Acessibilidade / Reportar erro

Partial characterization and anticoagulant activity of sulfated galactan from the green seaweed Halimeda opuntia

Abstract

The number of deaths associated with cardiovascular diseases (CVD) increases every year, leading to an intense search for new compounds that may be employed as anticoagulants. One of the classes of bioprospected molecules comprises sulfated polysaccharides (SP) from seaweed, as heparin displays many adverse effects associated with its use. The present study aimed to characterize and evaluate the anticoagulant potential of SP extracted from the green algae Halimeda opuntia. Four PS-rich fractions, F23, F44, F60 and F75, were obtained by proteolytic digestion in papain followed by ethanol precipitation. The presence of SP was confirmed by agarose gel electrophoresis, revealing different populations in each fraction. The F44 fraction is noteworthy compared to the other fractions, presenting a 5% yield compared to the initial algae weight and anticoagulant activity revealed by the activated partial thromboplastin time (APTT) assay (intrinsic/common coagulation pathway). Surprisingly, F44 purification (SP peak P1F44) resulted in prothrombin time (PT) activity (extrinsic coagulation pathway) at a 160 µg/mL, in addition to enhanced APTT activity. The P1F44 anticoagulant activity mechanism was shown to be dependent on two coagulations factors, IIa and Xa, more potent via IIa. Future assessments will be performed to assess this fraction in the medical clinic.

Key words
Marine algae; Sulfated polysaccharides; Blood clotting; Heparin-like; Factor IIa

INTRODUCTION

According to the World Health Organization (WHO), cardiovascular diseases (CVD) are currently the leading cause of death worldwide. About 17.9 million people lose their lives each year due to CVD, representing 31% of all global deaths, with a third of these deaths befalling people under 70 years old (WHO 2017WHO. 2017. Cardiovascular diseases. Available at: https://www.who.int/health-topics/cardiovascular-diseases. Accessed on: 14/04/2020.
https://www.who.int/health-topics/cardio...
). CVDs include coronary heart disease, cerebrovascular disease, rheumatic heart disease and other conditions generally associated with thromboembolic events due to the formation of thrombi in the circulatory system. Because of this, CVD diagnosis and treatment employing anticoagulant therapy is essential for people presenting CVD or at high cardiovascular risk (Mourão 2015MOURÃO PAS. 2015. Perspective on the use of sulfated polysaccharides from marine organisms as a source of new antithrombotic drugs. Mar Drugs 13: 2770-2784., WHO 2017WHO. 2017. Cardiovascular diseases. Available at: https://www.who.int/health-topics/cardiovascular-diseases. Accessed on: 14/04/2020.
https://www.who.int/health-topics/cardio...
).

The use of anticoagulant agents began with the discovery of unfractionated heparin (UFH), a sulfated polysaccharide (SP) belonging to the glycosaminoglycan (GAGs) family. This polymer, composed mainly of L-iduronic-2-O-sulfated acid and D-glucosamine-N-sulfated, displays several advantages due to its specificity and the fact that is does not cause anaphylaxis, and is considered one of the most important advances for the development of cardiac surgery (Olson et al. 1992OLSON ST, BJORK I, SHEFFER R, CRAIG PA, SHORE JD & CHOAY J. 1992. Role of the antithrombin-binding pentasaccharide in heparina acceleration conformational change contribution to heparin rate enhancement. J Bio Chem 267: 12528-12538., Sommers et al. 2011SOMMERS CD, YE H, KOLINSKI RE, NASR M, BUHSE LF, AL-HAKIM A & KEIRE DA. 2011. Characterization of currently marketed heparin products: analysis of molecular weight and heparinase-I digest patterns. Anal Bioanal Chem 401: 2445-2454.), followed by the discovery of hydroxycoumarin (warfarin). However, despite their effectiveness, both exhibit a series of limitations and adverse effects that restrict their clinical use (Alban 2005ALBAN S. 2005. The ‘precautionary principle’ as a guide for future drug development. Eur J Clin Investig 35: 33-44., White 2006WHITE RH. 2006. The epidemiology of venous thromboembolism. Thromb Haemost 21: 23-29., Pan et al. 2010PAN J, QIAN Y, ZHOU XD, PAZANDAK A, FRAZIER SB, WEISER P, LU H & ZHANG LJ. 2010. Oversulfated chondroitin sulfate is not the sole contaminant in heparin. Nat Biotechnol 28: 203-207., Mourão 2015MOURÃO PAS. 2015. Perspective on the use of sulfated polysaccharides from marine organisms as a source of new antithrombotic drugs. Mar Drugs 13: 2770-2784.). Because of this, low molecular weight heparin (LMWH) was developed in the 1980s, with the aim of reducing such limitations, and is still employed today. In fact, LMWH has been shown to be a safe and effective drug that does not require laboratory monitoring, with a longer half-life and predictable response. However, its administration, as well as that of UFH, remains parenteral, also presenting thrombocytopenia risks (Sattari & Lowelthal 2011SATTARI M & LOWENTHAL DT. 2011. Novel Oral Anticoagulants in Development: Dabigatran, Rivaroxaban, and Apixaban. Am J Ther 18: 332-338., Van Der Wall et al. 2017VAN DER WALL SJ, KLOK FA, DEN EXTER PL, BARRIOS D, MORILLO R, CANNEGIETER SC, JIMENEZ D & HUISMAN MV. 2017. Higher Adherence to Treatment With Low-Molecular-Weight-Heparin Nadroparin Than Enoxaparin Because of Side Effects in Cancer-Associated Venous Thromboembolism. HemaSphere 2:1.).

The development of semi-synthetic compounds, such as LMWH and pentasaccharides (fondaparinux), in recent decades has triggered the search for the ideal anticoagulant. In this regard, the need for equally effective, easy to use oral anticoagulants displaying better safety profiles has led to the development of novel molecules with mechanisms of action based on direct thrombin or factor Xa inhibition, such as dabigratan, rixoxaban, prasugrel and ticagrelor (Flato et al. 2010FLATO UAP, BUHATEM T, MERLUZZI T & BIANCO ACM. 2010. Novos anticoagulantes em cuidados intensivos. Rev Bras Ter Intensiva 23: 68-77., Yoshida et al. 2011YOSHIDA RDA, YOSHIDA WB & ROLLO HA. 2011. Novos anticoagulantes para a profilaxia do tromboembolismo venoso em cirurgias ortopédicas de grande porte. J Vasc Brasil 145-153., Lorga Filho et al. 2013LORGA FILHO AM ET AL. 2013. Diretrizes brasileiras de antiagregantes plaquetários e anticoagulantes em cardiologia. Arq Bras Cardiol 101: 3.). However, scientific data have reported adverse reactions in several patients, often leading to treatment discontinuation, as well as high costs (Sattari & Lowelthal 2011, Afonso et al. 2016AFONSO A, MARQUES G, GONÇALVES A, BARROSO P, GONZALEZ A, RODRIGUES H & FERREIRA MJA. 2016. A terapêutica antitrombótica: atual e em desenvolvimento. Angiologia e Cirurgia Vascular 12: 170-179., Van Der Wall et al. 2017VAN DER WALL SJ, KLOK FA, DEN EXTER PL, BARRIOS D, MORILLO R, CANNEGIETER SC, JIMENEZ D & HUISMAN MV. 2017. Higher Adherence to Treatment With Low-Molecular-Weight-Heparin Nadroparin Than Enoxaparin Because of Side Effects in Cancer-Associated Venous Thromboembolism. HemaSphere 2:1.). Currently, although heparin remains the ideal model, numerous substances are at different development stages, aimed at obtaining fully synthetic compounds from mimetic oligosaccharides and developing orally active heparin formulations (Correia-da-Silva et al. 2013CORREIA-DA-SILVA M, SOUSA E, MARQUES F & PINTO MM. 2013. Estado da arte na terapêutica anticoagulante: Novas abordagens. Acta Farm Port 2: 5-18., Afonso et al. 2016AFONSO A, MARQUES G, GONÇALVES A, BARROSO P, GONZALEZ A, RODRIGUES H & FERREIRA MJA. 2016. A terapêutica antitrombótica: atual e em desenvolvimento. Angiologia e Cirurgia Vascular 12: 170-179.). In this regard, many research groups have begun prospecting and investigating biologically active molecules with anticoagulant potential that can function as substitutive or alternative therapies to heparin in certain clinical indications (Melo et al. 2012MELO KRT, ALMEIDA-LIMA J, GOMES DL, DANTAS-SANTOS N CAMARA RGB & ROCHA HAO. 2012. Caracterização e atividade anticoagulante de PS extraídos da alga marrom Dictyopteris justii. Holos 1: 29-40., Sujian et al. 2019SUJIAN C, XIAOXI H, LING Q, MEIJIA H, YAJING Y, ZHICHUN L & WENJUN M. 2019. Anticoagulant and Antithrombotic Properties in vitro and in vivo of a Novel Sulfated Polysaccharide from Marine Green Alga Monostroma nitidum. Mar Drugs 17: 247.). Thus, SP constitute a complex group of macromolecules that have aroused great interest in the scientific community, mainly due to the fact that these compounds naturally display a polyanionic character capable of interacting with important functional proteins (Pomin & Mourão 2014POMIN VH & MOURÃO PA. 2014. Specific sulfation and glycosylation - a structural combination for the anticoagulation of marine carbohydrates. Front Cell Infect Mi 4: 1-8., Adrien et al. 2019ADRIEN A, BONNET A, DUFOUR D, BAUDOUIN S, MAUGARD T & BRIDIAU N. 2019. Anticoagulant Activity of Sulfated Ulvan Isolated from the Green Macroalga Ulva rígida. Mar Drugs 17: 291.), found in microorganisms (Kusaykin et al. 2006KUSAYKIN MI, CHIZHOV AO, GRACHEV AA, ALEKSEEVA SA, BAKUNINA IY, NEDASHKOVSKAYA OI, SOVA VV & ZVYAGINTSEVA TNA. 2006. A comparative study of specificity of fucoidanases from marine microorganisms and invertebrates. J Appl Phyc 18: 143-147.), animals (both vertebrates and invertebrates) (Nader et al. 2004NADER HB, LOPES CC, ROCHA HA, SANTOS EA, DIETRICH CP. 2004. Heparins and heparinoids: occurrence, structure and mechanism of antithrombotic and hemorrhagic activities. Curr Pharm Des 10: 951-966.), seaweed (Costa et al. 2010COSTA LS ET AL. 2010. Biological activities of sulfated polysaccharides from tropical seaweeds. Biomed Pharmacother 64: 21-28.) and some higher aquatic plants (Aquino et al. 2005AQUINO RS, LANDEIRA-FERNANDEZ AM, VALENTE AP, ANDRADE LR & MOURÃO PA. 2005. Occurrence of sulfated galactans in marine angiosperms: evolutionary implications. Glycobiology 15: 11-20., Dantas-santos et al. 2012DANTAS-SANTOS N ET AL. 2012. Freshwater plants synthesize sulfated polysaccharides: heterogalactans from water hyacinth (Eicchornia crassipes). Int J Mol Sci 13: 961-976.).

Marine algae are the main source of SP, which are often found in the extracellular matrix of these organisms, containing many active substances and metabolites (Silva & Neto 2009SILVA PM & NETO PC. 2009. Atividades biológicas de extratos de algas marinhas brasileiras – Master’s degree. Universidade de São Paulo, São Paulo, SP. (Unpublished).). SP are complex constituents, and their structure varies between different algae species, both concerning the type of constituent sugar and the position of the glycosidic bond and sulfation site, an important factor that determines their specific biological functions (Hu et al. 2012HU Y, YU GL, ZHAO XX, WANG YF, SUN XC, JIAO GL, ZHAO X & CHAI WG. 2012. Structural characterization of natural ideal 6-O-sulfated agarose from red alga Gloiopeltis furcata. Carbohydr Polym 89: 883-889., Mourão 2015MOURÃO PAS. 2015. Perspective on the use of sulfated polysaccharides from marine organisms as a source of new antithrombotic drugs. Mar Drugs 13: 2770-2784.). Furthermore, seaweed use increases every year (Calumpong et al. 2017CALUMPONG HP, WEST J & SEAWEEDS MG. 2017. In the First Global Integrated Marine Assessment: World Ocean Assessment I; United Nations, Cambridge University Press: Cambridge, UK., Ferdouse et al. 2018FERDOUSE F, LOVESTAD HOLDT S, SMITH R, MURUA P & YANG Z. 2018. The Global Status of Seaweed Production, Trade and Utilization; FAO Globefish Research Programme; Food and Agriculture Organization of the United Nations: Rome, Italy.), with the extraction of SP exhibiting numerous biological activities, i.e., anticoagulant (Anderson et al. 1976ANDERSON LO, BARROWCLIFFE TW, HOLMER E, JOHNSON EA & SIMS GFC. 1976. Anticoagulant properties of heparin fractionated by affinity chromatography on matrix-bound antithrombin-3 and by gel-filtration. Thromb Res 9: 575-580., Pereira et al. 1999PEREIRA MS, MULLOY B & MOURÃO PA. 1999. Structure and anticoagulant activity of sulfated fucans: Comparison between the regular repetitive and linear fucans from echinoderms with the more heterogeneous and branched polymers from brown algae. J Biol Chem 274: 7656-7667., 2002PEREIRA MS, MELO FR & MOURÃO PA. 2002. Is there a correlation between structure and anticoagulant action of sulfated galactans and sulfated fucans? Glycobiology 12: 573-580., 2005PEREIRA MG, BENEVIDES NM, MELO MR, VALENTE AP, MELO FR & MOURÃO PA. 2005. Structure and anticoagulant activity of a sulfated galactan from the red alga, Gelidium crinale. Is there a specific structural requirement for the anticoagulant action? Carbohydr Res 340: 2015-2023., Mourão & Pereira 2000MOURÃO PAS & PEREIRA MS. 2000. Searching from alternatives to heparin sulfated fucans from marine invertebrates. Trends Cardiovasc Med 9: 225-232.), antithrombotic (Mourão & Pereira 2000MOURÃO PAS & PEREIRA MS. 2000. Searching from alternatives to heparin sulfated fucans from marine invertebrates. Trends Cardiovasc Med 9: 225-232., Mourão 2004MOURÃO PAS. 2004. Estrutura versus atividade biológica de polissacarídeos sulfatados. In.: XXVI Reunião Anual sobre Evolução, Sistemática e Ecologia Micromolecular, 2004, Niterói.), antitumoral (Soeda et al. 1994SOEDA S, ISHIDA S, HONDA O, SHIMENO H & NAGAMATSU A. 1994. Amineted fucoidan promotes the invasion of 3 LL cells through reconstituted basement membrane: Its possible mechanism of action. Cancer Lett 85: 133-138.), antiproliferative (Costa et al. 2010COSTA LS ET AL. 2010. Biological activities of sulfated polysaccharides from tropical seaweeds. Biomed Pharmacother 64: 21-28.), antiviral (Damonte et al. 1994DAMONTE E ET AL. 1994. Antiviral activity of a sulphated polysaccharide from the red seaweed Nothogenia fastigiata. Biochem Pharmacol 47: 2187-2192.), antioxidant (Zhang et al. 2003ZHANG Q, LI N, ZHOU G, LU X, XU Z & LI Z. 2003. In vivo antioxidant activity of polysaccharide fraction from Porphyra haitanesis (Rhodophyta) in aging mice. Pharmacol Res 48: 151-155.), antiinflammatory (Berteau & Mulloy 2003BERTEAU O & MULLOY B. 2003. Sulfated fucans, fresh perspectives: structures, functions, and biological properties of sulfated fucans and an overview of enzymes active toward this class of polysaccharide. Glycobiology 13: 29R-40R.) and antinociceptive (Vieira et al. 2004VIEIRA LAP, FREITAS ALP, FEITOSA JPA, SILVA DC & VIANA GSB. 2004. The alga Bryothamnion seaforthii contains carbohydrates with antinociceptive activity. Braz J Med Bio Res 37: 1071-1079.) properties.

In this context, bioprospecting is undoubtedly of paramount importance for the specific knowledge of the chemical and pharmacological potential of marine natural products, especially concerning green algae, as only 1.3% of all compounds described from 1963 to 2013 belong to this phylum (Chlorophyta), against 7.3% extracted from Rodophyta (Blunt et al. 2015BLUNT JW, COPP BR, KEYZERS RA, MUNRO MH & PRINSEP MR. 2015. Marine natural products. Nat Prod Rep 32: 116-211.). In this regard, the green macroalgae Halimeda opuntia is noteworthy, as it is abundant throughout the Brazilian coast and easily cultivated, and scarce information is available concerning its anticoagulant activity. Although one study suggesting the structure of H. opuntia SP and anticoagulant activity has been recently published (Arata et al. 2015ARATA PX, QUINTANA I, CANELÓN DJ, VERA BE, COMPAGNONE RS & CIANCIA M. 2015. Chemical structure and anticoagulant activity of highly pyruvylated sulfated galactans from tropical green seaweeds of the order Bryopsidales. Carbohydr Polym 122: 376-386.), the extraction technique differs from the one proposed herein. Additionally, selective precipitation and purification methods were not carried out, leading to different results.

Therefore, this study aimed to extract and characterize SP from the seaweed H. opuntia and evaluate its anticoagulant activity concerning the intrinsic and/or common pathway and the extrinsic pathway, through the activated partial thromboplastin time (APTT) and prothrombin time (PT) assays, respectively, as well as the specific coagulation system antithrombin-dependent pathway (anti-Xa and anti-IIa).

MATERIALS AND METHODS

SP extraction and precipitation

Halimeda opuntia specimens were obtained at Cumuxuratiba beach (17˚ 6` 57.841” S and 39˚ 10` 18.779” O), in the Prado municipality (BA, Brazil), separated from other species, washed with distilled water and dried at room temperature. Subsequently, the crushed seaweed (3.5 g) was immersed and maintained in acetone for 24 h for delipidation and depigmentation. The collected sediment was dried at 60°C, suspended in an extraction buffer (0.1 M sodium acetate, 5.0 mM ethylenediamine tetraacetic acid (EDTA) and 5.0 mM cysteine at pH 6.0 with the addition of papain) and incubated at 60 °C for 24 h under agitation (200 rpm). The incubation mixture was then centrifuged (2500 x g, 20 min) at room temperature and the supernatant separated. The residue was resuspended in the same extraction buffer until the absence of SP in the supernatant, verified by the metachromatic property in dimethylethylene blue (DMB - Sigma-Aldrich) at A525nm. Positive supernatants were combined (1,2 L) and termed the crude extract (CE), which was then fractionated by precipitation in increasing ethanol concentrations (30, 80, 150 and 300%). Based on this methodology, final ethanol concentrations were considered as 23, 44, 60 and 75%, respectively. Subsequently, ethanol volumes were added to the solutions, which were maintained at 4 °C for 12 h. The precipitates were then separated from the solutions by centrifugation (10,000 x g, 10 min), dialyzed exhaustively against distilled, the SP were concentrated by freeze-drying, weighed and stored for further analysis. This procedure was repeated with increasing ethanol volumes up to the final concentration in a 75% solution. Each fraction obtained by precipitation was renamed using the number corresponding to the final applied ethanol concentration, namely F23 (25 mg), F44 (172 mg), F60 (566 mg) and F75 (302 mg). Yields were calculated compared to the initial seaweed weight (3.5 g).

Agarose gel electrophoresis

An initial SP analysis was performed by agarose gel electrophoresis. The obtained fractions were applied to a 0.5% agarose gel for 1 h at 110 V in 1,3-diaminopropane-acetate 0.05 M (pH 9.0). The SP were then fixed using a 0.1% N-cetyl-N,N,N-trimethylammonium bromide solution. After 12 h, the gel was dehydrated and stained with 0.1% toluidine blue in 0.1:5:5 (v/v) acetic acid-ethanol water (Dietrich & Dietrich 1976DIETRICH CP & DIETRICH SMC. 1976. Electrophoretic behaviour of acidic mucopolysaccharides in diamine buffers. Analytical Biochemistry, v.70, n.02, p.645-647.). Chondroitin sulfate (Sigma-Aldrich) and bovine heparin (200.5 IU mg−1 - National Institute of Standards and Biological Control/INPCB) were used as standards.

In vitro coagulation assays

To analyze the anticoagulant activity of each fraction, human plasma was collected in a 3.8% sodium citrate solution at a 9:1 ratio (Anderson et al. 1976ANDERSON LO, BARROWCLIFFE TW, HOLMER E, JOHNSON EA & SIMS GFC. 1976. Anticoagulant properties of heparin fractionated by affinity chromatography on matrix-bound antithrombin-3 and by gel-filtration. Thromb Res 9: 575-580.) and analyzed by the activated partial thromboplastin time (APTT) and prothrombin time (PT) assays. Both tests were carried out using commercial kits under experimental conditions recommended by the manufacturer (Labtest, São Paulo, SP, Brazil). A curve was performed using 100 µl volume of plasma plus sample, varying the volumes of SP samples by 2, 4, 6, 8 and 10 µ l. Clotting times were determined using a microcoagulometer (Amelug, model KC4A). The concentration of SP was calculated considering the final volume of the entire assay, being 300 µl in aPTT and 200 µl in PT. All assays were performed in triplicate and the results were compared to an unfractionated heparin (UFH) control (Melo et al. 2008MELO EI, PEREIRA MS, CUNHA RS, LEME MPS & MOURÃO PAS. 2008. Heparin quality control in the Brazilian market: implications in the cardiovascular surgery. Braz J Cardiovasc Surg 23: 169-174.).

SP purification

The F44 fraction (172 mg) was purified by ion exchange chromatography employing a Diethylaminoethyl (DEAE)-cellulose (BioRad) column. The column was equilibrated and washed with an equilibration buffer (20mM Tris-hydrochloric acid (HCl), 50 mM EDTA pH 7.4) and the SP were eluted under a linear sodium chloride (NaCl) gradient from 0 to 4.0 M, solubilized in an equilibrium buffer, at a 1 mL/min flow, and 2 mL fractions were collected and monitored at A525nm by means of the metachromatic property with DMB (Farndale et al. 1986FARNDALE RW, BUTTLE DJ & BARRETT AJ. 1986. Improved quantitation and discrimination of sulphated glycosaminoglycans by use of dimethylmethylene blue. Biochim Biophys Acta 883: 173-177.) and at A490nm to monitor the presence of hexoses by the phenol/sulfuric acid analysis. The obtained SP fractions were pooled, dialyzed exhaustively against distilled water and filtered through 3 kDa Amicon filters with Milli-Q water until complete absence of salts (monitored with AgNO3). Finally, the SP were concentrated by freeze-drying, weighed (20 mg) and stored at -20°C. Yields were calculated compared to the initial fraction weight (Farndale et al. 1986FARNDALE RW, BUTTLE DJ & BARRETT AJ. 1986. Improved quantitation and discrimination of sulphated glycosaminoglycans by use of dimethylmethylene blue. Biochim Biophys Acta 883: 173-177.).

Enzyme inhibition assays

Purified samples exhibiting anticoagulant potential detected by the APTT and TP assays were incubated in 96-well plates with 40 µL of a 20 mM Tris-HCl solution containing 150 mM NaCl and 0.1% polyethylene glycol (PEG) 8000, pH 7.5, AT (10 nM) and factor Xa (2 nM) or thrombin (2 nM) for 1 min at 37°C. Factor Xa and thrombin activities were determined by adding the chromogenic substrates S-2765 or S-2238 (100 μM), respectively. Kinetic test was performed and Substrate hydrolysis was detected (A405nm), accompanied by 5 minutes, using a microplate reader and the absorbance change rates were proportional to the residual thrombin or factor Xa activity remaining in the incubation solution (Glauser et al. 2009GLAUSER BF, REZENDE RM, MELO FR, PEREIRA MS, FRANCISCHETTI IMB, MONTEIRO RQ, REZAIE AR & MOURÃO PAS. 2009. Anticoagulant activity of a sulfated galactan: Serpin-independent effect and specific interaction with factor Xa. Thromb Haemost 102(06): 1183-1193.).

Chemical characterization

Sulfate content was determined according to the gelatin-barium method (Dodgson & Price 1962DODGSON KS & PRICE RG. 1962. A note on the determination of the ester sulphate content of sulphated polysaccharides. Biochem J 84: 106-110.) employing a standard sodium sulfate (1 μg/μL). The Bradford method (1976BRADFORD MM. 1976. A rapid and sensitive method for the quantification of microgram quantities of protein utilizing the principle of protein-dye binding. Anal Biochem 72: 248-254.) was performed for total protein determination using bovine albumin as standard (1 μg/μL). Uronic acid content was quantified according to Dische (1947)DISCHE Z. 1947. A specific color reaction for glucuronic acid. J Biol Chem 171: 725-730. in the presence of H2SO4 and borate, using glucuronolactone (0.1 μg/μL) as the standard. In all dosages, calibration curves were used with determined volumes of the standards.

Structural assessment

Infra-red analyses

About 0.1 mg of the purified and freeze-dried samples were added to 1 mg of potassium bromide and crushed until obtaining a homogeneous mixture. Next, pellets were prepared applying a force of 15 tons using a hydraulic press, which were then submitted to infrared analyses. Infrared spectra were recorded employing a Perkin Elmer Spectrum 65 FT-IR spectrophotometer between 400 and 4000 cm-1. Measurements were performed under a 4 cm-1 scanning resolution and normal environmental conditions (Chopin & Whalen 1993CHOPIN T & WHALEN E. 1993. A new rapad method for carrageenan indentification by FT IR diffuse reflectance spectroscopy directly on dried, ground algal material. Carbohydr Res 246: 51-59.).

Monosaccharide composition analyses

Initially, 5 mg of purified SP were hydrolyzed in the presence of trifluoroacetic acid at a final concentration of 5 M for 4 hours at 100 °C. After cleavage, the sugars were reduced with sodium borohydride and the produced alditols acetylated with acetic anhydride:pyridine (1:1, v/v). The acetylated alditols were then dissolved in chloroform and analyzed by gas chromatography (GC MS-QP2010 Shimadzu, Japan) using a Restek RTX-5MS column. The initial run temperature was 110 °C and the final one, 250 °C, at an increasing rate of 2 °C/min (Kircher 1960KIRCHER HW. 1960. Gas-liquid partition chromatography of methylated sugars. Anal Chem 32: 1103-1106.).

Statistical analyses

Results were expressed as the means ± standard error of the means (SEM) of an indicated number of experiments. Results were analyzed by the Student t test using the OriginPro 8 software. Statistical significance was set as p<0.05 and only curves with R > 0.98 were considered for the calculation purposes at the chemical dosages mentioned above.

RESULTS AND DISCUSSION

H. opuntia SP extraction and characterization

SP were extracted by enzymatic treatment and the obtained extracts following proteolysis were grouped (crude extract - CE) and precipitated with increasing and cumulative ethanol concentrations (23%, 44%, 60% and 75%). Sufficient amounts of SP were not detected in the fraction precipitated with 9%. Thus, only four fractions were considered, F23, F44, F60 and F75.

After the initial procedures, the seaweed material was freeze-dried and weighed. The dry weight obtained for each fraction was compared to the initial seaweed weight, allowing for SP yield calculations for each fraction. Fraction F60 exhibited the highest yield, at 16% of the total extracted mass, while the F23 fraction yield was less than 1%. Fractions F44 and F75 presented 5 and 8.5% yields, respectively. Several SP extraction methodologies are available, such as the use of hot and/or cold aqueous solutions (Adrien et al. 2017ADRIEN A, BONNET A, DUFOUR D, BAUDOUIN S, MAUGARD T & BRIDIAU N. 2017. Pilot production of ulvans from Ulva sp. and their effects on hyaluronan and collagen production in cultured dermal fibroblasts. Carbohydr Polym 157: 1306-1314.) under controlled acidic conditions and enzymatic extraction employing different proteolytic enzymes displaying nonspecific action, in addition to the papain used herein (Farias et al. 2000FARIAS WRL, VALENTE AP, PEREIRA MS & MOURÃO PAS. 2000. Structure and anticoagulant activity of sulfacted galactans. J Biol Chem 275: 29299-29307., 2008FARIAS EHC, POMIN VH, VALENTE AP, NADER HB, ROCHA HAO & MOURÃO PAS. 2008. A preponderantly 4-sulfated, 3-linked galactan from the green alga Codium isth-mocladum. Glycobiology 18: 250-259., Melo et al. 2012MELO KRT, ALMEIDA-LIMA J, GOMES DL, DANTAS-SANTOS N CAMARA RGB & ROCHA HAO. 2012. Caracterização e atividade anticoagulante de PS extraídos da alga marrom Dictyopteris justii. Holos 1: 29-40.). Each experimental protocol is capable of extracting compounds displaying different properties, chemical compositions, and yields. For example, Arata and collaborators extracted SP from H. opuntia using two different methods: aqueous solutions at room temperature and controlled acidic conditions. Total SP yields were low for both procedures, at 0.04% and 0.22%, respectively (Arata et al. 2015ARATA PX, QUINTANA I, CANELÓN DJ, VERA BE, COMPAGNONE RS & CIANCIA M. 2015. Chemical structure and anticoagulant activity of highly pyruvylated sulfated galactans from tropical green seaweeds of the order Bryopsidales. Carbohydr Polym 122: 376-386.). On the other hand, the yield obtained by proteolytic digestion employing papain was more efficient, at 30.5%.

Concerning the initial analysis of the sulfated constituents, each fraction was subjected to agarose gel electrophoresis and later stained with toluidine blue, allowing for the identification of SP in F23, F44 and F60. On the other hand, the F75 fraction, even when applied in large amounts, did not present the characteristic SP violet band in this electrophoresis system. The SP fractions extracted from H. opuntia presented different electrophoretic mobility patterns, although all bands were homogeneous and co-migrated close or slower than the heparin standard. The agarose gel electrophoresis initially suggested that SP with higher negative charges (high sulfation content) tend to display greater mobility, although the diaminopropane buffer present in the gel can interact with the sulfate groups of these molecules, resulting in negative charge neutralization, conformationally altering sulfate structure and exposure and, consequently, reducing molecule mobility. Therefore, this technique has been widely used in the study of carbohydrates, as it is capable of show marked differences in charge density between isolated SP fractions, which is why it is used in the identification of these molecules, allowing the preliminary analysis of extracted SP, indicating that different electrophoretic mobilities are associated to different structures. This would also explain why the F75 fraction is not visualized in this type of electrophoresis (Dietrich & Dietrich 1976DIETRICH CP & DIETRICH SMC. 1976. Electrophoretic behaviour of acidic mucopolysaccharides in diamine buffers. Analytical Biochemistry, v.70, n.02, p.645-647., Rodrigues et al. 2011RODRIGUES JAG, VANDERLEI ESO, BESSA EF, MAGALHÃES FA, PAULA RCM, LIMA V & BENEVIDES NMB. 2011. Anticoagulant Activity of a Sulfated Polysaccharide Isolated from the Green Seaweed Caulerpa cupressoides. Braz Arch Biol Technol 54(4): 691-700., Melo et al. 2012MELO KRT, ALMEIDA-LIMA J, GOMES DL, DANTAS-SANTOS N CAMARA RGB & ROCHA HAO. 2012. Caracterização e atividade anticoagulante de PS extraídos da alga marrom Dictyopteris justii. Holos 1: 29-40., Barbosa et al. 2019BARBOSA JS, PEREIRA MSS, MELO LFM, MEDEIROS MJC, PONTES DL, SCORTECCI KC & ROCHA HAO. 2019. In vitro immunostimulating activity of sulfated polysaccharides from Caulerpa cupressoides var. flabellata. Mar Drugs 17: 105.).

The anticoagulant activities of the four extracted fractions were tested in the activated partial thromboplastin time (APTT) and prothrombin time (PT) in vitro coagulation assays. These assays are used in clinical laboratory settings to determine the absence and/or functioning of coagulation factors present in the intrinsic and extrinsic (and/or common) pathways in the classic blood coagulation system model, respectively. Fractions F60 and F75 did not exhibit significant anticoagulant activity at any assessed concentration in the APTT assay (Figure 2a), while fractions F23 and F44, on the other hand, displayed similar anticoagulant effects, prolonging plasma clotting time by 300 seconds, at about 133 µg/mL using 1 µg/mL of unfractionated heparin (UFH) as a positive control. None of the precipitated fractions obtained from H. opuntia exhibited activity in the PT assay (Figure 2b).

Figure 1
Agarose gel electrophoresis of ethanol-precipitated SP fractions (F23, F44, F60 and F75) obtained from H. opuntia. UFH – Unfractionated Heparin; CS – Chondroitin sulfate; CE – Crude extract.
Figure 2
Anticoagulant activity of the precipitated SP fractions (F23, F44, F60 and F75) obtained from H. opuntia revealed by the APTT (2a) and PT (2b) assays. UFH – Unfractionated Heparin.

Purification of the F44 fraction obtained from H. opuntia and in vitro anticoagulant activity

SP purification of the precipitated F44 fraction and electrophoretic analysis

Although two fractions (F23 and F44) exhibited anticoagulant activity in the APTT assay, the F23 fraction resulted in a very low yield, as mentioned previously. Thus, only the F44 fraction was further investigated. The fraction was first subjected to ion exchange chromatography purification employing DEAE-cellulose under a linear NaCl gradient (0 to 4 M). The F44 fraction eluted in a single metachromatic peak (A525nm), termed P1F44 (Figure 3a). Furthermore, the presence of a neutral polysaccharide was observed by the exclusively positive reading noted at A490nm, by phenol/sulfuric acid analysis for total sugar, where a first major peak (LF44) was observed prior the start of the NaCl gradient, and a second peak was noted coinciding with the metachromatic peak, associated to SP (P1F44). The LF44 was discarded and the P1F44 fraction was subsequently dialyzed exhaustively against distilled water, filtered with Milli-Q water, freeze-dried and weighed, resulting in a 12% yield. This fraction was then submitted to agarose gel electrophoresis (Figure 3b), where an electrophoretic mobility pattern identical to the parent fraction was observed.

Figure 3
Purification of the F44 fraction by ion exchange chromatography employing DEAE-cellulose under a linear NaCl gradient (3a) with the detection of SP in DMB at A525nm and total hexoses at A490nm. Agarose gel electrophoresis (3b) of the F44 and P1F44 fractions obtained from H. opuntia. UFH – Unfractionated Heparin; CS – Chondroitin sulfate; CE – Crude extract.

In vitro anticoagulant activity in the APTT, PT and enzyme inhibition assays

The in vitro anticoagulant activity of the purified P1F44 fraction was reassessed by the APTT and PT assays. This fraction prolonged the clotting time in both in vitro assays (Figure 4). The P1F44 fraction doubled its activity as revealed by the APTT assay when compared to the precipitated fraction (F44), prolonging clotting time by 300 seconds at about 66 μg/mL of material, with an activity only 20-fold lower than the UFH. This behavior was expected, as the purification step is able to remove contaminants that do not display anticoagulant activity and, consequently, concentrate sample SP, increasing specific activity. It is important to highlight that sulfate presence and molecule position are extremely important for their anticoagulant activity. Furthermore, several studies concerning SP have demonstrated that their activities are also associated to their structural variability, given the wide diversification of their monosaccharide constitutions and types of glycosidic bonds, resulting in unique structural SP conformations with different patterns and anticoagulant potentials (Mendes et al. 2009MENDES SF ET AL. 2009. Sulfonation and anticoagulant activity of botryosphaeran from Botryosphaeria rhodina MAMB-05 grown on fructose. Int J Biol Macromol 45: 305-309., Fonseca et al. 2010FONSECA RJ, OLIVEIRA SN, POMIN VH, MECAWI AS, ARAUJO IG & MOURÃO PA. 2010. Effects of oversulfated and fucosylated chondroitin sulfates on coagulation. Thromb Haemost 103: 994-1004.).

Figure 4
Anticoagulant activity revealed by the APTT (4a) and PT (4b) assays for the purified SP fraction (P1F44) from H. opuntia compared to the fractions obtained by Arata et al. (2015)ARATA PX, QUINTANA I, CANELÓN DJ, VERA BE, COMPAGNONE RS & CIANCIA M. 2015. Chemical structure and anticoagulant activity of highly pyruvylated sulfated galactans from tropical green seaweeds of the order Bryopsidales. Carbohydr Polym 122: 376-386. (HO, PC, UF) and other authors. Caption: UFH – Unfractionated Heparin; HO – Halimeda opuntia; PC – Penicillus capitatus; UF – Udotea flabellum; MN – Monostroma nitidum; ML – Monostroma latissinum; CC – Codium cylindricum; CLC – Caulerpa cupressoides; CD – Codium divaricatum.

In addition, this fraction was also able to prolong the clotting time by 300 seconds at a concentration of 400 μg/mL via the PT assay, which is not common, as PS usually exhibit activity in only one pathway, intrinsic or extrinsic (Yang et al. 2002YANG J, DU Y, HUANG R, WAN Y & LI T. 2002. Chemical modification, characterization and structure-anticoagulant activity relationships of Chinese lacquer polysaccharides. Int J Biol Macromol 3: 55-62., Yoon et al. 2002YOON SJ, PEREIRA MS, PAVÃO MS, HWANG JK, PYUN YR & MOURÃO PA. 2002. The medicinal plant Porana volubilis contains polysaccharides with anticoagulant activity mediated by heparin cofactor II. Thromb Res 106: 51-58., 2007YOON SJ, PYUN YR, HWANG JK & MOURÃO PA. 2007. A sulfated fucan from the brown alga Laminaria cichorioides has mainly heparin cofactor II-dependent anticoagulant activity. Carbohydr Res 342: 2326-2330., Ronghua et al. 2003RONGHUA H, YUMIN D & JIANHONG Y. 2003. Preparation and in vitro anticoagulant activities of alginate sulfate and its quaterized derivatives. Carbohydr Polym 52: 19-24., Zhang et al. 2008ZHANG HJ, MAO WJ, FANG F, LI HY, SUN HH, CHEN Y & QI XH. 2008. Chemical characteristics and anticoagulant activities of a sulfated polysaccharide and its fragments from Monostroma latissimum. Carbohydr Polym 71: 428-434., Medeiros et al. 2008MEDEIROS VP, QUEIROZ KCS, CARDOSO ML, MONTEIRO GRG, OLIVEIRA FW, CHAVANTE SF, GUIMARAES LA, ROCHA HAO & LEITE EL. 2008. Sulfated galactofucan from Lobophora variegata: anticoagulant and anti-inflammatory properties. Biochemistry 73: 1018-1024., Silva et al. 2010SILVA FRF, DORE CMPG, MARQUES CT, NASCIMENTO MS, BENEVIDES NMB, ROCHA HAO, CHAVANTE SF & LEITE EL. 2010. Anticoagulant activity, paw edema and pleurisy induced carrageenan: Action of major types of commercial carrageenans. Carbohydr Polym 79: 26-33., Dore et al. 2013DORE CMPG ET AL. 2013. A sulfated polysaccharide, fucans, isolated from brown algae Sargassum vulgare with anticoagulant, antithrombotic, antioxidant and anti-inflammatory effects. Carbohydr Polym 91: 467-475.).

Figure 4 indicates the anticoagulant activities of several green algae SP as revealed by the APTT and PT assays. In general, the P1F44 fraction exhibits promising activity, since, in addition to activity detected via the APTT assay when compared to other algae, it also exhibits activity via the PT assay, while dual activity was noted only for Caulerpa cupressoides. It is important to note that the H. opuntia fraction assessed herein displayed better results than those reported by Arata et al. (2015)ARATA PX, QUINTANA I, CANELÓN DJ, VERA BE, COMPAGNONE RS & CIANCIA M. 2015. Chemical structure and anticoagulant activity of highly pyruvylated sulfated galactans from tropical green seaweeds of the order Bryopsidales. Carbohydr Polym 122: 376-386. as, in addition to being able to double the APTT assay results with an amount 6.5-fold lower than H. opuntia, it was also able to double the PT assay results, which was not reported by Arata et al. (Table I). This is probably due to the differences regarding the SP selective alcoholic extraction and precipitation processes applied herein. The algae Monostroma nitidum, Monostroma latissinum and Codium cylindricum exhibited better activity via the APTT assay, but no activity was observed via the PT assay at the concentrations evaluated, while it was possible to observe activity at the concentration of 400 μg/mL of the P1F44 fraction.

Table I
Comparison between the activities of different SP in the APTT and PT assays.

If, on the one hand, activity via the APTT assay is common, on the other hand, activity via the PT assay is unusual and is rarely reported in the literature. The double positive activity of the investigated fraction (aPTT and PT) is, thus, extraordinary, although some reports have indicated that SP extracted from the brown algae Ecklonia kurome (Nishino et al. 1991NISHINO T, AIZU Y & NAGUMO T. 1991. The relationship between the molecular weight and the anticoagulant activity of two types of fucan sulfates from the brown seaweed Ecklonia kurome. Agric Biol Chem 55: 791-796.), the green algae Codium dwarkense (Siddhanta et al. 1999SIDDHANTA AK, SHANMUGAM M, MODY KH, GOSWAMI AM & RAMAVAT BK. 1999. Sulphated polysaccharides of Codium dwarkense Boergs. from the west coast of India: chemical composition and blood anticoagulant activity. Int J Biol Macromol 26: 151-154.) and the fucoidan obtained from the brown seaweed Fucus vesiculosus (Azevedo et al. 2009AZEVEDO TCG, BEZERRA MEB, SANTOS MDGDL, SOUZA LA, MARQUES CT, BENEVIDES NMB & LEITE EL. 2009. Heparinoids algal and their anticoagulant, hemorrhagic activities and platelet aggregation. Biomed Pharmacother 63: 477-483.) exhibit dual activities.

To explore the anticoagulant mechanism of action, P1F44 was tested alongside factors isolated from the blood coagulation system. The results indicate that the P1F44 fraction interfered with the coagulation system by inhibiting thrombin (IIa) (Figure 5a) and factor Xa (Figure 5b) activities, both mediated by antithrombin (AT).

Figure 5
Inhibition of coagulation factors IIa (a) and Xa (b) by the P1F44 fraction obtained from H. opuntia in the presence of antithrombin. UFH – Unfractionated Heparin.

Table II displays the mean inhibitory concentration (EC50) of SP extracted from the red algae Botryocladia occidentalis and Gellidium crinale and from the sea urchins Strongylocentrotus purpuratus I, Strongylocentrotus purpuratus II and Arbacia lixula compared to the P1F44 fraction from H. opuntia with the reference drug (heparin) in relation to factors IIa and Xa in the presence of AT. All the listed SP exhibited lower EC50 in tests in which their anticoagulant potential via IIa/AT was compared to Xa/AT assays.

Table II
Inhibitory concentration of the P1F44 fraction obtained from H. opuntia compared to other polysaccharides of marine origin and to the heparin standard.

There is growing scientific evidence that suggests different interactions between SP and specific coagulation system proteins based on their different structural and sulfation patterns, forming a particular complex between the plasma inhibitor and the target protease. Some SP from marine organisms are capable of anticoagulant activity independent of interactions with coagulation proteins, as in the case of the sulfated galactan present in the seaweed Botryocladia occidenalis, which display activity through serpin-dependent and independent mechanisms. This SP exhibited excellent results in the in vitro APTT and PT coagulation assays and in enzymatic assays performed with purified proteases and serpins, displaying anti-factor IIa and anti-factor Xa activities mediated by AT. In addition to the examples previously cited in Table I, one SP isolated from the green alga Codium cylindricum by Matsubara et al. (2001)MATSUBARA K, MATSUBARA Y, BASIC A, LIAO ML, HORI K, MIYAZAWA K. 2001. Anticoagulant properties of a sulfated galactan preparation from a marine green alga, Codium cylindricum. Int J Biol Macromol 28: 395-399. also exhibited anticoagulant activity with a direct thrombin inhibitory mechanism, independent of the presence of AT and heparin cofactor II (Matsubara et al. 2001MATSUBARA K, MATSUBARA Y, BASIC A, LIAO ML, HORI K, MIYAZAWA K. 2001. Anticoagulant properties of a sulfated galactan preparation from a marine green alga, Codium cylindricum. Int J Biol Macromol 28: 395-399., Rodrigues et al. 2013RODRIGUES JAG, NETO EM, TEIXEIRA LAC, DE PAULA RCM, MOURAO PAS & BENEVIDES NMB. 2013. Structural features and inactivation of coagulation proteases of a sulfated polysaccharidic fraction from Caulerpa cupressoides var. lycopodium (Caulerpaceae, Chlorophyta). Acta Sci Technol 35: 611–619.). This possible multiplicity of anticoagulant action mechanisms should be better investigated to further understand the biotechnological potential of this class of molecule.

Chemical characterization of SP obtained from H. opuntia

The chemical characterizations of the F44 and P1F44 fractions as presented in Table III. The sulfate percentages in each fraction were 8.9% and 26%, respectively. The purification process may have removed sample interferences, generating better composition determinations. Sulfate contents, on the other hand, were higher than in the fraction reported by Arata et al. (2015)ARATA PX, QUINTANA I, CANELÓN DJ, VERA BE, COMPAGNONE RS & CIANCIA M. 2015. Chemical structure and anticoagulant activity of highly pyruvylated sulfated galactans from tropical green seaweeds of the order Bryopsidales. Carbohydr Polym 122: 376-386. of 21.7%. This difference corroborates the data obtained in the coagulation tests and shows the importance of the presence of sulfates for the anticoagulant activity of SP, hence the difference also observed for in vitro activity (ATTP and PT) between the SP obtained herein and by Arata et al. (2015)ARATA PX, QUINTANA I, CANELÓN DJ, VERA BE, COMPAGNONE RS & CIANCIA M. 2015. Chemical structure and anticoagulant activity of highly pyruvylated sulfated galactans from tropical green seaweeds of the order Bryopsidales. Carbohydr Polym 122: 376-386.. The uronic acid content detected herein (1,6%) was low when compared to other green algae, such as Ulva rigid (30%) and Caulerpa cupressoides (7.18%) (Rodrigues et al. 2013RODRIGUES JAG, NETO EM, TEIXEIRA LAC, DE PAULA RCM, MOURAO PAS & BENEVIDES NMB. 2013. Structural features and inactivation of coagulation proteases of a sulfated polysaccharidic fraction from Caulerpa cupressoides var. lycopodium (Caulerpaceae, Chlorophyta). Acta Sci Technol 35: 611–619., Adrien et al. 2019ADRIEN A, BONNET A, DUFOUR D, BAUDOUIN S, MAUGARD T & BRIDIAU N. 2019. Anticoagulant Activity of Sulfated Ulvan Isolated from the Green Macroalga Ulva rígida. Mar Drugs 17: 291.). No protein content was detected, indicating the efficiency of the applied extraction method.

Table III
Chemical characterization of SP from obtained from H.opuntia.

The monosaccharide composition analysis of the P1F44 fraction revealed the presence of mostly galactose (77.5%), followed by mannose (21.1%) and glucose (1.4%). Arata et al. (2015)ARATA PX, QUINTANA I, CANELÓN DJ, VERA BE, COMPAGNONE RS & CIANCIA M. 2015. Chemical structure and anticoagulant activity of highly pyruvylated sulfated galactans from tropical green seaweeds of the order Bryopsidales. Carbohydr Polym 122: 376-386. found the same constituent monosaccharides, with similar galactose contents (77.2%) and different mannose (9.8%) and glucose (6.1%) contents, also detecting the presence of ramnose (1%), fucose (1.3%), arabinose (2.3%) and xylose (2.1%). The results reported herein corroborate literature data, demonstrating that green algae exhibit more heterogeneous SP, usually rich in galactose, mannose, xylose, arabinose and glucose and/or uronic acids (Chevolot et al. 2001CHEVOLOT L, MULLOY B, RATISKOL J, FOUCAULT A & COLIEC-JOUAULT S. 2001. A disaccharide repeat unit is the major structure in fucoidans from two species of brown algae. Carbohyd Res 330: 529-535.).

Finally, an infrared (IR) spectroscopy assay was performed to analyze the primary structural characteristics of the SP present in the P1F44 fraction (Figure 6). The findings indicate characteristic carbohydrate bands, such as at 1250 cm⁻¹, confirming the bond between monomers due to typical C—O—C bond stretching. The most prominent band was observed between 3500 - 3300 cm⁻¹, corresponding to O—H bond stretching characteristic of monosaccharides.

Figure 6
Infrared spectrum of the P1F44 fraction extracted from H. opuntia.

Characteristic signs of sulfate grouping were also observed, such as S=O bond stretching between 1050 - 950 cm-1, C-O-S bond stretching at 850 cm-1 and C-O bond stretching at 600 cm-1, as exhibited in Table IV (Abreu 1997ABREU HS. 1997. Estimativa por infravermelho da concentração da unidade estrutural β-O-4 em ligninas de angiospermas tropicais. Quim Nova 20: 592-598., Silverstein et al. 2006SILVERSTEIN RM, WEBSTER FX & KIEMLE DJ. 2006. Identificação espectrométrica de compostos orgânicos. 7ª ed., Rio de Janeiro: LTC., Wu et al. 2013WU N, YE X, GUO X, LIAO N, YIN X, HU Y, SUN Y, LIU D & CHEN S. 2013. Depolymerization of fucosylated chondroitin sulfate from sea cucumber, Pearsonothuria graeffei, via 60 Co irradiation. Carbohydr Polym 93: 604-614.). Bands referring to C—H bond stretching around 2900 cm-1 and COO- or OH around 1650 cm-1 were also observed, the latter associated with the carboxylic acid group of uronic acid, confirming the data obtained in the chemical uronic acid analysis reported in Table I (Abreu 1997ABREU HS. 1997. Estimativa por infravermelho da concentração da unidade estrutural β-O-4 em ligninas de angiospermas tropicais. Quim Nova 20: 592-598., Silverstein et al. 2006SILVERSTEIN RM, WEBSTER FX & KIEMLE DJ. 2006. Identificação espectrométrica de compostos orgânicos. 7ª ed., Rio de Janeiro: LTC., Wu et al. 2013WU N, YE X, GUO X, LIAO N, YIN X, HU Y, SUN Y, LIU D & CHEN S. 2013. Depolymerization of fucosylated chondroitin sulfate from sea cucumber, Pearsonothuria graeffei, via 60 Co irradiation. Carbohydr Polym 93: 604-614.). The small band present at 1400 cm-1 corresponds to the carboxyl group belonging to pyruvic acid (Estevez et al. 2009ESTEVEZ JM, FERNANDEZ PV, KASULIN L, DUPREE P & CIANCA M. 2009. Chemical and in situ characterization macromolecular components of cell walls from the green seaweed Codium fragile. Glycobiology 19: 212-228.), thus confirming the data reported by Arata et al (2015), who describe the composition of SP present in H. opuntia as a sulphated pyruvated galactan.

Table IV
Infrared spectrum attributions of the SP obtained from H. opuntia.

CONCLUSIONS

A proteolytic extraction methodology followed by selective alcohol precipitation was effective in the extraction of four distinct and SP-rich fractions obtained from the green algae H. opuntia, two of which (F23 and F44) were active in the APTT (intrinsic coagulation system pathway) assay. F44 purification (P1F44) increased anticoagulant activity compared to the precipitated fraction (F44), doubling its intrinsic pathway action and also indicating extrinsic pathway activity (PT). The anticoagulant activity mechanism presented by P1F44 was shown to be dependent on both IIa and Xa factors, and more potent via IIa mediated by AT.

REFERENCES

  • ABREU HS. 1997. Estimativa por infravermelho da concentração da unidade estrutural β-O-4 em ligninas de angiospermas tropicais. Quim Nova 20: 592-598.
  • ADRIEN A, BONNET A, DUFOUR D, BAUDOUIN S, MAUGARD T & BRIDIAU N. 2017. Pilot production of ulvans from Ulva sp. and their effects on hyaluronan and collagen production in cultured dermal fibroblasts. Carbohydr Polym 157: 1306-1314.
  • ADRIEN A, BONNET A, DUFOUR D, BAUDOUIN S, MAUGARD T & BRIDIAU N. 2019. Anticoagulant Activity of Sulfated Ulvan Isolated from the Green Macroalga Ulva rígida. Mar Drugs 17: 291.
  • AFONSO A, MARQUES G, GONÇALVES A, BARROSO P, GONZALEZ A, RODRIGUES H & FERREIRA MJA. 2016. A terapêutica antitrombótica: atual e em desenvolvimento. Angiologia e Cirurgia Vascular 12: 170-179.
  • ALBAN S. 2005. The ‘precautionary principle’ as a guide for future drug development. Eur J Clin Investig 35: 33-44.
  • ANDERSON LO, BARROWCLIFFE TW, HOLMER E, JOHNSON EA & SIMS GFC. 1976. Anticoagulant properties of heparin fractionated by affinity chromatography on matrix-bound antithrombin-3 and by gel-filtration. Thromb Res 9: 575-580.
  • AQUINO RS, GRATIVOL C & MOURÃO PA. 2011. Rising from the sea: correlations between sulfated polysaccharides and salinity in plants. PLoS One 6: 1-7.
  • AQUINO RS, LANDEIRA-FERNANDEZ AM, VALENTE AP, ANDRADE LR & MOURÃO PA. 2005. Occurrence of sulfated galactans in marine angiosperms: evolutionary implications. Glycobiology 15: 11-20.
  • ARATA PX, QUINTANA I, CANELÓN DJ, VERA BE, COMPAGNONE RS & CIANCIA M. 2015. Chemical structure and anticoagulant activity of highly pyruvylated sulfated galactans from tropical green seaweeds of the order Bryopsidales. Carbohydr Polym 122: 376-386.
  • AZEVEDO TCG, BEZERRA MEB, SANTOS MDGDL, SOUZA LA, MARQUES CT, BENEVIDES NMB & LEITE EL. 2009. Heparinoids algal and their anticoagulant, hemorrhagic activities and platelet aggregation. Biomed Pharmacother 63: 477-483.
  • BARBOSA JS, PEREIRA MSS, MELO LFM, MEDEIROS MJC, PONTES DL, SCORTECCI KC & ROCHA HAO. 2019. In vitro immunostimulating activity of sulfated polysaccharides from Caulerpa cupressoides var. flabellata. Mar Drugs 17: 105.
  • BERTEAU O & MULLOY B. 2003. Sulfated fucans, fresh perspectives: structures, functions, and biological properties of sulfated fucans and an overview of enzymes active toward this class of polysaccharide. Glycobiology 13: 29R-40R.
  • BLUNT JW, COPP BR, KEYZERS RA, MUNRO MH & PRINSEP MR. 2015. Marine natural products. Nat Prod Rep 32: 116-211.
  • BRADFORD MM. 1976. A rapid and sensitive method for the quantification of microgram quantities of protein utilizing the principle of protein-dye binding. Anal Biochem 72: 248-254.
  • CALUMPONG HP, WEST J & SEAWEEDS MG. 2017. In the First Global Integrated Marine Assessment: World Ocean Assessment I; United Nations, Cambridge University Press: Cambridge, UK.
  • CAO S, HE X, QIN L, HE M, YANG Y, LIU Z & MAO W. 2019. Anticoagulant and antithrombotic properties in vitro and in vivo of a novel sulfated polysaccharide from marine green alga Monostroma nitidum. Mar Drugs 17: 247.
  • CHEVOLOT L, MULLOY B, RATISKOL J, FOUCAULT A & COLIEC-JOUAULT S. 2001. A disaccharide repeat unit is the major structure in fucoidans from two species of brown algae. Carbohyd Res 330: 529-535.
  • CHOPIN T & WHALEN E. 1993. A new rapad method for carrageenan indentification by FT IR diffuse reflectance spectroscopy directly on dried, ground algal material. Carbohydr Res 246: 51-59.
  • CORREIA-DA-SILVA M, SOUSA E, MARQUES F & PINTO MM. 2013. Estado da arte na terapêutica anticoagulante: Novas abordagens. Acta Farm Port 2: 5-18.
  • COSTA LS ET AL. 2010. Biological activities of sulfated polysaccharides from tropical seaweeds. Biomed Pharmacother 64: 21-28.
  • DAMONTE E ET AL. 1994. Antiviral activity of a sulphated polysaccharide from the red seaweed Nothogenia fastigiata. Biochem Pharmacol 47: 2187-2192.
  • DANTAS-SANTOS N ET AL. 2012. Freshwater plants synthesize sulfated polysaccharides: heterogalactans from water hyacinth (Eicchornia crassipes). Int J Mol Sci 13: 961-976.
  • DIETRICH CP & DIETRICH SMC. 1976. Electrophoretic behaviour of acidic mucopolysaccharides in diamine buffers. Analytical Biochemistry, v.70, n.02, p.645-647.
  • DISCHE Z. 1947. A specific color reaction for glucuronic acid. J Biol Chem 171: 725-730.
  • DODGSON KS & PRICE RG. 1962. A note on the determination of the ester sulphate content of sulphated polysaccharides. Biochem J 84: 106-110.
  • DORE CMPG ET AL. 2013. A sulfated polysaccharide, fucans, isolated from brown algae Sargassum vulgare with anticoagulant, antithrombotic, antioxidant and anti-inflammatory effects. Carbohydr Polym 91: 467-475.
  • ESTEVEZ JM, FERNANDEZ PV, KASULIN L, DUPREE P & CIANCA M. 2009. Chemical and in situ characterization macromolecular components of cell walls from the green seaweed Codium fragile. Glycobiology 19: 212-228.
  • FARIAS EHC, POMIN VH, VALENTE AP, NADER HB, ROCHA HAO & MOURÃO PAS. 2008. A preponderantly 4-sulfated, 3-linked galactan from the green alga Codium isth-mocladum. Glycobiology 18: 250-259.
  • FARIAS WRL, VALENTE AP, PEREIRA MS & MOURÃO PAS. 2000. Structure and anticoagulant activity of sulfacted galactans. J Biol Chem 275: 29299-29307.
  • FARNDALE RW, BUTTLE DJ & BARRETT AJ. 1986. Improved quantitation and discrimination of sulphated glycosaminoglycans by use of dimethylmethylene blue. Biochim Biophys Acta 883: 173-177.
  • FERDOUSE F, LOVESTAD HOLDT S, SMITH R, MURUA P & YANG Z. 2018. The Global Status of Seaweed Production, Trade and Utilization; FAO Globefish Research Programme; Food and Agriculture Organization of the United Nations: Rome, Italy.
  • FLATO UAP, BUHATEM T, MERLUZZI T & BIANCO ACM. 2010. Novos anticoagulantes em cuidados intensivos. Rev Bras Ter Intensiva 23: 68-77.
  • FONSECA RJ, OLIVEIRA SN, POMIN VH, MECAWI AS, ARAUJO IG & MOURÃO PA. 2010. Effects of oversulfated and fucosylated chondroitin sulfates on coagulation. Thromb Haemost 103: 994-1004.
  • GLAUSER BF, REZENDE RM, MELO FR, PEREIRA MS, FRANCISCHETTI IMB, MONTEIRO RQ, REZAIE AR & MOURÃO PAS. 2009. Anticoagulant activity of a sulfated galactan: Serpin-independent effect and specific interaction with factor Xa. Thromb Haemost 102(06): 1183-1193.
  • HU Y, YU GL, ZHAO XX, WANG YF, SUN XC, JIAO GL, ZHAO X & CHAI WG. 2012. Structural characterization of natural ideal 6-O-sulfated agarose from red alga Gloiopeltis furcata. Carbohydr Polym 89: 883-889.
  • KIRCHER HW. 1960. Gas-liquid partition chromatography of methylated sugars. Anal Chem 32: 1103-1106.
  • KUSAYKIN MI, CHIZHOV AO, GRACHEV AA, ALEKSEEVA SA, BAKUNINA IY, NEDASHKOVSKAYA OI, SOVA VV & ZVYAGINTSEVA TNA. 2006. A comparative study of specificity of fucoidanases from marine microorganisms and invertebrates. J Appl Phyc 18: 143-147.
  • LI H ET AL. 2011. Structural characterization of an anticoagulant-active sulfated polysaccharide isolated from green alga Monostroma latissimum. Carbohydr Polym 85: 394-400.
  • LI N ET AL. 2015. Structural characterization and anticoagulant activity of a sulfated polysaccharide from the green alga Codium divaricatum. Carbohydr Polym 121: 175-182.
  • LORGA FILHO AM ET AL. 2013. Diretrizes brasileiras de antiagregantes plaquetários e anticoagulantes em cardiologia. Arq Bras Cardiol 101: 3.
  • MATSUBARA K, MATSUBARA Y, BASIC A, LIAO ML, HORI K, MIYAZAWA K. 2001. Anticoagulant properties of a sulfated galactan preparation from a marine green alga, Codium cylindricum. Int J Biol Macromol 28: 395-399.
  • MEDEIROS VP, QUEIROZ KCS, CARDOSO ML, MONTEIRO GRG, OLIVEIRA FW, CHAVANTE SF, GUIMARAES LA, ROCHA HAO & LEITE EL. 2008. Sulfated galactofucan from Lobophora variegata: anticoagulant and anti-inflammatory properties. Biochemistry 73: 1018-1024.
  • MELO EI, PEREIRA MS, CUNHA RS, LEME MPS & MOURÃO PAS. 2008. Heparin quality control in the Brazilian market: implications in the cardiovascular surgery. Braz J Cardiovasc Surg 23: 169-174.
  • MELO KRT, ALMEIDA-LIMA J, GOMES DL, DANTAS-SANTOS N CAMARA RGB & ROCHA HAO. 2012. Caracterização e atividade anticoagulante de PS extraídos da alga marrom Dictyopteris justii. Holos 1: 29-40.
  • MENDES SF ET AL. 2009. Sulfonation and anticoagulant activity of botryosphaeran from Botryosphaeria rhodina MAMB-05 grown on fructose. Int J Biol Macromol 45: 305-309.
  • MOURÃO PAS. 2004. Estrutura versus atividade biológica de polissacarídeos sulfatados. In.: XXVI Reunião Anual sobre Evolução, Sistemática e Ecologia Micromolecular, 2004, Niterói.
  • MOURÃO PAS. 2015. Perspective on the use of sulfated polysaccharides from marine organisms as a source of new antithrombotic drugs. Mar Drugs 13: 2770-2784.
  • MOURÃO PAS & PEREIRA MS. 2000. Searching from alternatives to heparin sulfated fucans from marine invertebrates. Trends Cardiovasc Med 9: 225-232.
  • NADER HB, LOPES CC, ROCHA HA, SANTOS EA, DIETRICH CP. 2004. Heparins and heparinoids: occurrence, structure and mechanism of antithrombotic and hemorrhagic activities. Curr Pharm Des 10: 951-966.
  • NISHINO T, AIZU Y & NAGUMO T. 1991. The relationship between the molecular weight and the anticoagulant activity of two types of fucan sulfates from the brown seaweed Ecklonia kurome. Agric Biol Chem 55: 791-796.
  • OLSON ST, BJORK I, SHEFFER R, CRAIG PA, SHORE JD & CHOAY J. 1992. Role of the antithrombin-binding pentasaccharide in heparina acceleration conformational change contribution to heparin rate enhancement. J Bio Chem 267: 12528-12538.
  • PAN J, QIAN Y, ZHOU XD, PAZANDAK A, FRAZIER SB, WEISER P, LU H & ZHANG LJ. 2010. Oversulfated chondroitin sulfate is not the sole contaminant in heparin. Nat Biotechnol 28: 203-207.
  • PEREIRA MG, BENEVIDES NM, MELO MR, VALENTE AP, MELO FR & MOURÃO PA. 2005. Structure and anticoagulant activity of a sulfated galactan from the red alga, Gelidium crinale. Is there a specific structural requirement for the anticoagulant action? Carbohydr Res 340: 2015-2023.
  • PEREIRA MS, MELO FR & MOURÃO PA. 2002. Is there a correlation between structure and anticoagulant action of sulfated galactans and sulfated fucans? Glycobiology 12: 573-580.
  • PEREIRA MS, MULLOY B & MOURÃO PA. 1999. Structure and anticoagulant activity of sulfated fucans: Comparison between the regular repetitive and linear fucans from echinoderms with the more heterogeneous and branched polymers from brown algae. J Biol Chem 274: 7656-7667.
  • POMIN VH & MOURÃO PA. 2014. Specific sulfation and glycosylation - a structural combination for the anticoagulation of marine carbohydrates. Front Cell Infect Mi 4: 1-8.
  • RODRIGUES JAG, NETO EM, TEIXEIRA LAC, DE PAULA RCM, MOURAO PAS & BENEVIDES NMB. 2013. Structural features and inactivation of coagulation proteases of a sulfated polysaccharidic fraction from Caulerpa cupressoides var. lycopodium (Caulerpaceae, Chlorophyta). Acta Sci Technol 35: 611–619.
  • RODRIGUES JAG, VANDERLEI ESO, BESSA EF, MAGALHÃES FA, PAULA RCM, LIMA V & BENEVIDES NMB. 2011. Anticoagulant Activity of a Sulfated Polysaccharide Isolated from the Green Seaweed Caulerpa cupressoides. Braz Arch Biol Technol 54(4): 691-700.
  • RONGHUA H, YUMIN D & JIANHONG Y. 2003. Preparation and in vitro anticoagulant activities of alginate sulfate and its quaterized derivatives. Carbohydr Polym 52: 19-24.
  • SATTARI M & LOWENTHAL DT. 2011. Novel Oral Anticoagulants in Development: Dabigatran, Rivaroxaban, and Apixaban. Am J Ther 18: 332-338.
  • SIDDHANTA AK, SHANMUGAM M, MODY KH, GOSWAMI AM & RAMAVAT BK. 1999. Sulphated polysaccharides of Codium dwarkense Boergs. from the west coast of India: chemical composition and blood anticoagulant activity. Int J Biol Macromol 26: 151-154.
  • SILVA FRF, DORE CMPG, MARQUES CT, NASCIMENTO MS, BENEVIDES NMB, ROCHA HAO, CHAVANTE SF & LEITE EL. 2010. Anticoagulant activity, paw edema and pleurisy induced carrageenan: Action of major types of commercial carrageenans. Carbohydr Polym 79: 26-33.
  • SILVA PM & NETO PC. 2009. Atividades biológicas de extratos de algas marinhas brasileiras – Master’s degree. Universidade de São Paulo, São Paulo, SP. (Unpublished).
  • SILVERSTEIN RM, WEBSTER FX & KIEMLE DJ. 2006. Identificação espectrométrica de compostos orgânicos. 7ª ed., Rio de Janeiro: LTC.
  • SOEDA S, ISHIDA S, HONDA O, SHIMENO H & NAGAMATSU A. 1994. Amineted fucoidan promotes the invasion of 3 LL cells through reconstituted basement membrane: Its possible mechanism of action. Cancer Lett 85: 133-138.
  • SOMMERS CD, YE H, KOLINSKI RE, NASR M, BUHSE LF, AL-HAKIM A & KEIRE DA. 2011. Characterization of currently marketed heparin products: analysis of molecular weight and heparinase-I digest patterns. Anal Bioanal Chem 401: 2445-2454.
  • SUJIAN C, XIAOXI H, LING Q, MEIJIA H, YAJING Y, ZHICHUN L & WENJUN M. 2019. Anticoagulant and Antithrombotic Properties in vitro and in vivo of a Novel Sulfated Polysaccharide from Marine Green Alga Monostroma nitidum. Mar Drugs 17: 247.
  • VAN DER WALL SJ, KLOK FA, DEN EXTER PL, BARRIOS D, MORILLO R, CANNEGIETER SC, JIMENEZ D & HUISMAN MV. 2017. Higher Adherence to Treatment With Low-Molecular-Weight-Heparin Nadroparin Than Enoxaparin Because of Side Effects in Cancer-Associated Venous Thromboembolism. HemaSphere 2:1.
  • VIEIRA LAP, FREITAS ALP, FEITOSA JPA, SILVA DC & VIANA GSB. 2004. The alga Bryothamnion seaforthii contains carbohydrates with antinociceptive activity. Braz J Med Bio Res 37: 1071-1079.
  • WHITE RH. 2006. The epidemiology of venous thromboembolism. Thromb Haemost 21: 23-29.
  • WHO. 2017. Cardiovascular diseases. Available at: https://www.who.int/health-topics/cardiovascular-diseases Accessed on: 14/04/2020.
    » https://www.who.int/health-topics/cardiovascular-diseases
  • WU N, YE X, GUO X, LIAO N, YIN X, HU Y, SUN Y, LIU D & CHEN S. 2013. Depolymerization of fucosylated chondroitin sulfate from sea cucumber, Pearsonothuria graeffei, via 60 Co irradiation. Carbohydr Polym 93: 604-614.
  • YANG J, DU Y, HUANG R, WAN Y & LI T. 2002. Chemical modification, characterization and structure-anticoagulant activity relationships of Chinese lacquer polysaccharides. Int J Biol Macromol 3: 55-62.
  • YOON SJ, PEREIRA MS, PAVÃO MS, HWANG JK, PYUN YR & MOURÃO PA. 2002. The medicinal plant Porana volubilis contains polysaccharides with anticoagulant activity mediated by heparin cofactor II. Thromb Res 106: 51-58.
  • YOON SJ, PYUN YR, HWANG JK & MOURÃO PA. 2007. A sulfated fucan from the brown alga Laminaria cichorioides has mainly heparin cofactor II-dependent anticoagulant activity. Carbohydr Res 342: 2326-2330.
  • YOSHIDA RDA, YOSHIDA WB & ROLLO HA. 2011. Novos anticoagulantes para a profilaxia do tromboembolismo venoso em cirurgias ortopédicas de grande porte. J Vasc Brasil 145-153.
  • ZHANG HJ, MAO WJ, FANG F, LI HY, SUN HH, CHEN Y & QI XH. 2008. Chemical characteristics and anticoagulant activities of a sulfated polysaccharide and its fragments from Monostroma latissimum. Carbohydr Polym 71: 428-434.
  • ZHANG Q, LI N, ZHOU G, LU X, XU Z & LI Z. 2003. In vivo antioxidant activity of polysaccharide fraction from Porphyra haitanesis (Rhodophyta) in aging mice. Pharmacol Res 48: 151-155.

Publication Dates

  • Publication in this collection
    17 Feb 2023
  • Date of issue
    2023

History

  • Received
    14 July 2021
  • Accepted
    29 Nov 2021
Academia Brasileira de Ciências Rua Anfilófio de Carvalho, 29, 3º andar, 20030-060 Rio de Janeiro RJ Brasil, Tel: +55 21 3907-8100 - Rio de Janeiro - RJ - Brazil
E-mail: aabc@abc.org.br