of seaweed Laminaria japonica extracts on skin moisturizing activity in vivo. J. Cosmet Sci. 2013, 64, 193–205.
  39. 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.
  40. Kim, S.-K.; Babitha, S.; Kim, E.-K. Effect of Marine Cosmeceuticals on the Pigmentation of Skin. In Marine Cosmeceuticals; CRC Press: Boca Ratón, FL, USA, 2011; pp. 63–65.
  41. Wang, H.D.; Chen, C.; Huynh, P.; Chang, J. Exploring the potential of using algae in cosmetics. Bioresour. Technol. 2015, 184, 355–362.
  42. Sahin, S.C. The potential of Arthrospira platensis extract as a tyrosinase inhibitor for pharmaceutical or cosmetic applications. S. Afr. J. Bot. 2018, 119, 236–243.
  43. 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.
  44. Bak, S.S.; Ahn, B.N.; Kim, J.A.; Shin, S.H.; Kim, J.C.; Kim, M.K.; Sung, Y.K.; Kim, S.K. Ecklonia cava promotes hair growth. Clin. Exp. Dermatol. 2013, 38, 904–910.
  45. Atzori, G.; Nissim, W.G.; Rodolfi, L.; Niccolai, A.; Biondi, N.; Mancuso, S.; Tredici, M.R. Algae and Bioguano as promising source of organic fertilizers. J. Appl. Phycol. 2020, 32, 3971–3981.
  46. Akila, V.; Manikandan, A.; Sahaya Sukeetha, D.; Balakrishnan, S.; Ayyasamy, P.M.; Rajakumar, S. Biogas and biofertilizer production of marine macroalgae: An effective anaerobic digestion of Ulva sp. Biocatal. Agric. Biotechnol. 2019, 18, 101035.
  47. Hashem, H.A.; Mansour, H.A.; El-Khawas, S.A.; Hassanein, R.A. The potentiality of marine macro-algae as bio-fertilizers to improve the productivity and salt stress tolerance of canola (Brassica napus L.) plants. Agronomy 2019, 9, 146.
  48. Stirk, W.A.; Rengasamy, K.R.R.; Kulkarni, M.G.; Staden, J. Plant Biostimulants from Seaweed. In The Chemical Biology of Plant Biostimulants; Geelen, D., Lin, X., Eds.; Wiley: Chichester, UK, 2020; pp. 31–55.
  49. Leandro, A.; Pereira, L.; Gonçalves, A.M.M. Diverse applications of marine macroalgae. Mar. Drugs 2020, 18, 17.
  50. Maia, M.R.G.; Fonseca, A.J.M.; Cortez, P.P.; Cabrita, A.R.J. In vitro evaluation of macroalgae as unconventional ingredients in ruminant animal feeds. Algal Res. 2019, 40, 101481.
  51. Bansemer, M.S.; Qin, J.G.; Harris, J.O.; Duong, D.N.; Hoang, T.H.; Howarth, G.S.; Stone, D.A.J. Growth and feed utilisation of greenlip abalone (Haliotis laevigata) fed nutrient enriched macroalgae. Aquaculture 2016, 452, 62–68.
  52. Sáez, M.I.; Vizcaíno, A.; Galafat, A.; Anguís, V.; Fernández-Díaz, C.; Balebona, M.C.; Alarcón, F.J.; Martínez, T.F. Assessment of long-term effects of the macroalgae Ulva ohnoi included in diets on Senegalese sole (Solea senegalensis) fillet quality. Algal Res. 2020, 47, 101885.
  53. Valente, L.M.P.; Gouveia, A.; Rema, P.; Matos, J.; Gomes, E.F.; Pinto, I.S. Evaluation of three seaweeds Gracilaria bursa-pastoris, Ulva rigida and Gracilaria cornea as dietary ingredients in European sea bass (Dicentrarchus labrax) juveniles. Aquaculture 2006, 252, 85–91.
  54. Passos, R.; Correia, A.P.; Ferreira, I.; Pires, P.; Pires, D.; Gomes, E.; do Carmo, B.; Santos, P.; Simões, M.; Afonso, C.; et al. Effect on health status and pathogen resistance of gilthead seabream (Sparus aurata) fed with diets supplemented with Gracilaria gracilis. Aquaculture 2021, 531, 735888.
  55. Boletín Oficial del Estado. Real Decreto 630/2013, de 2 de Agosto, por el que se Regula el Catálogo Español de Especies Exóticas Invasoras; Boletín Oficial del Estado: Madrid, Spain, 2013.
  56. Mulas, M.; Bertocci, I. Devil’s tongue weed (Grateloupia turuturu Yamada) in northern Portugal: Passenger or driver of change in native biodiversity? Mar. Environ. Res. 2016, 118, 1–9.
  57. Altamirano, M.; Muñoz, A.R.; de la Rosa, J.; Barrajón-Mínguez, A.; Barrajón-Domenech, A.; Moreno-Robledo, C.; del Arroyo, M.C. The invasive species Asparagopsis taxiformis (Bonnemaisoniales, Rhodophyta) on andalusian coasts (Southern Spain): Reproductive stages, new records and invaded communities. Acta Botánica Malacit. 2008, 23, 5–15.
  58. Bellissimo, G.; Galfo, F.; Nicastro, A.; Costantini, R.; Castriota, L. First record of the invasive green alga Codium fragile ssp. fragile (Chlorophyta, Bryopsidales) in Abruzzi waters, central Adriatic sea. Acta Adriat. 2018, 59, 207–212.
  59. Gennaro, P.; Piazzi, L. The indirect role of nutrients in enhancing the invasion of Caulerpa racemosa var cylindracea. Biol. Invasions 2014, 16, 1709–1717.
  60. Ornano, L.; Sanna, C.; Serafini, M.; Bianco, A.; Donno, Y.; Ballero, M. Phytochemical study of Caulerpa racemosa (Forsk.) J. Agarth, an invading alga in the habitat of La Maddalena Archipelago. Nat. Prod. Res. 2014, 28, 1795–1799.
  61. Capdevila-Argüelles, L.; Zilletti, B.; Suárez Álvarez, V.Á. Plan. Extratéxico galego de Xestión das Especies Exóticas Invasoras e Para o Desenvolvemento Dun Sistema Esandarizado de Análise de Riscos Para as Especies Exóticas en Galicia; Xunta de Galicia: Santiago, Chile; Galicia, Spain, 2012.
  62. Marshall, R.A.; Hamilton, J.T.G.; Dring, M.J.; Harper, D.B. Do vesicle cells of the red alga Asparagopsis (Falkenbergia stage) play a role in bromocarbon production? Chemosphere 2003, 52, 471–475.
  63. McConnell, O.; Fenical, W. Halogen chemistry of the red alga asparagopsis. Phytochemistry 1977, 16, 367–374.
  64. Woolard, F.X.; Moore, R.E.; Roller, P.P. Halogenated acetic and acrylic acids from the red alga asparagopsis taxiformis. Phytochemistry 1979, 18, 617–620.
  65. Kolsi, R.B.A.; Salah, H.B.; Hamza, A.; El feki, A.; Allouche, N.; El feki, L.; Belguith, K. Characterization and evaluating of antioxidant and antihypertensive properties of green alga (Codium fragile) from the coast of Sfax. J. Pharmacogn. Phytochem. 2017, 6, 186–191.
  66. Ortiz, J.; Uquiche, E.; Robert, P.; Romero, N.; Quitral, V.; Llantén, C. Functional and nutritional value of the Chilean seaweeds Codium fragile, Gracilaria chilensis and Macrocystis pyrifera. Eur. J. Lipid Sci. Technol. 2009, 320–327.
  67. Santos, S.A.O.; Vilela, C.; Freire, C.S.R.; Abreu, M.H.; Rocha, S.M.; Silvestre, A.J.D. Chlorophyta and Rhodophyta macroalgae: A source of health promoting phytochemicals. Food Chem. 2015, 183, 122–128.
  68. 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.
  69. Kendel, M.; Barnathan, G.; Fleurence, J.; Rabesaotra, V.; Wielgosz-Collin, G. Non-methylene interrupted and hydroxy fatty acids in polar lipids of the alga Grateloupia turuturu over the four seasons. Lipids 2013, 48, 535–545.
  70. Kendel, M.; Couzinet-Mossion, A.; Viau, M.; Fleurence, J.; Barnathan, G.; Wielgosz-Collin, G. Seasonal composition of lipids, fatty acids, and sterols in the edible red alga Grateloupia turuturu. J. Appl. Phycol. 2013, 25, 425–432.
  71. Haslin, C.; Lahaye, M.; Pellegrini, M.; Chermann, J.C. In Vitro Anti-HIV Activity of Sulfated Cell-Wall Polysaccharides from Gametic, Carposporic and Tetrasporic Stages of the Mediterraean Red Alga Asparagopsis armata. Planta Med. 2001, 67, 301–305.
  72. Bouhlal, R.; Riadi, H.; Bourgougnon, N. Antiviral activity of the extracts of Rhodophyceae from Morocco. Afr. J. Biotechnol. 2010, 9, 7968–7975.
  73. Andrade, P.B.; Barbosa, M.; Pedro, R.; Lopes, G.; Vinholes, J.; Mouga, T.; Valentão, P. Valuable compounds in macroalgae extracts. Food Chem. 2013, 138, 1819–1828.
  74. Kolsi, R.B.A.; Fakhfakh, J.; Sassi, S.; Elleuch, M.; Gargouri, L. Physico-chemical characterization and beneficial effects of seaweed sulfated polysaccharide against oxydatif and cellular damages caused by alloxan in diabetic rats. Int. J. Biol. Macromol. 2018, 117, 407–417.
  75. Kolsi, R.B.A.; Jardak, N.; Hajkacem, F.; Chaaben, R.; Jribi, I.; El Feki, A.; Rebai, T.; Jamoussi, K.; Fki, L.; Belghith, H.; et al. Anti-obesity effect and protection of liver-kidney functions by Codium fragile sulphated polysaccharide on high fat diet induced obese rats. Int. J. Biol. Macromol. 2017, 102, 119–129.
  76. Wang, L.; Oh, J.Y.; Je, J.G.; Jayawardena, T.U.; Kim, Y.S.; Ko, J.Y.; Fu, X.; Jeon, Y.J. Protective effects of sulfated polysaccharides isolated from the enzymatic digest of Codium fragile against hydrogen peroxide-induced oxidative stress in in vitro and in vivo models. Algal Res. 2020, 48, 101891.
  77. Athukorala, Y.; Lee, K.W.; Kim, S.K.; Jeon, Y.J. Anticoagulant activity of marine green and brown algae collected from Jeju Island in Korea. Bioresour. Technol. 2007, 98, 1711–1716.
  78. Ciancia, M.; Quintana, I.; Vizcargüénaga, M.I.; Kasulin, L.; de Dios, A.; Estevez, J.M.; Cerezo, A.S. Polysaccharides from the green seaweeds Codium fragile and C. vermilara with controversial effects on hemostasis. Int. J. Biol. Macromol. 2007, 41, 641–649.
  79. Surayot, U.; You, S.G. Structural effects of sulfated polysaccharides from Codium fragile on NK cell activation and cytotoxicity. Int. J. Biol. Macromol. 2017, 98, 117–124.
  80. Park, S.H.; Kim, J.L.; Jeong, S.; Kim, B.R.; Na, Y.J.; Jo, M.J.; Yun, H.K.; Jeong, Y.A.; Kim, D.Y.; Kim, B.G.; et al. Codium fragile F2 sensitize colorectal cancer cells to TRAIL-induced apoptosis via c-FLIP ubiquitination. Biochem. Biophys. Res. Commun. 2019, 508, 1–8.
  81. Lee, J.B.; Ohta, Y.; Hayashi, K.; Hayashi, T. Immunostimulating effects of a sulfated galactan from Codium fragile. Carbohydr. Res. 2010, 345, 1452–1454.
  82. Shi, Q.; Wang, A.; Lu, Z.; Qin, C.; Hu, J.; Yin, J. Overview on the antiviral activities and mechanisms of marine polysaccharides from seaweeds. Carbohydr. Res. 2017, 453–454, 1–9.
  83. Fernández, P.V.; Arata, P.X.; Ciancia, M. Polysaccharides from Codium Species; Elsevier: Amsterdam, The Netherlands, 2014; Volume 71, ISBN 9780124080621.
  84. Yang, Y.; Park, J.; You, S.G.; Hong, S. Immuno-stimulatory effects of sulfated polysaccharides isolated from Codium fragile in olive flounder, Paralichthys olivaceus. In Fish Shellfish Immunology; Elsevier Inc.: Alpharetta, GA, USA, 2019; Volume 87, pp. 609–614.
  85. Zhang, W.; Hwang, J.; Park, H.; Lim, S.; Go, S. Human Peripheral Blood Dendritic Cell and T Cell Activation by Codium fragile Polysaccharide. Mar. Drugs 2020, 18, 535.
  86. Sánchez-Camargo, P.; Montero, L.; Stiger-pouvreau, V.; Tanniou, A.; Cifuentes, A.; Herrero, M.; Ibáñez, E. Considerations on the use of enzyme-assisted extraction in combination with pressurized liquids to recover bioactive compounds from algae. Food Chem. 2016, 192, 67–74.
  87. Flórez-Fernández, N.; Domínguez, H.; Torres, M.D. A green approach for alginate extraction from Sargassum muticum brown seaweed using ultrasound-assisted technique. Int. J. Biol. Macromol. 2019, 124, 451–459.
  88. Martínez-Lüscher, J.; Holmer, M. Potential effects of the invasive species Gracilaria vermiculophylla on Zostera marina metabolism and survival. Mar. Environ. Res. 2010, 69, 345–349.
  89. Pereira, L. Edible Seaweeds of the World; CRC Press: Boca Ratón, FL, USA, 2016.
  90. Cardoso, I.; Cotas, J.; Rodrigues, A.; Ferreira, D.; Osório, N.; Pereira, L. Extraction and analysis of compounds with antibacterial potential from the red alga Grateloupia turuturu. J. Mar. Sci. Eng. 2019, 7, 220.
  91. Sheu, J.; Huang, S.; Duh, C. Cytotoxic Oxygenated Desmosterols of the Red Alga Galaxaura marginata. J. Nat. Prod. 1996, 59, 23–26.
  92. Pinteus, S.; Lemos, M.F.L.; Alves, C.; Silva, J.; Pedrosa, R. The marine invasive seaweeds Asparagopsis armata and Sargassum muticum as targets for greener antifouling solutions. Sci. Total Environ. 2021, 750, 141372.
  93. Pinteus, S.; Lemos, M.F.L.; Simões, M.; Alves, C.; Silva, J.; Gaspar, H.; Martins, A.; Rodrigues, A.; Pedrosa, R. Marine invasive species for high-value products’ exploration—Unveiling the antimicrobial potential of Asparagopsis armata against human pathogens. Algal Res. 2020, 52, 102091.
  94. Lee, C.; Park, G.H.; Ahn, E.M.; Kim, B.A.; Park, C.I.; Jang, J.H. Protective effect of Codium fragile against UVB-induced pro-inflammatory and oxidative damages in HaCaT cells and BALB/c mice. Fitoterapia 2013, 86, 54–63.
  95. Kim, A.D.; Lee, Y.; Kang, S.H.; Kim, G.Y.; Kim, H.S.; Hyun, J.W. Cytotoxic effect of clerosterol isolated from Codium fragile on A2058 human melanoma cells. Mar. Drugs 2013, 11, 418–430.
  96. Ganesan, P.; Noda, K.; Manabe, Y.; Ohkubo, T.; Tanaka, Y.; Maoka, T.; Sugawara, T.; Hirata, T. Siphonaxanthin, a marine carotenoid from green algae, effectively induces apoptosis in human leukemia (HL-60) cells. Biochim. Biophys. Acta Gen. Subj. 2011, 1810, 497–503.
  97. Silva, J.; Martins, A.; Alves, C.; Pinteus, S.; Gaspar, H.; Alfonso, A.; Pedrosa, R. Natural Approaches for Neurological Disorders-The Neuroprotective Potential of Codium tomentosum. Molecules 2020, 25, 5478.
  98. Sun, Y.; Xu, Y.; Liu, K.; Hua, H.; Zhu, H.; Pei, Y. Gracilarioside and gracilamides from the Red alga Gracilaria asiatica. J. Nat. Prod. 2006, 69, 1488–1491.
  99. Barceló-villalobos, M.; Figueroa, F.L.; Korbee, N. 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.
  100. Ganesan, P.; Matsubara, K.; Ohkubo, T.; Tanaka, Y.; Noda, K.; Sugawara, T.; Hirata, T. Anti-angiogenic effect of siphonaxanthin from green alga, Codium fragile. Phytomedicine 2010, 17, 1140–1144.
  101. Lourenço-Lopes, C.; Garcia-Oliveira, P.; Carpena, M.; Fraga-Corral, M.; Jimenez-Lopez, C.; Pereira, A.G.; Prieto, M.A.; Simal-Gandara, J. Scientific approaches on extraction, purification and stability for the commercialization of fucoxanthin recovered from brown algae. Foods 2020, 9, 1113.
  102. Heo, S.J.; Yoon, W.J.; Kim, K.N.; Ahn, G.N.; Kang, S.M.; Kang, D.H.; Affan, A.; Oh, C.; Jung, W.K.; Jeon, Y.J. Evaluation of anti-inflammatory effect of fucoxanthin isolated from brown algae in lipopolysaccharide-stimulated RAW 264.7 macrophages. Food Chem. Toxicol. 2010, 48, 2045–2051.
  103. Yang, E.-J.; Moon, J.-Y.; Kim, S.S.; Yang, K.-W.; Lee, W.J.; Lee, N.H.; Hyun, C.-G. Jeju seaweeds suppress lipopolysaccharide-stimulated proinflammatory response in RAW 264. 7 murine macrophages. Asian Pac. J. Trop. Biomed. 2014, 4, 529–537.
  104. Le Guillard, C.; Dumay, J.; Donnay-Moreno, C.; Bruzac, S.; Ragon, J.Y.; Fleurence, J.; Bergé, J.P. Ultrasound-assisted extraction of R-phycoerythrin from Grateloupia turuturu with and without enzyme addition. Algal Res. 2015, 12, 522–528.
  105. 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.
  106. Zubia, M.; Fabre, M.; Deslandes, E.; Shannon, C. Antioxidant and cytotoxic activities of some red algae (Rhodophyta) from Brittany coasts (France). Botenica Mar. 2009, 52, 268–277.
  107. Klejdus, B.; Plaza, M.; Snóblová, M.; Lojková, L. Development of new efficient method for isolation of phenolics from sea algae prior to their rapid resolution liquid chromatographic—tandem mass spectrometric determination. J. Pharm. Biomed. Anal. 2017, 135, 87–96.
  108. Balboa, E.M.; Luisa, M.; Nogueira, D.R.; González-lópez, N.; Conde, E.; Moure, A.; Pilar, M. Potential of antioxidant extracts produced by aqueous processing of renewable resources for the formulation of cosmetics. Ind. Crop. Prod. 2014, 58, 104–110.
  109. 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.
  110. Casas, M.P.; Conde, E.; Domínguez, H.; Moure, A. Ecofriendly extraction of bioactive fractions from Sargassum muticum. Process Biochem. 2019, 79, 166–173.
  111. Pérez-Larrán, P.; Torres, M.D.; Flórez-Fernández, N.; Balboa, E.M.; Moure, A.; Domínguez, H. Green technologies for cascade extraction of Sargassum muticum bioactives. J. Appl. Phycol. 2019, 31, 2481–2495.
  112. Mata, L.; Silva, J.; Schuenhoff, A.; Santos, R. The effects of light and temperature on the photosynthesis of the Asparagopsis armata tetrasporophyte (Falkenbergia rufolanosa), cultivated in tanks. Aquaculture 2006, 252, 12–19.
  113. Jacinto, M.S.C.; Monteiro, H.R.; Lemos, M.F.L. Impact of the invasive macroalgae Asparagopsis armata on coastal environments: An ecotoxicological assessment. Curr. Opin. Biotechnol. 2013, 24S, S75.
  114. De Nys, R.; Steinberg, P.D.; Willemsen, P.; Dworjanyn, S.A.; Gabelish, C.L.; King, R.J. Broad spectrum effects of secondary metabolites from the red alga delisea pulchra in antifouling assays. Biofouling 1995, 8, 259–271.
  115. Schuenhoff, A.; Mata, L.; Santos, R. The tetrasporophyte of Asparagopsis armata as a novel seaweed biofilter. Aquaculture 2006, 252, 3–11.
  116. Paul, N.; Nys, R. De Chemical defence against bacteria in the red alga Asparagopsis armata: Linking structure with function. Mar. Ecol. Prog. Ser. 2006, 306, 87–101.
  117. Choi, J.H.; Sapkota, K.; Park, S.E.; Kim, S.; Kim, S.J. Thrombolytic, anticoagulant and antiplatelet activities of codiase, a bi-functional fibrinolytic enzyme from Codium fragile. Biochimie 2013, 95, 1266–1277.
More
Information
Contributor MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register :
View Times: 686
Revisions: 2 times (View History)
Update Date: 21 Apr 2021
1000/1000