Open access peer-reviewed chapter

Seaweeds as Source of New Bioactive Prototypes

Written By

Caio Cesar Richter Nogueira and Valéria Laneuville Teixeira

Submitted: 21 September 2015 Reviewed: 07 March 2016 Published: 29 June 2016

DOI: 10.5772/62913

From the Edited Volume

Algae - Organisms for Imminent Biotechnology

Edited by Nooruddin Thajuddin and Dharumadurai Dhanasekaran

Chapter metrics overview

2,484 Chapter Downloads

View Full Metrics

Abstract

Living organisms endowed with natural benefits have been used for millions of years in the medical practice. Seaweeds have been widely used around the world for the production of agar and food; however, the pharmaceutical industry has drawn attention to the activities of these natural products. In this chapter, we present some bioactive metabolites of the three phyla of seaweed (green, brown, and red algae) along with their potential for drug development.

Keywords

  • Seaweeds
  • bioactive compounds
  • natural products
  • drug development
  • bioactivities

1. Introduction

The use of natural resources for medicinal purposes in the treatment and prevention of diseases is one of the oldest practices of mankind. The earliest historical report describing the use of natural derivatives was written and found in Nagpur, India, and is approximately 5000 years old. These records comprise 12 recipes for drug preparation and refer to more than 250 plants [1]. Another historical example is the book written by Emperor Shen Nung circa 2500 BC. This Chinese book describes the use of more than 365 parts of medicinal plants; among these are camphor, the great yellow gentian, ginseng, jimson weed, cinnamon bark, and ephedra [2]. The Ebers Papyrus is one of the oldest and most important medical treatises known in the world. Written in ancient Egypt, it is dated at around 1550 BC and contains more than 700 species of plants and drugs used in therapy, such as pomegranate, castor oil, aloe, senna, garlic, onions, figs, willow, coriander, juniper, and common centaury [3].

The first initiative in the search of natural products of marine origin with pharmacological potential began at a conference in Rhode Island, USA, under the name of “Drugs from the Sea” in 1967. Since this important date, researchers from around the world pledged in search of primary and secondary metabolites of various marine organisms. In 1980, the University of Utah in the USA discovered a toxin derived from cone snail that was able to block a voltage‐gated calcium channel. Based on the initial data, a peptide was synthesized and developed by Elan Corporation. The FDA authorized the sale in 2004 of the first drug derived from a marine natural product under the trade name Prialt for the treatment of chronic pain in spinal cord injuries. In 2007, the second drug derived from marine organisms was developed. The isoquinoline derived from the sea squirt Ecteinascidia turbinate was approved by the European Union for the treatment of soft tissue sarcoma with the name of Trabectedin / Yondelis [4]. Even today, researchers from around the world are working to isolate, identify, and test promising natural products derived from marine organisms to develop new drugs.

Seaweed is used extensively for development in the industries of cosmetics, fuel production, agar production, and it also serves as animal and human food. However, due to advances in the isolation and structural elucidation of its primary and secondary metabolites, the pharmaceutical industry is turning its attention toward algae. In this section, we describe the natural products derived from seaweed, which have potential for drug development.

Advertisement

2. Marine seaweeds as a source of new bioactive prototypes

2.1. Green seaweeds as a source of new bioactive prototype

The primary metabolites in green seaweeds are more exploited for the development of drugs than the secondary metabolites. Molecules such as peptides, glycolipids, and sulfated polysaccharides have shown interesting results and are in an advanced phase in the drug test. As an example, the depsipeptide Kahalalide F (KF) isolated from the mollusk Elysia rufescens that feeds on green seaweeds of the genus Bryopsis. For isolation of the peptide, the ethanol extract of animals was chromatographed on a silica flash column, from which the peptide mixture was eluted with EtOAc/MeOH (1:l). A new column using high‐performance liquid chromatography (HPLC) on C18 reversed‐phase was performed, obtaining the isolation of KF. When measuring their biological activities, KF showed IC50 values of 2.5, 0:25, and <1.0 µg/mL against A‐549, HT‐29, and LOVO cells, respectively. Also observed was interesting antiviral activity against Herpes simplex virus type 2 (HSV‐2) and antifungal against Aspergillus oryzae, Penicillium chrysogenum (also known as P. notatum), Trichophyton mentagrophytes, Saccharomyces cerevisiae, and Candida albicans [5]. In 1996, five new depsipeptides (Kahalalide A‐E) were isolated from Elysia rufescens, which feeds on green seaweed Bryopsis sp. [6]. Due to the results obtained previously, the target for the anticancer peptide, KF, has been studied in cultured cells. During the experiment, it was observed that the cells become swollen, due to the formation of large vacuoles, which appeared to be the consequence of changes in lysosomal membranes. Thus, lysosomes are a target for KF action [7]. However, over the years, other mechanisms of action have been proposed. In 2000, KF has also shown activity against human prostate cancer xenografts in animal models in vivo [8]. To facilitate the advancement of studies of anticancer activity and toxicity of KF, its synthesis was described [9]. In 2001, a stable parenteral formulation of KF was developed, to be used in early clinical trials [10]. In 2002, preclinical toxicity studies of KF using single‐ and multiple‐dose schedules were done in rats [11]. In 2005, a study was initiated with the objective of determining the maximum tolerated dose, profile of adverse events, and dose‐limiting toxicity of KF in patients with androgen refractory prostate cancer. The study concluded that the peptide can be given safely as a one‐hour i.v. infusion during five days at a dose of 560 µg/m2 per day once every three weeks [12]. Also in the Phase I clinical stage, another group found that the maximum tolerated dose was 800 µg/m2 to patients with advanced solid tumors [13]. Recently, a group evaluated the effect of demographics and pathophysiologically relevant factors on KF pharmacokinetic parameters, however, no clinically relevant covariates were identified [14].

Activity‐directed isolation of the n‐hexane and dichloromethane fractions from the Capsosiphon fulvescens resulted in obtaining three glycolipids (Capsofulvesins A‐C) of pharmacological interest. These compounds exhibited IC50 values of 53.13 ± 2.83, 51.38 ± 0.90, and 82.54 ± 0.88 µM when measuring AChE inhibitory activity, respectively, and IC50 values of >132.28, 114.75 ± 4.13, and 185.55 ± 6.95 µM in BChE assay [15]. A screening for Aldose reductase inhibitors using the ethanol extract of 22 algae was held in South Korea. The green seaweed Capsosiphon fulvescens had one of the best results and the fractionation of its extract also resulted in the isolation of Capsofulvesins. Capsofulvesin A and B showed potential rat lens aldose reductase inhibitory activity with the IC50 values of 52.53 and 101.92, respectively [16].

Also in primary metabolism, green seaweeds produce sulfated polysaccharides, which are said to ulvans. These molecules have shown interesting immunomodulatory and anticoagulant activities. For example, the sulfated polysaccharides of green seaweed Enteromorpha prolifera were used to determine their in vitro and in vivo immunomodulatory activities. In vitro, fractions rich in sulfated polysaccharides increase nitric oxide production and cytokine (TNF‐α, IL‐6, IL‐10, and COX‐2) release in Raw 264.7 cells. In vivo, the sulfated polysaccharides increase Con A‐induced splenocyte proliferation and IFN‐γ and IL‐2 secretions [17]. Anticoagulant activity of Sulfated polysaccharides of green seaweed is also disclosed in literature [18].

Among the secondary metabolites, green seaweeds synthesize mainly sterols, alkaloids, and prenylated bromohydroquinones. Historically, the ether extract from green seaweed Cymopolia barbata has shown antibiotic and antifungal activities, but no specific compounds were isolated. In 1976, the fractionation was carried out in the same sample, resulting in the isolation of the first seven bromohydroquinones prenylated [19]. In 1987, another eight bromohydroquinones were isolated in the same seaweed, starting the study of biological activities of these substances [20]. A study in north coast of Puerto Rico was conducted with the green seaweed Cymopolia barbata leading to the isolation of two bromohydroquinones guided by antimutagenic assay. These compounds were active in inhibiting 2AN mutagenicity toward Salmonella typhimurium at doses of 300, 150, and 75 µg/plate [21]. Other prenylated bromohydroquinones have been described over the years [22]. In 2012, the green seaweed C. barbata was again collected from Jamaica at Fairy Hill Beach and its extract prepared in dichloromethane:methanol (1:1). The fractionation of the extract resulted in the isolation of two prenylated bromohydroquinones. The anticancer activity of both products was carried out using CCD18 Co, HT29, HepG, and MCF‐7 cells. The compound 1 obtained IC50 values of 55.65 ± 3.28 and 19.82 ± 0.46 to CCD18 Co and HT29, respectively. Compound 2 did not prove active in all tested cells. Furthermore, the ability of compounds to inhibit the enzyme cytochrome P450 also was evaluated. Compounds 1 and 2 showed an IC50 value of 0.93 ± 0.26 and 0.39 ± 0.05 μM, respectively [23].

Alkaloids may be defined as a compound that has nitrogen atom(s) in a cyclic ring. In marine algae, these substances can be classified into: a) Phenylethylamine alkaloids; b) Indole alkaloids; or c) Other alkaloids [24]. In green seaweeds, the indole alkaloids are the main natural products isolated. Caulerpin (3) was the first alkaloid isolated from Caulerpa genus [25] and its isolation from the substance has shown amazing results. In India, the methanolic from the seaweed Caulerpa racemosa was fractionated using column of silica gel (60–120 mesh) and was eluted successively with various percentages of solvent mixtures containing petroleum ether, chloroform, and methanol. The fractions eluted with chloroform‐petroleum ether (1: 1) resulted in the isolation of caulerpin, which was used to evaluate the anticorrosion activity using polarization, impedance, and atomic force microscopy assays. A protective layer on the steel surface was observed by electrochemical impedance spectroscopy when treated with caulerpin, demonstrating an anticorrosion effect [26].

In Brazil, caulerpin was used for investigation of their cytotoxicity on Vero cells and antiviral activity against Herpes simplex virus type 1 (HSV‐1) KOS strain. Caulerpin demonstrated a selectivity index better than the reference drug Acyclovir, with CC50 of 1167 µM and EC50 of 1.29 µM values. In addition, its mechanism of action was studied on the virus replication cycle. Caulerpin seems to inhibit the alpha and beta phase of replication of HSV‐1 virus [27]. Recently, the synthesis of Caulerpin and its analogues has been proposed along its assessment of anti‐tuberculosis activity. All compounds exhibited activity against bacillus Mycobacterium tuberculosis strain H37Rv. However, Caulerpin demonstrated the best IC50 value compared to their analogues and reference drug Rifampin [28].

Steroids are triterpenic compounds having a tetracyclic system; its A, B, and C rings have six carbons while ring D has five carbons. The vast majority of green seaweeds synthesize sterols of 28 and 29 carbons, as an example, with one of the first compounds isolated from Codium fragile [29]. Over the years, several sterols were isolated from green seaweeds and have shown interesting biological activities, however, the anticancer activity is the most explored for this type of metabolite. The green seaweed Tydemania expeditionis were collected from the Yellow Sea in Weihai, Shandong Province of China. Its partition prepared in EtOAc was subjected to silica gel (200–300 mesh) and eluted with cyclohexane‐acetic ether in various proportions. Refractionation using Sephadex LH‐20 column resulted in the isolation of four sterols, which have been used to evaluate the anticancer activity in human prostate cancer cells lines (DU145, PC3, and LNCaP). Compound 4 exhibited inhibitory activity against the prostate cancer cells DU145, PC3, and LNCaP with IC50 values of 12.38 ± 2.47, 2.14 ± 0.33, and 1.38 ± 0.07 μM, respectively. Compounds 5 showed IC50 values of 31.27 ± 1.50, 40.59 ± 3:10, 19.80 ± 3.84 μM. It was noted by researchers that the presence of hydroxyl at C‐3 increased the cytotoxic activity of sterols, however, the presence of the hydroxyl in C‐24 diminished activity. To investigate if the inhibitory activities against prostate cancer cells were due to inhibition of androgen receptor, the binding affinity of sterols was evaluated. Competitive binding assay showed that compound 5 exhibited significant affinity to the androgen receptor with an IC50 value of 7.19 ± 0.45 μM, while the compound 4 was inactive [30].

The partition prepared in EtOAc of the green seaweed Codium iyengarii, was collected from Karachi cost of Arabian, was subjected to fractionation using silica gel and hexane, chloroform, and methanol at binary mixture or pure, which resulted in the isolation of four sterols. The compound 6 showed the best IC50 values when tested against Corynebacterium diptheriae, Klebsiella pneumonia, Snigella dysentri, Staphylococcus aureus, Pseudomonas aeruginosa, Salmonella typhi, and Streptococcus pyogenes microorganisms [31].

Other products found less in green algae also exhibit biological activities described; for example, bromophenols (BPs) found in Avrainvillea genre presented inhibitory activity for HMG‐CoA reductase enzyme [32].

Figure 1.

Bioactive compounds isolated from the green seaweeds.

2.2. Brown seaweeds as a source of new bioactive prototypes

Among the primary metabolites of brown seaweeds, molecules as glycolipids and sulfated polysaccharides are used in the search for bioactivities to develop new drugs.

The brown seaweed Sargassum muticum was collected in France and its extract prepared using organic solvent chloroform. The chloroform extract was subjected to fractionation using column vacuum chromatography and high pressure liquid chromatography; this resulted in the isolation of galactoglycerolipids with antibacterial activity against Shewanella putrefaciens and Polaribacter irgensii, and antifungical activity against Pleurochrysis roscoffensis, Exanthemachrysis gayraliae, Cylindrotheca closterium, Navicula jeffreyii, Halosphaeriopsis mediosetigera, Asteromyces cruciatus, Lulworthia uniseptata, and Monodictys pelagica [33].

The brown seaweed Lobophora variegata was collected in Mexico and its extract prepared using dichloromethane‐methanol (7:3). The extract was dissolved in methanol‐water (9:1) and subjected to partitioning using hexane, chloroform, ethyl acetate, and n‐butanol. The chloroform fraction was subjected to column chromatography on Sephadex LH20, eluted with hexane‐chloroform‐methanol (3:2:1), resulting in isolation of three sulfoquinovosyldiacylglycerides. The mixture of sulfoquinovosyldiacylglycerides showed activity against Entamoeba histolytica, Trichomonas vaginalis, and Giardia intestinalis with value of IC50 of 3.9 ± 0.03 μg/mL, 8.0 ± 0.42 μg/mL, and 20.9 ± 0.89 μg/mL, respectively [34]. Other activities, such as anticancer and inhibitor of DNA polymerase, have also been described for glycolipids [35, 36].

An experiment performed with fucoidans from the Laminaria saccharina was evaluated for its biological activities. Fucoidans from the L. saccharina showed the inhibition of neutrophil extravasation into peritoneal cavity in an acute peritonitis rat model at a dose of about 4 mg/kg. Anticoagulant activity was measured as the activated partial thromboplastin time related to the heparin standard. Fucoidans the L. saccharina showed an APTT value of 33.0 ± 2 U/mg [37]. In the subsequent article, the mixture fucoidans was fractionated by ion‐exchange chromatography that produced two differing fractions. The first fraction consisted of sulfated fucomannoglucuronan and the second fraction consisted mainly of sulfated fucans. The sulfated fucan showed an increased anticoagulant activity with APTT values of 29.2 ± 1.6 [38]. A study in Phase I clinical with fucans from the Laminaria japonica investigated orally administered effects on hemostatic parameters in healthy volunteers [39].

Among the secondary metabolites, brown seaweeds synthesize different types of terpenes and phenolic compounds. Among the diterpenes, the brown seaweeds synthesize secondary metabolites with different carbon frameworks including dolabellane, dolastane, prenylated guaiane diterpenes, and meroditerpenes skeletons of interest in drug development.

The brown seaweed, Dictyota pfaffii, was collected in Rocas Atoll reef, Brazil, and its extract prepared using the mixture of organic solvents dichloromethane/methanol (7:3). This extract was subjected to silica gel column chromatography eluted with hexane, dichloromethane, ethyl acetate, and methanol at pure or binary mixture. The fraction eluted with dichloromethane pure and dichloromethane/ethyl acetate (9:1) resulting in compound 7, which was recrystallized from n‐hexane and the fraction eluted with dichloromethane/ethyl acetate (6:4) was purified using silica gel column chromatography resulting in compound 8. Compound 9 was obtained by addition of hydroxyl groups by chemical reaction. The cytotoxicity on Vero cells and HSV‐1 antiviral activity of the compounds was evaluated. Compounds 79 demonstrate CC50 values of 185 ± 5.0, 189 ± 1.2, 184 ± 3.4 µM, and TCID50 values of 89 ± 4.5, 87 ± 3.9, 81 ± 4.1%, respectively. Compared to the drug acyclovir, the compounds demonstrated an effective antiviral activity; however, they also showed high cytotoxicity [40]. The mechanism of action of the compound on HSV‐1 replication cycle was studied and compound 9 inhibits the initial events in HSV‐1 replication and decreases the levels of some early proteins of HSV‐1, such as UL‐8, RL‐1, UL‐12, UL‐30, and UL‐9 [41]. Due to the presence of an interesting antiviral activity, the substance Dolabelladienetriol (9) was used to inhibit human immunodeficiency virus (HIV) RT enzyme, with an IC50 value of 16.5 ± 4.3 µM. This same compound was also used to assess their ability to inhibit HIV‐1 replication in peripheral blood mononuclear cell and macrophages. Compound 9 showed an EC50 value of 8.4 µM in PBMC and 1.85 µM in macrophages. Knowing that HIV entry into cells may be mediated by various coreceptor, dolabelladienetriol was used to determine its inhibition of HIV replication at different strains. When used at a concentration of 25 µM, the compound was able to inhibit more than 80% of the replication of all strains of HIV. Tests indicate that the compound 9 can inhibit HIV‐1 replication at a posttranscriptional step [42]. Subsequent studies demonstrated that compound 9 blocked the synthesis and integration of HIV‐1 provirus and acts as a Non‐Nucleoside Reverse Transcriptase Inhibitors (NNRTIs) [43]. In order to develop an antiviral drug, compound 9 was evaluated for their toxicity in vivo. Mice deaths were not observed at any dose during the ten‐day period. Significant changes were not observed in the concentrations of urea nitrogen, creatinine, alanine aminotransferase, uric acid, and total protein [44]. In 2014, three news dolabellane-type diterpenoids named dolabelladienols A–C were isolated from the brown seaweed Dictyoya pffafi. These compounds have also been shown to inhibit HIV‐1 replication in MT‐2 cells [45]. The compound Dolabelladienetriol has also proven to be effective in the inhibition of Leishmania amazonensis replication in peripheral blood mononuclear cells (IC50 = 43.9 µM), even in the presence of factors that exacerbate parasite growth, such as IL‐10, TGF‐b, and HIV‐1 coinfection [46].

The brown seaweed Canistrocarpus cervicornis was collected in Rio de Janeiro, Brazil, and its extract prepared using the organic solvent dichloromethane. This extract was subjected to column chromatography resulting in the isolation of two dolastane diterpenes. Both diterpenes have been shown to inhibit the activity of organ crude homogenates and purified Na+K+-ATPase of the kidney and brain. Compound 10 showed the best results [47]. The other two dolastane diterpenes were isolated from brown seaweed C. cervicornis and evaluated their cytotoxicity in Vero cells and antiviral activity (HSV‐1). The compounds demonstrated a CC50 value of 1423 and 706, and the percentage of inhibition of viral replication of 90% and 99% when used 50 µM, respectively. The position of the double bond and the hydroxyl group seems to interfere with the cytotoxicity and antiviral activity of the substances [48]. From the same seaweed, other dalastane diterpene was isolated from the Brazilian coast and showed anticoagulant and antiplatelet effects [49].

The extract prepared in dichloromethane/methanol (1:1) of brown seaweed Dictyota menstrualis collected on the Brazilian coast was used to evaluate its antiviral activity (HIV‐1). Through isolation guided activity, it was possible to isolate two diterpenes prenylated guaianes. When evaluating the antiviral activity of compounds 19 and 20 by the p24 antigen dosage in the supernatant of PM‐1 cells, it was possible to obtain an EC50 value of 40 and 70 µM, respectively. Both compounds showed cytotoxicity with values above 200 µM. Seeking to understand the mechanism of action of the substances led to other tests being performed. When treated with 100 µM of the two substances, the synthesis of viral DNA was inhibited. Furthermore, compounds 11 and 12 show IC50 values of 10 and 35 µM when evaluating their ability to inhibit the enzyme reverse transcriptase [50]. The results obtained in the previous study investigated the mechanism of action of two diterpenes on the enzyme reverse transcriptase. The kinetic analyses of the HIV‐1 RT demonstrate that both prenylated guaianes have similar mechanisms of inhibition of RNA‐dependent DNA‐polymerase activity, with the compound 11 being more effective in inhibiting [51]. Subsequently, the compound 11 has also shown antiviral activity against herpes simplex virus type 1 with CC50 value of 1000 ± 83 µM and EC50 of 1.60 ± 0.08 [52].

The brown seaweed Padina pavonia collected from the Red Sea coast in Hurghada, Egypt, was subjected to extraction using methanol 80%, where this extract was partitioned using n‐hexane. The n‐hexane fraction was chromatographed resulting in the isolation of two xenicane diterpenes, which were used in the assessment of anticancer activity. Compounds 13 and 14 showed IC50 values of 13.2 µg/mL and 18.4 µg/mL in H460 cells, respectively, and IC50 values greater than 20 µg/mL in HepG2 cells [53].

The extract prepared in dichloromethane of brown seaweed Stypopodium zonale was collected on the coast of Tenerife, Spain and fractionated by flash chromatography. The fraction eluted with hexane‐ethyl acetate (8:2) resulted in three meroditerpenes. Compound 15 exhibits cytotoxic activity with IC50 values of <2.5 µg/mL in HT‐29 cells and IC50 value of 2.5 µg/mL in H‐116 and A549 cells [54]. Meroditerpenes isolated from brown seaweed Taonia atomaria also feature an interesting anticancer activity [55]. Meroditerpenes isolated from Stypopodium zonale collected on the Brazilian coast showed antiviral activity [56].

Brown seaweeds also synthesize products such as phlorotannins and sesquiterpenes, which can be used in the development of drugs. The brown seaweed Eisenia bicyclis was bought in Japan and its extract prepared using the organic solvent Methanol. The extract was partitioned by column system using Diaion HP‐2MG. After successive column chromatographic and reversed‐phase HPLC, it was possible to obtain the isolation of phlorotannins. Compound 16 exhibits inhibitory activity on glycation and α‐amylase enzymes [57], the angiotensin‐converting enzyme I inhibitory [58] and inhibits the protein tyrosine phosphatase 1B and α‐glucosidase [59].

The brown seaweed Dictyopteris divaricata collected at the coast of Qingdao, China, was subjected to extraction using ethanol where this extract was partitioned using ethyl acetate. The EtOAc fraction was chromatographed resulting in the isolation of four sesquiterpenes that were used in the evaluation of anticancer activity. These compounds have shown a moderate anticancer activity, with IC50 values above 10 µ against several human cancer cell lines including lung adenocarcinoma (A549), stomach cancer (BGC‐823), breast cancer (MCF‐7), hepatocellular carcinoma (Bel7402), and colon cancer (HCT‐8) cell lines [60]. Six other sesquiterpenes showed a similar IC50 value against several human cancer cell lines [61].

Figure 2.

Bioactive compounds isolated from the brown seaweeds.

2.3. Red seaweeds as a source of new bioactive prototypes

Among the primary metabolites of red seaweeds, some molecules, such as glycolipids and sulfated polysaccharides, are being used in the search for bioactivities in order to develop new drugs.

Activity‐guided isolation of red seaweed Palmaria palmata resulted in the isolation of ten polar lipids, among them, two sulfoquinovosyldiacylglycerides (SQDGs). The bioactive compounds 25 and 26 were demonstrated nitricoxide inhibitory activity in macrophage RAW264.7 cells with IC50 values of 36.5 and 11.0 μM. Moreover, the compound 26 also has been shown to inhibit the production of nitric oxide synthase in a dose‐dependent manner [62]. Other activities, such as antiviral (HSV‐1 and HSV‐2) and anticancer, have also been described for glycolipids [63, 64].

Antiviral activity has been a major focus in the study of biological activities of polysaccharides of red algae, because its polysaccharides have shown a low cytotoxicity and high efficiency [65]. The sulfated polysaccharides of red seaweeds Sphaerococcus coronopifolius and Boergeseniella thyoides collected on the coast of Morocco were used in the investigation of antiviral activity against HSV‐1 and HIV‐1. Sulfated polysaccharides were capable of inhibiting the HSV‐1 on Vero cells with values of EC50 of 4.1 and 17.2 µg/mL, respectively. After investigation of the mechanism of action of these substances, sulfated polysaccharides appear to inhibit viral adsorption step of HSV‐1. The polysaccharides of S. coronopifolius prevents HIV‐induced syncytium formation at the lowest concentration tested (12.5 µg/mL), however, the polysaccharides of B. thyoides did not demonstrate the same efficiency [66]. Some studies involving the use of carrageenan are in development; for example, a Phase II trial study in the USA has the objective of developing a vaginal gel for reducing the rate of human papilloma virus (HPV) infection [67].

Among the secondary metabolism, red seaweeds synthesize substances of different chemical classes such as terpenes, phenols, and acetogenins. Additionally, species of Rhodophyta are skilled in the incorporation of chlorine and bromine atoms.

The monoterpenes are substances with ten carbons formed by two isoprene units, and can be cyclic or aliphatic (acyclic) [68]. Halogenated monoterpenes are found in genres Plocamium, Portieria, and Ochtodes [69]. In general, it is believed that in the marine environment, halogenated monoterpenes serve as chemical defense in response to stressor agents, especially herbivores. The production and storage of these metabolites must be related to the survival of algae in the marine environment, and therefore may be potential prototypes for important pharmacological activities [70]. An example of the activity of red algae monoterpenes was described by Chilean researchers from studies with hexane extract of Plocamium cartilagineum collected on the coast of Antarctica. The hexane extract was subjected to column chromatography and fractions purified by HPLC, resulting in the isolation of four cyclic halogenated monoterpenes. One of the products 27 showed an intense insecticidal activity against Heliothis virescens larvae and moderate activity against Spodoptera frugiperda. In the same study, the other monoterpene 28 showed antibacterial activity, resulting in a zone of inhibition of 19.35 mm when used to Gram‐negative bacterial strain Porphyromonas gingivalis, a major organism responsible for chronic periodontitis [71]. Other activities of monoterpenes belonging to Rhodophyta have been described, for example, to inhibit DNA methyltransferase activity [72], as anticancer [73, 74], and antifungal activities [75].

Sesquiterpenes are natural products with 15 carbons, formed from three isoprene units. The sesquiterpenes are the class of natural products more produced by phylum Rhodophyta, especially by species of Laurencia. These secondary metabolites can have various types of carbon skeletons, such as bisabolane, brasilane, chamigrane, cuparane, eudesmane, laurane, and snyderane.

Halogenated sesquiterpenes with bisabolane skeleton are mainly synthesized by species Laurencia aldingensis and Laurencia catarinensis. Recently, the in vitro production of bisabolanes was developed through genetic engineering techniques and Saccharomyces cerevisiae metabolism [76]. This fact aroused the interest of biotechnologists, in view of the possibility of large‐scale production. The halogenated bisabolene sesquiterpene (29) showed anthelmintic activity against parasitant stage (L4) of Nippostrongilus brasiliensis [77].

The chamigrane sesquiterpenes exhibit a spiro ring attached to a five‐carbon ring. The chamigranes can be divided into: those that contain an epoxide between carbons 5 and 10 [78] and others that do not [79]. The total synthesis of chamigrane sesquiterpene elatol was described, which further stimulated the search for their biological activities [80], taking into account that the elatol is the most potent known natural product with antifouling activity [81, 82]. The elatol (30), isolated from Laurencia dendroidea, collected on the northern Rio de Janeiro coast, was tested against the Y strain of Trypanosoma cruzi. Elatol showed a dose‐dependent effect against the epimastigote, trypomastigote, and amastigote forms, with IC50 values of 45.4, 1.38, and 1.01 µM, respectively [83]. The elatol was also obtained from Laurencia microcladia and evaluated for their anticancer activity in vitro and in vivo. This substance showed of IC50 1.1 μM in L929 cells and IC50 of 10.1 µM in B16F10 cells. It also caused a delay in the transition from the G1/S phase of the cell cycle and induces apoptosis [84].

The cuparanes sesquiterpenes are rarely described in red seaweeds. The great majority of isolates is formed by an aromatic ring attached to a ring structure of five carbon atoms and may or may not have double bonds in its interior. In 1996, the synthesis of Cuparene and Cuparenol metabolites has been described from β‐cyclogeraniol [85]. Currently, these substances are marketed, which brings a lot of interest as prototypes with biological activities. Two cuparane sesquiterpenes (31 and 32) showed good cytotoxicity against two cell types of lung cancer (NSCLC‐N6 and A549). This antitumor activity seems to be related to the presence of the phenolic group and the double bond in the five‐carbon ring [86].

The eudesmane sesquiterpenes are formed by two rings of six carbons with the isopropyl group at the carbon 7 and a bromine atom at carbon 1. Eudesmane sesquiterpenes also been isolated from brown seaweeds, for example, from Dictyopteris divaricata [87]. In red seaweeds, the metabolites 1‐bromoselindiene and 9‐bromoselindiene (33 and 34) were isolated species Laurencia composita and demonstrate potent toxicity to brine shrimp [88].

The laurane sesquiterpenes are similar in chemical structure to cuparanes. These molecules have a phenolic group attached to a cyclopentane through carbon 6, where the vast majority has an addition of bromine to carbon 10 or 12. This metabolite class also shows the first examples of iodinated naturally occurring substances [89]. In 2004, a major study was conducted to evaluate the antibacterial activity of secondary metabolites isolated from the genus Laurencia. The Dibromohydroxylaurene substance (35) showed a minimum inhibitory concentration (MIC) of 1.56 μg/mL against Streptococcus pyogenes, Moraxella catarrhalis, and Streptococcus pneumoniae strains [90]. In Greece, 12 sesquiterpenes were isolated from red seaweed Laurencia microcladia being investigated for anticancer activity of these metabolites. The 7‐hydroxylaurene substance (36) showed the best results against K562, MCF7, PC3, HeLa, A431, and CHO cells [91].

The snideranes form the second largest group of sesquiterpenes described for the algae Laurencia genre. These sesquiterpenes are widely studied from an ecological point of view since they are constantly being found in the digestive tract of species of mollusc Aplysia [92]. Some biological activities have been described for 8‐Bromo‐10‐epi‐snyderol substance (37), which showed an IC50 value of 2700 and 4000 ng/mL against strains D6 and W2 clones of Plasmodium falciparum, respectively [93].

The brasilanes sesquiterpenes were isolated only in the species Laurencia obtuse (Hudson) JV Lamouroux, collected on Greek Island Simi [94] and in southern Turkey [95]. Only four brasilanes are known and their biological activities have not been explored.

The diterpenes are composed of 20 carbon atoms, which correspond to four isoprene units. Among the known red seaweeds are more than 20 kinds of diterpene skeletons, with irieane, labdane, and parguerane [96] being the main ones. The first irieane of red seaweeds were isolated in 1975 from the chloroform extract of the genus Laurencia and identified by x‐ray spectroscopy [97]. In 2010, five populations of an unidentified species of Laurencia were collected in Malaysia for chemical research and evaluation of its antibacterial activity. The diterpene 10‐acetoxyangasiol (38) was isolated and demonstrated an MIC of 250 μg/mL and 100 μg/mL against Staphylococcus aureus and Vibrio cholerae, respectively [98].

The labdane diterpenes are bicyclic, usually with ramification on carbon 9. Historically, brominated diterpene Ent‐13‐epiconcinndiol was found in the specie Chondria tenuissima [99]. However, today they are found in the genus Laurencia in great abundance [100]. Despite already having several isolated molecules, their biological activities have been little explored.

The parguerane diterpenes are formed by a tricyclic structure with six carbons in each cycle. All described skeletons have a standard addition of the bromine atom at carbon 15. The vast majority of pargueranes was isolated from red algae, and has a double bond between carbons 9 and 11 with the addition of a hydroxyl group on carbon 16. In a study from Theuri Island, Japan, the isolation of parguerane-type diterpenes was performed and its anticancer activity was tested in P388 and HeLa cells. The monoacetate parguerane diterpene showed the best results with IC50 values of 0.3 and 1.1 µg/mL for HeLa and P388 cells, respectively. The acetoxy group at C2 and bromine at C15 are important for anticancer activity [101].

The triterpenes have 30 carbons in their structure and are derived from six isoprene units [102]. Over 20,000 triterpenoids were isolated and identified in nature, where their structures can be classified into different chemical skeletons, such as squalene, lanostane, dammarane, lupane, oleanane, ursane, and others [103].

The halogenated triterpenoids found in red seaweeds are the type squalene and are known to exhibit excellent anticancer activity. As an example, we can highlight the triterpenoid isolated from red seaweed Laurencia mariannensis collected in China. This compound exhibited significant cytotoxic activity against cancer cells P‐388, with CC50 value of 0.6 µg/mL [104]. In order to obtain products with biological activity, synthesis of some halogenated triterpenes have been proposed, for example, the synthesis of Thyrsiferol (39) [105]. Using low concentrations of this product (3 μM), we observed 60% inhibition of Hypoxia‐inducible factors‐1 in T47D human breast cancer cells. Furthermore, the same natural product has been shown to inhibit the production of messenger RNA of VEGF and GLUT‐1 [106].

The acetogenins are derived from the metabolism of fatty acids. The first acetogenin halogenated C15 of red seaweed was isolated from the methanol extract of seaweed Laurencia glandulifera [107]. Halogenated aromatic polyketides can be classified as linear or cyclic, where the ring of the cyclic metabolites can range from five to twelve atoms in their structure [108]. On the coast of Spain, Laurencia marilzae was collected in the intertidal zone for chemical and biological research evaluation. Its extract was obtained in CH2Cl2/MeOH (1: 1) and subjected to column chromatography using Sephadex and HPLC chromatography. The linear acetogenin Adrienyne (Figure 6A) and its isomer were isolated. The biological activity was tested using the A2780 cell line, HBL‐100, HeLa, SW1573, T‐47D, and WiDr. After the incubation period, the substance showed CC50 greater than 10 µg/mL [108]. Acetogenins cyclic halogenated also have anticancer activity described in the literature [109].

The BPs are substances formed by one or more benzene rings linked to at least one bromine atom. The first BPs isolated from marine organisms were found in red seaweed Rhodomela larix [110]. The BPs are known to have various biological activities [111]. The BPs isolated from the polar extract (MeOH:H2O ‐ 95:5) of Rhodomela confervoides, collected off the coast of China were tested against strains of Staphylococcus aureus, Staphylococcus epidermidis, Escherichia coli, and Pseudomonas aeruginosa. The ether 2,3‐dibromo‐4,5‐dihydroxybenzyl ether (41) showed the best results [112]. Inhibiting the activity of the enzyme Glucose‐6‐phosphate dehydrogenase also was described by BPs [113].

Figure 3.

Bioactive compounds isolated from the red seaweeds.

Advertisement

3. Conclusion

Based on the work described in this chapter, it is clear that seaweed is endowed with a variety of structurally and chemically diverse metabolites having a broad spectrum of biological activities. Of all natural products presented, KF peptide from green seaweed appears to be the most promising in the development of a new drug, since it has excellent biological activity and a known synthesis pathway. We also believe that Dolabelladienetriol, a dolabellane diterpene isolated from Dictyota pfaffii, can be used as an antiviral drug in the future.

References

  1. 1. Peter EL, Rumisha SF, Mashoto KO and Malebo HM. Ethno‐medicinal knowledge and plants traditionally used to treat anemia in Tanzania: a cross sectional survey. Journal of Ethnopharmacology. 2014; 154: 767–773. DOI: 10.1016/j.jep.2014.05.002
  2. 2. Natnoo SA, Mohammed J. Reference flora—A source of traditional medicine in jammu and kashmir. International Journal of Recent Scientific Research. 2014; 5: 2286–2288.
  3. 3. Petrovska BB. Historical review of medicinal plants’ usage. Pharmacognosy Reviews. 2012; 6: 1–5. DOI: 10.4103/0973‐7847.95849
  4. 4. Molinski TF, Dalisay DS, Lievens SL and Saludes JP. Drug development from marine natural products. Nature Reviews Drug Discovery 2009; 8: 69–85. DOI: 10.1038/nrd2487
  5. 5. Hamann MT, Scheuer PJ. Kahalalide F A Bioactive depsipeptide from the sacoglossaa mollusk Elysia rufescens and the green alga Bryopsis sp. Journal of the American Chemical Society. 1993; 115: 5825–5826. DOI: 10.1021/ja00066a061
  6. 6. Hamann MT, Otto CS, Scheuer PJ. Kahalalides: Bioactive peptides from a marine mollusk Elysia rufescens and its algal diet Bryopsis sp. The Journal of Organic Chemistry. 1996; 61: 6594–6600. DOI: 10.1021/jo960877
  7. 7. Garcia‐Rocha M, Bonay P, Avila J. The antitumoral compound Kahalalide F acts on cell lysosomes. Cancer Letters. 1996; 99: 43–50. DOI: 10.1016/0304‐3835(95)04036‐6
  8. 8. Faircloth GT, Grant W, Smith B, Supko J, Brown A, Geldof A, Jimeno J. Preclinical development of Kahalalide F, a new marine compound selected for clinical studies. Proceedings of the American Association for Cancer Research. 2000; 41: 600.
  9. 9. López‐Macià A, Jiménez JC, Royo M, Giralt E, Albericio F. Synthesis and structure determination of Kahalalide F. Journal of the American Chemical Society. 2001; 123: 11398–11401. DOI: 10.1021/ja0116728
  10. 10. Nuijen B, Bouma M, Talsma H, Manada C, Jimeno JM, Lopez‐Lazaro L, Bult A, Beijnen JH. Development of a lyophilized parenteral pharmaceutical formulation of the investigational polypeptide marine anticancer agent Kahalalide F. Drug Development and Industrial Pharmacy. 2001; 27: 767–780. DOI: 10.1081/DDC‐100107240
  11. 11. Brown AP, Morrissey RL, Faircloth GT, Levine BS. Preclinical toxicity studies of kahalalide F, a new anticancer agent: single and multiple dosing regimens in the rat. Cancer Chemotherapy and Pharmacology. 2002; 50: 333–340. DOI: 10.1007/s00280‐002‐0499‐2
  12. 12. Rademaker‐Lakhai JM, Horenblas S, Meinhardt W, Stokvis E, Reijke TM, Jimeno JM, Lopez‐Lazaro L, Martin JAL, Beijnen JH, Schellens JHM. Phase I clinical and pharmacokinetic study of Kahalalide F in patients with advanced androgen refractory prostate cancer. Clinical Cancer Research. 2005; 11: 1854–1862. DOI: 10.1158/1078‐0432.CCR‐04‐1534
  13. 13. Pardo B, Paz‐Ares L, Tabernero J, Ciruelos E, GarcI¤a M, Salazar R, López A, Blanco M, Nieto A, Jimeno J, Izquierdo MA and Trigo JM. Phase I clinical and pharmacokinetic study of Kahalalide F administered weekly as a 1‐hour infusion to patients with advanced solid tumors. Clinical Cancer Research. 2008; 14: 1116–1123. DOI: 10.1158/1078‐0432.CCR‐07‐4366
  14. 14. Miguel‐Lillo B, Valenzuela B, Peris‐Ribera JE, Soto‐Matos A, Pérez‐Ruixo JJ. Population pharmacokinetics of kahalalide F in advanced cancer patients. Cancer Chemotherapy and Pharmacology. 2015; 76: 365–374. DOI: 10.1007/s00280‐015‐2800‐1
  15. 15. Fang Z, Jeong SY, Jung HA, Choi JS, Min BS, Woo MH. Capsofulvesins A–C, cholinesterase Inhibitors from Capsosiphon fulvescens. Chemical and Pharmaceutical Bulletin. 2012; 60: 1351–1358. DOI: 10.1248/cpb.c12‐00268
  16. 16. Islam MN, Choi SH, Moon HE, Park JJ, Jung HA, Woo MH, Woo HC, Choi JS. The inhibitory activities of the edible green alga Capsosiphon fulvescens on rat lens aldose reductase and advanced glycation end products formation. European Journal of Nutrition. 2014; 53: 233–242. DOI: 10.1007/s00394‐013‐0521‐y
  17. 17. Kima JK, Chob ML, Karnjanapratumb S, Shinb S, You SG. In vitro and in vivo immunomodulatory activity of sulfated polysaccharides from Enteromorpha prolifera. International Journal of Biological Macromolecules. 2011; 49: 1051–1058. DOI: 10.1016/j.ijbiomac.2011.08.032
  18. 18. Arata PX, Quintana I, Canelón DJ, Vera BE, Compagnone RS, Ciancia M. Chemical structure and anticoagulant activity of highly pyruvylatedsulfated galactans from tropical green seaweeds of the order Bryopsidales. Carbohydrate Polymers. 2015; 122: 376–386. DOI: 10.1016/j.carbpol.2014.10.030
  19. 19. Högberg HE, Thomson RH, King TJ. The cymopols, a group of prenylated bromohydroquinones from the green calcareous alga Cympolia barbata. Journal of the Chemical Society, Perkin Transactions 1. 1976; 6: 1696–1701. DOI: 10.1039/P19760001696
  20. 20. Estrada DM, Martin JD, Perez C. A new brominated monoterpenoid quinol from Cymopolla barbata. Journal of Natural Products. 1987; 50: 735–737. DOI: 10.1021/np50052a028
  21. 21. Wall ME, Wani MC, Manikumar G, Taylor H, Hughes TJ, Gaetano K, Gerwick WH, McPhail AT, McPhail DR. Plant Antimutagenic Agents 7. Structure and Antimutagenic Properties of Cymobarbatol and 4‐Isocymbarbatol, New Cymopols from Green Alga (Cymopolia barbata). Journal of Natural Products. 1989; 52: 1092–1099. DOI: 10.1021/np50065a028
  22. 22. Dorta E, Darias J, Martín AS, Cueto M. New prenylated bromoquinols from the green alga Cymopolia barbata. Journal of Natural Products. 2002; 65: 329–333. DOI: 10.1021/np010418q
  23. 23. Badal S, Gallimore W, Huang G, Tzeng TR, Delgoda R. Cytotoxic and potent CYP1 inhibitors from the marine algae Cymopolia barbata. Organic and Medicinal Chemistry Letters. 2012; 2: 1–8. DOI: 10.1186/2191‐2858‐2‐21
  24. 24. Güven KC, Percot A, Sezik E. Alkaloids in marine algae. Marine Drugs. 2010; 8: 269–284. DOI: 10.3390/md8020269
  25. 25. Aguilar‐Santos G. Caulerpin, a new red pigment from green algae of the genus Caulerpa. Journal of the Chemical Society, Perkin Transactions 1. 1970; 6: 842–843.
  26. 26. Kamal C, Sethuraman MG. Caulerpin—A bis‐Indole alkaloid as a green inhibitor for the corrosion of mild steel in 1 M HCl solution from the marine alga Caulerpa racemosa. Industrial & Engineering Chemistry Research. 2012; 51: 10399–10407. DOI: 10.1021/ie3010379
  27. 27. Macedo NRPV, Ribeiro MS, Villaça RC, Ferreira W, Pinto AN, Teixeira VL, Cirne‐Santos C, Paixão ICNP, Giongo V. Caulerpin as a potential antiviral drug herpes simplex virus type 1. Revista Brasileira de Farmacognosia. 2012; 22: 861–867. DOI: 10.1590/S0102‐695X2012005000072
  28. 28. Chay CIC, Cansino RG, Pinzón CIE, Torres‐Ochoa RO, Martínez R. Synthesis and anti‐tuberculosis activity of the marine natural product caulerpin and its analogues. Marine drugs. 2014; 12: 1757–1772. DOI: 10.3390/md12041757
  29. 29. Rubinstein I, Goad LJ. Sterols of the siphonous marine alga Codium fragile. Phytochemistry.1974; 13: 481–484. DOI: 10.1016/S0031‐9422(00)91238‐X
  30. 30. Zhang JL, Tian HY, Li J, Jin L, Luo C, Ye WC, Jiang RW. Steroids with inhibitory activity against the prostate cancer cells and chemical diversity of marine alga Tydemania expeditionis. Fitoterapia. 2012; 83: 973–978. DOI: 10.1016/j.fitote.2012.04.019
  31. 31. Ali MS, Saleem M, Yamdagni R, Ali MA. Steroid and antibacterial steroidal glycosides from marine green alga Codium iyengarii Borgesen. Natural Product Letters. 2002; 16: 407–413. DOI: 10.1080/10575630290034249
  32. 32. Carte BK, Troupe N, Chan JA, Westley JW, Faulkner DJ. Rawsonol, an inhibitor of HMG‐CoA reductase from the tropical green alga Avrainvillea rawsoni. Phytochemistry. 1989; 28: 2917–2919. DOI: 10.1016/0031‐9422(89)80253‐5
  33. 33. Plouguerné E, Ioannou E, Georgantea P, Vagias C, Roussis V, Hellio C, Kraffe E, Stiger‐Pouvreau V. Anti‐microfouling activity of lipidic metabolites from the invasive brown alga Sargassum muticum (Yendo) Fensholt. Marine Biotechnology. 2010; 12: 52–61. DOI: 10.1007/s10126‐009‐9199‐9
  34. 34. Cantillo‐Ciau Z, Moo‐Puc R, Quijano L, Freile‐Pelegrín Y. The tropical brown alga Lobophora variegata: a source of antiprotozoal compounds. Marine drugs. 2010; 16: 1292–1304. DOI: 10.3390/md8041292
  35. 35. Hossain Z, Kurihara H, Hosokawa M, Takahashi K. Growth inhibition and induction of differentiation and apoptosis mediated by sodium butyrate in Caco‐2 cells with algal glycolipids. In Vitro Cellular & Developmental Biology. 2005; 41: 154–159. DOI: 10.1290/0409058.1
  36. 36. Mizushina Y, Sugiyama Y, Yoshida H, Hanashima S, Yamazaki T, Kamisuki S, Ohta K, Takemura M, Yamaguchi T, Matsukage A, Yoshida S, Saneyoshi M, Sugawara F, Sakagauchi K. Galactosyldiacylglycerol, a mammalian DNA polymerase alpha‐specific inhibitor from a sea alga, Petalonia bingbamiae. Biological and Pharmaceutical Bulletin. 2001; 24: 982–987. DOI: 10.1248/bpb.24.982
  37. 37. Cumashi A, Ushakova NA, Preobrazhenskaya ME, D'Incecco A, Piccoli A, Totani L, Tinari N, Morozevich GE, Berman AE, Bilan MI, Usov AI, Ustyuzhanina NE, Grachev AA, Sanderson CJ, Kelly M, Rabinovich GA, Iacobelli S, Nifantiev NE. A comparative study of the anti‐inflammatory, anticoagulant, antiangiogenic, and antiadhesive activities of nine different fucoidans from brown seaweeds. Glycobiology. 2007; 17: 541–552. DOI: 10.1093/glycob/cwm014
  38. 38. Croci DO, Cumashi A, Ushakova NA, Preobrazhenskaya ME, Piccoli A, Totani L, Ustyuzhanina NE, Bilan MI, Usov AI, Grachev AA, Morozevich GE, Berman AE, Sanderson CJ, Kelly M, Di Gregorio P, Rossi C, Tinari N, Iacobelli S, Rabinovich GA, Nifantiev NE. Fucans, but not fucomannoglucuronans, determine the biological activities of sulfated polysaccharides from Laminaria saccharina brown seaweed. PLoS One. 2011; 6: 1–10. DOI: 10.1371/journal.pone.0017283
  39. 39. Clinical trials.gov. Phase 1 Dosing Study of BAX 513 in Healthy Volunteers [Internet]. 2010. Available from: https://clinicaltrials.gov/ct2/show/NCT01063101?term=NCT01063101&rank=1 [Accessed: 2015‐12‐14].
  40. 40. Barbosa JP, Pereira RC, Abrantes JL, Cirne dos Santos CC, Rebello MA, Frugulhetti IC, Texeira VL. In vitro antiviral diterpenes from the Brazilian brown alga Dictyota pfaffii. Planta Medica. 2004; 70: 856–860. DOI: 10.1055/s‐2004‐827235
  41. 41. Abrantes JL, Barbosa J, Cavalcanti D, Pereira RC, Frederico Fontes CL, Teixeira VL, Moreno Souza TL, Paixão IC. The effects of the diterpenes isolated from the Brazilian brown algae Dictyota pfaffii and Dictyota menstrualis against the herpes simplex type‐1 replicative cycle. Planta Medica. 2010; 76: 339–344. DOI: 10.1055/s‐0029‐1186144
  42. 42. Cirne‐Santos CC, Teixeira VL, Castello‐Branco LR, Frugulhetti IC, Bou‐Habib DC. Inhibition of HIV‐1 replication in human primary cells by a dolabellane diterpene isolated from the marine algae Dictyota pfaffii. Planta Medica. 2006; 72: 295–299. DOI: 10.1055/s‐2005‐916209
  43. 43. Cirne‐Santos CC, Souza TM, Teixeira VL, Fontes CF, Rebello MA, Castello‐Branco LR, Abreu CM, Tanuri A, Frugulhetti IC, Bou‐Habib DC. The dolabellane diterpene Dolabelladienetriol is a typical noncompetitive inhibitor of HIV‐1 reverse transcriptase enzyme. Antiviral Research. 2008; 77: 64–71. DOI: 10.1016/j.antiviral.2007.08.006
  44. 44. Garrido V, Teixeira GAPB, Teixeira VL, Ocampo P, Ferreira WJ, Cavalcanti DN, Campos SMN, Pedruzzi MMB, Olaya P, Santos CCC, Giongo V, Paixão ICP. Evaluation of the acute toxicity of dolabelladienotriol, a potential antiviral from the brown alga Dictyota pfaffii, in BALB/c mice. Brazilian Journal of Pharmacognosy. 2011; 21: 209–215. DOI: 10.1590/S0102‐695X2011005000053
  45. 45. Pardo‐Vargas A, de Barcelos Oliveira I, Stephens PR, Cirne‐Santos CC, de Palmer Paixão IC, Ramos FA, Jiménez C, Rodríguez J, Resende JA, Teixeira VL and Castellanos L. Dolabelladienols A‐C, new diterpenes isolated from Brazilian brown alga Dictyota pfaffii. Marine Drugs. 2014; 12: 4247–4259. DOI: 10.3390/md12074247
  46. 46. Soares DC, Calegari‐Silva TC, Lopes UG, Teixeira VL, de Palmer Paixão IC, Cirne‐Santos C, Bou‐Habib DC, Saraiva EM. Dolabelladienetriol, a compound from Dictyota pfaffii algae, inhibits the infection by Leishmania amazonensis. PLOS Neglected Tropical Diseases. 2012; 6: e1787. DOI: 10.1371/journal.pntd.0001787
  47. 47. Garcia DG, Bianco EM, Santos Mda C, Pereira RC, Faria MV, Teixeira VL, Burth P. Inhibition of mammal Na(+)K(+)‐ATPase by diterpenes extracted from the Brazilian brown alga Dictyota cervicornis. Phytotherapy Research. 2009; 23: 943–947. DOI: 10.1002/ptr.2600
  48. 48. Vallim MA, Barbosa JE, Cavalcanti DN, De‐Paula JC, da Silva VAGG, Teixeira VL, Paixão ICNP. In vitro antiviral activity of diterpenes isolated from the Brazilian brown alga Canistrocarpus cervicornis. Journal of Medicinal Plants Research. 2010; 4: 2379–2382. DOI: 10.5897/JMPR10.564
  49. 49. de Andrade Moura L, Bianco EM, Pereira RC, Teixeira VL, Fuly AL. Anticoagulation and antiplatelet effects of a dolastane diterpene isolated from the marine brown alga Canistrocarpus cervicornis. Journal of Thrombosis and Thrombolysis. 2011; 31: 235–240. DOI: 10.1007/s11239‐010‐0545‐6
  50. 50. Pereira HS, Leão‐Ferreira LR, Moussatché N, Teixeira VL, Cavalcanti DN, Costa LJ, Diaz R, Frugulhetti IC. Antiviral activity of diterpenes isolated from the Brazilian marine alga Dictyota menstrualis against human immunodeficiency virus type 1 (HIV‐1). Antiviral Research. 2004; 64: 69–76. DOI: 2004 Oct;64(1):69–76
  51. 51. de Souza Pereira H, Leão‐Ferreira LR, Moussatché N, Teixeira VL, Cavalcanti DN, da Costa LJ, Diaz R, Frugulhetti IC. Effects of diterpenes isolated from the Brazilian marine alga Dictyota menstrualis on HIV‐1 reverse transcriptase. Planta Medica. 2005; 71: 1019–1024.
  52. 52. Abrantes JL, Barbosa J, Cavalcanti D, Pereira RC, Fontes CFL, Teixeira VL, Souza TML and Paixão IC. The effects of the diterpenes isolated from the Brazilian brown algae Dictyota pfaffii and Dictyota menstrualis against the herpes simplex type–1 replicative cycle. Planta Medica. 2010; 76: 339–344. DOI: 10.1055/s-0029-1186144
  53. 53. Awad NE, Selim MA, Metawe HM, Matloub AA. Cytotoxic xenicane diterpenes from the brown alga Padina pavonia (L.) Gaill. Phytotherapy Research. 2008; 22: 1610–1613. DOI: 10.1002/ptr.2532
  54. 54. Dorta E, Cueto M, Brito I, Darias J. New terpenoids from the brown alga Stypopodium zonale. Journal of Natural Products. 2002; 65: 1727–1730. DOI: 10.1021/np020090g
  55. 55. Abatisa D, Vagiasa C, Galanakisb D, Norrisc JN, Moreaud D, Roussakisd C, Roussis V. Atomarianones A and B: two cytotoxic meroditerpenes from the brown alga Taonia atomaria. Tetrahedron Letters. 2005; 46: 8525–8529. DOI: 10.1016/j.tetlet.2005.10.007
  56. 56. Mendes G, Soares AR, Sigiliano L, Machado F, Kaiser C, Romeiro N, Gestinari L, Santos N, Romanos MT. In vitro anti‐HMPV activity of meroditerpenoids from marine alga Stypopodium zonale (dictyotales). Molecules. 2011; 16: 8437–8450. DOI: 10.3390/molecules16108437
  57. 57. Okada Y, Ishimaru A, Suzuki R, Okuyama T. A new phloroglucinol derivative from the brown alga Eisenia bicyclis: potential for the effective treatment of diabetic complications. Journal of Natural Products. 2004; 67: 103–105. DOI: 10.1021/np030323j
  58. 58. Jung HA, Hyun SK, Kim HR, Choi JS. Angiotensin‐converting enzyme I inhibitory activity of phlorotannins from Ecklonia stolonifera. Fisheries Science. 2006; 72: 1292–1299. DOI: 10.1111/j.1444‐2906.2006.01288.x
  59. 59. Moon HE, Islam N, Ahn BR, Chowdhury SS, Sohn HS, Jung HA and Choi JS. Protein tyrosine phosphatase 1B and α‐glucosidase inhibitory phlorotannins from edible brown algae, Ecklonia stolonifera and Eisenia bicyclis. Bioscience, Biotechnology, and Biochemistry. 2011; 75: 1472–1480. DOI: 10.1271/bbb.110137
  60. 60. Song F, Xu X, Li S, Wang S, Zhao J, Cao P, Yang Y, Fan X, Shi J, He L, Lü Y. Norsesquiterpenes from the brown alga Dictyopteris divaricata. Journal of Natural Products. 2005; 68: 1309–1313. DOI: 10.1021/np040227y
  61. 61. Song F, Xu X, Li S, Wang S, Zhao J, Yang Y, Fan X, Shi J, He L. Minor sesquiterpenes with new carbon skeletons from the brown alga Dictyopteris divaricata. Journal of Natural Products. 2006; 69: 1261–1266. DOI: 10.1021/np060076u
  62. 62. Banskota AH, Stefanova R, Sperker S, Lall SP, Craigie JS, Hafting JT, Critchley AT. Polar lipids from the marine macroalga Palmaria palmata inhibit lipopolysaccharide‐induced nitric oxide production in RAW264.7 macrophage cells. Phytochemistry. 2014; 101: 101–108. DOI: 10.1016/j.phytochem.2014.02.004
  63. 63. de Souza LM, Sassaki GL, Romanos MT, Barreto‐Bergter E. Structural characterization and anti‐HSV‐1 and HSV‐2 activity of glycolipids from the marine algae Osmundaria obtusiloba isolated from Southeastern Brazilian coast. Marine Drugs. 2012; 10: 918–931. DOI: 10.3390/md10040918
  64. 64. Tsai CJ, Sun Pan B. Identification of sulfoglycolipid bioactivities and characteristic fatty acids of marine macroalgae. Journal of Agricultural and Food Chemistry. 2012; 60: 8404–8410. DOI: 10.1021/jf302241d
  65. 65. Ghosh T, Chattopadhyay K, Marschall M, Karmakar P, Mandal P, Ray B. Focus on antivirally active sulfated polysaccharides: from structure‐activity analysis to clinical evaluation. Glycobiology. 2009; 19: 2–15. DOI: 10.1093/glycob/cwn092
  66. 66. Bouhlal R, Haslin C, Chermann JC, Colliec‐Jouault S, Sinquin C, Simon G, Cerantola S, Riadi H, Bourgougnon N. Antiviral activities of sulfated polysaccharides isolated from Sphaerococcus coronopifolius (Rhodophytha, Gigartinales) and Boergeseniella thuyoides (Rhodophyta, Ceramiales). Marine Drugs. 2011; 9: 1187–1209. DOI: 10.3390/md9071187
  67. 67. Clinical trials.gov. Carrageenan‐Containing Gel in Reducing the Rate of HPV Infection in Healthy Participants [Internet]. 2015. Available from: https://clinicaltrials.gov/ct2/show/NCT02382419?term=seaweed&rank=11 [Accessed: 2015‐12‐14].
  68. 68. Wolwer‐Rieck U, May B, Lankes C, Wust M. Methylerythritol and mevalonate pathway contributions to biosynthesis of mono‐, sesqui‐, and diterpenes in glandular trichomes and leaves of Stevia rebaudiana Bertoni. Journal of Agricultural and Food Chemistry. 2014; 62: 2428–2435. DOI: 10.1021/jf500270s
  69. 69. Getrey L, Krieg T, Hollmann F, Schrader J, Holtmann D. Enzymatic halogenation of the phenolic monoterpenes thymol and carvacrol with chloroperoxidase. Green Chemistry. 2014; 16: 1104–1108. DOI: 10.1039/C3GC42269K
  70. 70. Teixeira V. Produtos Naturais de Algas Marinhas Bentônicas. Revista Virtual de Química. 2013; 5: 343–362. DOI: 10.5935/1984‐6835.20130033
  71. 71. Rovirosa J, Soler A, Blanc V, León R, San‐Martín A. Bioactive monoterpenes from antarctic Plocamium cartilagineum. Journal of the Chilean Chemical Society. 2013; 58: 2025–2026. DOI: 10.4067/S0717‐97072013000400026
  72. 72. Andrianasolo EH, France D, Cornell‐Kennon S and Gerwick WH. DNA methyl transferase inhibiting halogenated monoterpenes from the Madagascar red marine alga Portieria hornemannii. Journal of Natural Products. 2006; 69: 576–579.
  73. 73. Fuller RW, Cardellina JH II, Jurek J, Scheuer PJ, Alvarado‐Lindner B, McGuire M, Gray GN, Steiner JR, Clardy J, Menez E, et al. Isolation and structure/activity features of halomon‐related antitumor monoterpenes from the red alga Portieria hornemannii. Journal of Medicinal Chemistry. 1994; 37: 4407–4411. DOI: 10.1021/jm00051a019
  74. 74. Kladi M, Xenaki H, Vagias C, Papazafiri P, Roussis V. New cytotoxic sesquiterpenes from the red algae Laurencia obtusa and Laurencia microcladia. Tetrahedron. 2006; 62: 182–189. DOI: 10.1002/chin.200621178
  75. 75. König GM, Wright AD, Sticher O. A New polyhalogenated monoterpene from the red alga Plocamium cartilagineum. Journal of Natural Products. 1990; 53: 1615–1618. DOI: 10.1021/np50072a041
  76. 76. Rodriguez S, Kirby J, Denby CM, Keasling JD. Production and quantification of sesquiterpenes in Saccharomyces cerevisiae, including extraction, detection and quantification of terpene products and key related metabolites. Nature Protocols. 2014; 9 (8): 1980–1996. DOI: 10.1038/nprot.2014.132
  77. 77. Davyt D, Fernandez R, Suescun L, Mombru AW, Saldana J, Dominguez L, Fujii MT, Manta E. Bisabolanes from the red alga Laurencia scoparia. Journal of Natural Products. 2006; 69: 1113–1116. DOI: 10.1021/np060235+
  78. 78. Ahmad VU, Ali MS. Terpenoids from marine red alga Laurencia pinnatifida. Phytochemistry. 1991; 30: 4172–4174. DOI: 10.1016/0031‐9422(91)83493‐5
  79. 79. Davyt D, Fernandez R, Suescun L, Mombrú AW, Saldaña J, Domínguez L, Coll J, Fujii MT, Manta E. New sesquiterpene derivatives from the red alga Laurencia scoparia. Isolation, Structure Determination, and Anthelmintic Activity. Journal of Natural Products. 2001; 64: 1552–1555. DOI: 10.1021/np0102307
  80. 80. White DE, Stewart IC, Grubbs RH and Stoltz BM. The catalytic asymmetric total synthesis of elatol. Journal of the American Chemical Society. 2007; 130: 810–811. DOI: 10.1021/ja710294k
  81. 81. Granado I, Caballero P. Chemical defense in the seaweed Laurencia obtusa (Hudson) Lamouroux. Scientia Marina. 1995; 59: 31–39.
  82. 82. De Nys R, Leya T, Maximilien R, Afsar A, Nair PSR and Steinberg PD. The need for standardised broad scale bioassay testing: A case study using the red alga Laurencia rigida. Biofouling. 1996; 10: 213–224. DOI: 10.1080/08927019609386281
  83. 83. Veiga‐Santos P, Pelizzaro‐Rocha KJ, Santos AO, Ueda‐Nakamura T, Dias Filho BP, Silva SO, Sudatti DB, Bianco EM, Pereira RC, Nakamura CV. In vitro anti‐trypanosomal activity of elatol isolated from red seaweed Laurencia dendroidea. Parasitology. 2010; 137: 1661–1670. DOI: 10.1017/S003118201000034X
  84. 84. Campos A, Souza CB, Lhullier C, Falkenberg M, Schenkel EP, Ribeiro‐do‐Valle RM, Siqueira JM. Anti‐tumour effects of elatol, a marine derivative compound obtained from red algae Laurencia microcladia. Journal of Pharmacy and Pharmacology. 2012; 64: 1146–1154. DOI: 10.1111/j.2042‐7158.2012.01493.x
  85. 85. Abad A, Agulló C, Arnó M, Cuñat AC, García MT, Zaragozá RJ. Enantioselective Synthesis of Cuparane Sesquiterpenes. Synthesis of (-)‐Cuparene and (-)‐δ‐Cuparenol. The Journal of Organic Chemistry. 1996; 61: 5916–5919. DOI: 10.1021/jo960463g
  86. 86. Kladi M, Vagias C, Furnari G, Moreau D, Roussakis C, Roussis V. Cytotoxic cuparene sesquiterpenes from Laurencia microcladia. Tetrahedron Letters. 2005; 46: 5723–5726. DOI: 10.1016/j.tetlet.2005.06.076
  87. 87. Ji NY, Wen W, Li XM, Xue QZ, Xiao HL and Wang BG. Brominated Selinane Sesquiterpenes from the Marine Brown Alga Dictyopteris divaricata. Marine Drugs. 2009; 7: 355–360. DOI: 10.3390/md7030355
  88. 88. Li XD, Miao FP, Yin XL, Liu JL, Ji NY. Sesquiterpenes from the marine red alga Laurencia composita. Fitoterapia. 2012; 83: 1191–1195. DOI: 10.1016/j.fitote.2012.07.001
  89. 89. Kladi M, Vagias C, Papazafiri P, Furnari G, Serio D, Roussis V. New sesquiterpenes from the red alga Laurencia microcladia. Tetrahedron. 2007; 63: 7606–7611. DOI: 10.1016/j.tet.2005.09.113
  90. 90. Vairappan CS, Kawamoto T, Miwa H, Suzuki M. Potent antibacterial activity of halogenated compounds against antibiotic‐resistant bacteria. Planta Medica. 2004; 70: 1087–1090. DOI: 10.1055/s‐2004‐832653
  91. 91. Kladi M, Xenaki H, Vagias C, Papazafiri P, Roussis V. New cytotoxic sesquiterpenes from the red algae Laurencia obtusa and Laurencia microcladia. Tetrahedron. 2006; 62: 182–189. DOI: 10.1016/j.tet.2005.09.113
  92. 92. Rogers CN, De Nys R, Charlton TS, Steinberg PD. Dynamics of Algal Secondary Metabolites in Two Species of Sea Hare. Journal of Chemical Ecology. 2000; 26: 721–744. DOI: 10.1023/A:1005484306931
  93. 93. Topcu G, Aydogmus Z, Imre S, Goren AC, Pezzuto JM, Clement JA, Kingston DG. Brominated sesquiterpenes from the red alga Laurencia obtusa. Journal of Natural Products. 2003; 66: 1505–1508. DOI: 10.1021/np030176p
  94. 94. Iliopoulou D, Vagias C, Galanakis D, Argyropoulos D, Roussis V. Brasilane‐Type Sesquiterpenoids from Laurencia obtusa. Organic Letters. 2002; 4: 3263–3266. DOI: 10.1021/ol026506z
  95. 95. Aydoğmus Z, Imre S, Ersoy L, Wray V. Halogenated secondary metabolites from Laurencia Obtusa. Natural Product Research. 2004; 18: 43–49. DOI: 10.1080/1057563031000122086
  96. 96. Kornprobst JM, Al‐Easa HS. Brominated diterpenes of marine origin. Current Organic Chemistry. 2003; 7: 1181–1229. DOI: 10.2174/1385272033486530
  97. 97. Fenical W, Howard B, Gifkins KB, Clardy J. Irieol A and iriediol, dibromoditerpenes of a new skeletal class from Laurencia. Tetrahedron Letters. 1975; 16: 3983–3986. DOI: 10.1016/S0040‐4039(00)91215‐2
  98. 98. Vairappan CS, Ishii T, Lee TK, Suzuki M, Zhaoqi Z. Antibacterial activities of a new brominated diterpene from Borneon Laurencia spp. Marine Drugs. 2010; 8: 1743–1749. DOI: 10.3390/md8061743
  99. 99. Öztunç A, Imre S, Lotter H, Wagner H. Ent‐13‐epiconcinndiol from the red alga Chondria tenuissima and its absolute configuration. Phytochemistry. 1989; 28: 3403–3404. DOI: 10.1016/0031‐9422(89)80356‐5
  100. 100. Suzuki M, Kawamoto T, Vairappan CS, Ishii T, Abe T, Masuda M. Halogenated metabolites from Japanese Laurencia spp. Phytochemistry. 2005; 66: 2787–2793. DOI: 10.1016/j.phytochem.2005.08.008
  101. 101. Takeda S, Kurosawa E, Komiyama K, Suzuki T. The Structures of Cytotoxic Diterpenes Containing Bromine from the Marine Red Alga Laurencia obtusa (Hudson) Lamouroux. Bulletin of the Chemical Society of Japan. 1990; 63: 3066–3072. DOI: 10.1246/bcsj.63.3066
  102. 102. Connolly JD, Hill RA. Triterpenoids. Natural Product Reports. 2002; 19: 494–513. DOI: 10.1039/B110404G
  103. 103. Hill RA, Connolly JD. Triterpenoids. Natural Product Reports. 2013; 30: 1028–1065. DOI: 10.1039/C3NP70032A
  104. 104. Ji NY, Li XM, Xie H, Ding J, Li K, Ding LP, Wang BG. Highly Oxygenated Triterpenoids from the Marine Red Alga Laurencia mariannensis (Rhodomelaceae). Helvetica Chimica Acta. 2008; 91: 1940–1946. DOI: 10.1002/hlca.200890207
  105. 105. Hashimoto M, Kan T, Nozaki K, Yanagiya M, Shirahama H, Matsumoto T. Total syntheses of (+)‐thyrsiferol, (+)‐thyrsiferyl 23‐acetate, and (+)‐venustatriol. The Journal of Organic Chemistry. 1990; 55: 5088–5107. DOI: 10.1021/jo00304a022
  106. 106. Mahdi F, Falkenberg M, Ioannou E, Roussis V, Zhou YD, Nagle DG. Thyrsiferol Inhibits Mitochondrial Respiration and HIF‐1 Activation. Phytochemistry Letters. 2011; 4: 75–78. DOI: 10.1016/j.phytol.2010.09.003
  107. 107. Irie T, Suzuki M, Masamune T. Laurencin, a constituent from laurencia species. Tetrahedron Letters. 1965; 6: 1091–1099. DOI: 10.1016/S0040‐4039(00)90038‐8
  108. 108. Gutierrez‐Cepeda A, Fernandez JJ, Gil LV, Lopez‐Rodriguez M, Norte M, Souto ML. Nonterpenoid C15 acetogenins from Laurencia marilzae. Journal of Natural Products. 2011; 74: 441–448. DOI: 10.1021/np100866g
  109. 109. Kladi M, Vagias C, Papazafiri P, Brogi S, Tafi A, Roussis V. Tetrahydrofuran acetogenins from Laurencia glandulifera. Journal of Natural Products. 2009; 72: 190–193.
  110. 110. 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. DOI: 10.1016/0040‐4020(67)85068‐3
  111. 111. Liu M, Hansen PE, Lin X. Bromophenols in Marine Algae and Their Bioactivities. Marine Drugs. 2011; 9: 1273–1292. DOI: 10.3390/md9071273
  112. 112. Xu N, Fan X, Yan X, Li X, Niu R, Tseng CK. Antibacterial bromophenols from the marine red alga Rhodomela confervoides. Phytochemistry. 2003; 62: 1221–1224. DOI: 10.1016/S0031‐9422(03)00004‐9
  113. 113. Mikami D, Kurihara H, Kim SM, Takahashi K. Red Algal Bromophenols as Glucose 6‐Phosphate Dehydrogenase Inhibitors. Marine Drugs. 2013; 11: 4050–4057. DOI: 10.3390/md11104050

Written By

Caio Cesar Richter Nogueira and Valéria Laneuville Teixeira

Submitted: 21 September 2015 Reviewed: 07 March 2016 Published: 29 June 2016