[go: up one dir, main page]

 
 

Marine Polysaccharides

A topical collection in Marine Drugs (ISSN 1660-3397).

Viewed by 897015
Printed Edition Available!
A printed edition of this Special Issue is available here.

Editor


E-Mail Website
Collection Editor
Institute for Polymers, Composites and Biomaterials, CNR, via Campi Flegrei 34, 80078 Pozzuoli, Italy
Interests: macromolecular chemistry; natural polysaccharides; drug delivery; nanoparticles surface modification; active targeting
Special Issues, Collections and Topics in MDPI journals

Topical Collection Information

Dear Colleagues,

Biopolymers, as natural polysaccharides, are considered benign polymers for what concerns the environment. This is not a new invention, but at best a renaissance: the first type of polymers used by human kind were animal hides, cellulose, silk, wool. Among benefits of natural occurring biopolymers there are potential biocompatibility, renewable resources, low processing costs, tailoring of structure by genetic manipulation, and, as said, environmentally compatibility. Limits are, sometimes, premature degradation and high production costs due to the very high purity required for medical uses. Polysaccharides are not drugs by themselves, but their use in pharmaceutical field, for example as drug carriers or antimicrobial, anti-inflammatory or anticoagulant agents, is increasingly promising. Marine polysaccharides include chitin, chitosan, alginate, agar and carrageenans. Chitosan is a cationic carbohydrate biopolymer derived from chitin, the second most abundant polysaccharides present in nature after cellulose. The main sources of chitin are the shell wastes of shrimps, lobsters and crabs. For its characteristics, chitosan founds particular application as non viral vector in gene delivery. Films from chitosan are very tough and long lasting. Alginates derive from seaweed extraction (pheophyceae), and are mainly used in drug delivery and as hydrogels for immobilizing cells and enzymes, due to the mild conditions of cross-linking through bivalent cations (Ca2+). Agar (or agar-agar) and carrageenans are linear polysaccharides from red seaweeds. They are highly reactive chemically and are peculiar for thermoreversible gel formation. Exopolysaccharides (EPS), substantial components of the extracellular matrix of many cells of marine origin, also have to be mentioned for their potential interest in pharmaceuticals, and new EPS producing bacteria, particularly from extreme marine environments, are being isolated.
The possibility of chemical modification, blending and addition of biodegradable additives allows to tailor the final properties of polysaccharides and opens the doors to wider applications, particularly in pharmaceutical area. This collection is intended to explore any new potentiality of marine polysaccharides, as those above mentioned, deriving from chemical or chemical-physical modifications, and the scaling-up of their pharmaceutical applications.

Dr. Paola Laurienzo
Collection Editor

Manuscript Submission Information

Manuscripts for the topical collection can be submitted online at www.mdpi.com by registering and logging in to this website. Once you are registered, click here to go to the submission form. All papers will be peer-reviewed. Accepted papers will be published continuously in the journal (as soon as accepted) and will be listed together on this website. The topical collection considers regular research articles, short communications and review articles. A guide for authors and other relevant information for submission of manuscripts is available on the Instructions for Authors page.

Please visit the Instructions for Authors page before submitting a manuscript. The article processing charge (APC) for publication in this open access journal is 2900 CHF (Swiss Francs).


Keywords

  • chitosan
  • alginate
  • agar
  • carrageenans
  • exopolysaccharides
  • chemical modification
  • drug delivery
  • gene delivery

Published Papers (102 papers)

2024

Jump to: 2022, 2021, 2020, 2019, 2018, 2017, 2016, 2015, 2014, 2013, 2012, 2011, 2010

18 pages, 7206 KiB  
Article
Extraction Optimization and Anti-Tumor Activity of Polysaccharides from Chlamydomonas reinhardtii
by Zhongwen Liang, Lan Xiong, Ying Zang, Zhijuan Tang, Zhenyu Shang, Jingyu Zhang, Zihan Jia, Yanting Huang, Xiaoyu Ye, Hongquan Liu and Mei Li
Mar. Drugs 2024, 22(8), 356; https://doi.org/10.3390/md22080356 - 2 Aug 2024
Viewed by 906
Abstract
Chlamydomonas reinhardtii polysaccharides (CRPs) are bioactive compounds derived from C. reinhardtii, yet their potential in cancer therapy remains largely unexplored. This study optimized the ultrasound-assisted extraction conditions using response surface methodology and proceeded with the isolation and purification of these polysaccharides. The [...] Read more.
Chlamydomonas reinhardtii polysaccharides (CRPs) are bioactive compounds derived from C. reinhardtii, yet their potential in cancer therapy remains largely unexplored. This study optimized the ultrasound-assisted extraction conditions using response surface methodology and proceeded with the isolation and purification of these polysaccharides. The optimal extraction conditions were identified as a sodium hydroxide concentration of 1.5%, ultrasonic power of 200 W, a solid-to-liquid ratio of 1:25 g/mL, an ultrasonic treatment time of 10 min, and a water bath duration of 2.5 h, yielding an actual extraction rate of 5.71 ± 0.001%, which closely aligns with the predicted value of 5.639%. Infrared analysis revealed that CRP-1 and CRP-2 are α-pyranose structures containing furoic acid, while CRP-3 and CRP-4 are β-pyranose structures containing furoic acid. Experimental results demonstrated that all four purified polysaccharides inhibited the proliferation of cervical (HeLa) hepatoma (HepG-2) and colon (HCT-116) cancer cells, with CRP-4 showing the most significant inhibitory effect on colon cancer and cervical cancer, achieving inhibition rates of 60.58 ± 0.88% and 40.44 ± 1.44%, respectively, and significantly reducing the migration of HeLa cells. DAPI staining confirmed that the four purified polysaccharides inhibit cell proliferation and migration by inducing apoptosis in HeLa cells. CRP-1 has the most significant inhibitory effect on the proliferation of liver cancer cells. This study not only elucidates the potential application of C. reinhardtii polysaccharides in cancer therapy but also provides a scientific basis for their further development and utilization. Full article
Show Figures

Figure 1

Figure 1
<p>Single-factor test. (<b>a</b>) NaOH mass fraction. (<b>b</b>) Solid-to-liquid ratio. (<b>c</b>) Ultrasound time. (<b>d</b>) Ultrasonic power. (<b>e</b>) Water bath time. The data are expressed as mean ± standard deviation of at least 3 independent experiments (3 replicates).</p>
Full article ">Figure 2
<p>The effects of mass fraction of NaOH, ultrasonic power, and solid–liquid ratio on the extraction rate of polysaccharide were studied. A: mass fraction of NaOH; B: ultrasonic power; C: solid–liquid ratio.</p>
Full article ">Figure 3
<p>The elution profile of crude polysaccharide from <span class="html-italic">C. reinhardtii</span> on DEAE-cellulose anion exchange chromatography column. The black curve represents the elution curve of polysaccharides measured by the anthrone sulfuric acid method. The red curve represents different concentrations of NaCl.</p>
Full article ">Figure 4
<p>Ultraviolet spectra of <span class="html-italic">C. reinhardtii</span> polysaccharides. There is no absorption peak at 260 nm or 280 nm.</p>
Full article ">Figure 5
<p>Infrared spectra of four purified polysaccharides. (<b>a</b>) CRP-1. (<b>b</b>) CRP-2. (<b>c</b>) CRP-3. (<b>d</b>) CRP-4.</p>
Full article ">Figure 6
<p>Toxic effect of <span class="html-italic">C. reinhardtii</span> polysaccharide on Vero cells. There was no cytotoxicity to Vero cells after treatment with CRP-1, CRP-2, CRP-3 and CRP-4 for 24 h. The data are expressed as mean ± standard deviation of at least 3 independent experiments (3 replicates).</p>
Full article ">Figure 7
<p>Effects of <span class="html-italic">C. reinhardtii</span> polysaccharide on the proliferation of different tumor cells. HCT-116, HeLa, and HepG2 cells were treated with four purified polysaccharides for 24 h, and the proliferation of tumor cells was inhibited. (<b>a</b>) HCT-116 cells. (<b>b</b>) HeLa cells. (<b>c</b>) HepG-2 cells. The data are expressed as mean ± standard deviation of at least 3 independent experiments (3 replicates).</p>
Full article ">Figure 8
<p>The effect of CRP-4 on the morphology of cervical cancer HeLa cells. (<b>a</b>) Control; (<b>b</b>) 0.1 mg/mL CRP-4; (<b>c</b>) 0.3 mg/mL CRP-4; (<b>d</b>) 0.5 mg/mL CRP-4; (<b>e</b>) 1 mg/mL CRP-4. After HeLa cells were treated with CRP-4 for 24 h, HeLa cells showed different morphological changes compared with the control group, including morphological abnormalities, increased cell gap, cell shrinkage, blurred contour, and darkening of luster. Microscope magnification is 200×.</p>
Full article ">Figure 9
<p>Effect of <span class="html-italic">C. reinhardtii</span> polysaccharide on migration of cervical cancer HeLa cells. Compared with the control group, the migration ability of HeLa cells was inhibited after treatment with four purified polysaccharides for 24 h. Microscope magnification is 100×.</p>
Full article ">Figure 10
<p>The effects of four purified polysaccharides on the migration of cervical cancer HeLa cells. The scratches were quantitatively analyzed by Image J 1.8 software and similar results were obtained from two other independent experiments. The data are expressed as mean ± standard deviation of at least 3 independent experiments (3 replicates). An asterisk (*) indicates a significant difference (<span class="html-italic">p</span> &lt; 0.05), while three asterisks (***) indicate a highly significant difference (<span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 11
<p>DAPI staining was used to observe the effect of <span class="html-italic">C. reinhardtii</span> polysaccharide on the apoptosis morphology of cervical cancer HeLa cells. Red arrows represent chromatin condensation, nuclear fragmentation, formation of apoptotic bodies, bright blue aggregates or dots.</p>
Full article ">
27 pages, 14164 KiB  
Article
A Comparative Analysis of the Anti-Tumor Activity of Sixteen Polysaccharide Fractions from Three Large Brown Seaweed, Sargassum horneri, Scytosiphon lomentaria, and Undaria pinnatifida
by Lin Song, Yunze Niu, Ran Chen, Hao Ju, Zijian Liu, Bida Zhang, Wancui Xie and Yi Gao
Mar. Drugs 2024, 22(7), 316; https://doi.org/10.3390/md22070316 - 16 Jul 2024
Viewed by 784
Abstract
Searching for natural products with anti-tumor activity is an important aspect of cancer research. Seaweed polysaccharides from brown seaweed have shown promising anti-tumor activity; however, their structure, composition, and biological activity vary considerably, depending on many factors. In this study, 16 polysaccharide fractions [...] Read more.
Searching for natural products with anti-tumor activity is an important aspect of cancer research. Seaweed polysaccharides from brown seaweed have shown promising anti-tumor activity; however, their structure, composition, and biological activity vary considerably, depending on many factors. In this study, 16 polysaccharide fractions were extracted and purified from three large brown seaweed species (Sargassum horneri, Scytosiphon lomentaria, and Undaria pinnatifida). The chemical composition analysis revealed that the polysaccharide fractions have varying molecular weights ranging from 8.889 to 729.67 kDa, and sulfate contents ranging from 0.50% to 10.77%. Additionally, they exhibit different monosaccharide compositions and secondary structures. Subsequently, their anti-tumor activity was compared against five tumor cell lines (A549, B16, HeLa, HepG2, and SH-SY5Y). The results showed that different fractions exhibited distinct anti-tumor properties against tumor cells. Flow cytometry and cytoplasmic fluorescence staining (Hoechst/AO staining) further confirmed that these effective fractions significantly induce tumor cell apoptosis without cytotoxicity. qRT-RCR results demonstrated that the polysaccharide fractions up-regulated the expression of Caspase-3, Caspase-8, Caspase-9, and Bax while down-regulating the expression of Bcl-2 and CDK-2. This study comprehensively compared the anti-tumor activity of polysaccharide fractions from large brown seaweed, providing valuable insights into the potent combinations of brown seaweed polysaccharides as anti-tumor agents. Full article
Show Figures

Figure 1

Figure 1
<p>Fourier transform infrared (FT-IR) spectra of polysaccharide fractions from three types of seaweed. (<b>A</b>) <span class="html-italic">Sargassum horneri</span> polysaccharide (SHP) fractions, (<b>B</b>) <span class="html-italic">Scytosiphon lomentaria</span> polysaccharide (SLP) fractions, and (<b>C</b>) <span class="html-italic">Undaria pinnatifida</span> polysaccharide (UPP) fractions.</p>
Full article ">Figure 2
<p>The triple-helix conformation analysis of polysaccharide fractions from three types of seaweed. (<b>A</b>) <span class="html-italic">S. horneri</span> polysaccharide (SHP) fractions, (<b>B</b>) <span class="html-italic">S. lomentaria</span> polysaccharide (SLP) fractions, and (<b>C</b>) <span class="html-italic">U. pinnatifida</span> polysaccharide (UPP) fractions.</p>
Full article ">Figure 3
<p>The inhibitory effects of seven fractions of the <span class="html-italic">S. horneri</span> polysaccharide (SHP) at different concentrations (100, 200, and 400 μg/mL) on the proliferation of (<b>A</b>) A549, (<b>B</b>) B16, (<b>C</b>) HeLa, (<b>D</b>) HepG<sub>2</sub>, and (<b>E</b>) SH-SY5Y cells (mean  ±  standard deviation). The letters a–c indicate significant differences (<span class="html-italic">p</span>  &lt;  0.05) in the inhibitory rate of each fraction.</p>
Full article ">Figure 4
<p>The inhibitory effects of five fractions of the <span class="html-italic">S. lomentaria</span> polysaccharide (SLP) at different concentrations (100, 200, and 400 μg/mL) on the proliferation of (<b>A</b>) A549, (<b>B</b>) B16, (<b>C</b>) HeLa, (<b>D</b>) HepG<sub>2</sub>, and (<b>E</b>) SH-SY5Y cells (mean  ±  SD). The letters a–c indicate significant differences (<span class="html-italic">p</span>  &lt;  0.05) in the inhibitory rate of each fraction.</p>
Full article ">Figure 5
<p>The inhibitory effects of four fractions of the <span class="html-italic">U. pinnatifida</span> polysaccharide (UPP) at different concentrations (100, 200, and 400 μg/mL) on the proliferation of (<b>A</b>) A549, (<b>B</b>) B16, (<b>C</b>) HeLa, (<b>D</b>) HepG<sub>2</sub>, and (<b>E</b>) SH-SY5Y cells (mean  ±  SD). The letters a–c indicate significant differences (<span class="html-italic">p</span>  &lt;  0.05) in the inhibitory rate of each fraction.</p>
Full article ">Figure 6
<p>The effect of polysaccharide fractions on tumor cells through flow cytometry analysis. (<b>A</b>) Control group of HeLa cells, (<b>B</b>) 200 μg/mL SHP-1.7 treatment, (<b>C</b>) control group of B16 cells, (<b>D</b>) 400 μg/mL SHP-1.0 treatment, (<b>E</b>) control group of HepG<sub>2</sub> cells, (<b>F</b>) 200 μg/mL SLP-1.3 treatment, (<b>G</b>) control group of A549 cells, (<b>H</b>) 100 μg/mL SLP-1.3 treatment, (<b>I</b>) control group of SH-SY5Y cells, and (<b>J</b>) 100 μg/mL UPP-0.5 treatment. The proportion of each area is displayed as a percentage.</p>
Full article ">Figure 7
<p>The effects of four polysaccharide fractions at different concentrations (200, 400, and 800 μg/mL) on the proliferation of macrophage (RAW264.7) cells (mean  ±  SD). The letters a–c indicate a significant difference (<span class="html-italic">p</span>  &lt;  0.05) between the proliferation rate of each fraction.</p>
Full article ">Figure 8
<p>Hoechst 33342 staining in five cancer cell lines treated with the most effective fractions (SHP-1.7/HeLa, SHP-1.0/B16, SLP-1.3/A549, SLP-1.3/HepG<sub>2</sub>, and UPP-0.5/SH-SY5Y) for 48 h. The left side represents the control group, and the right side represents the experimental group.</p>
Full article ">Figure 9
<p>Acridine Orange (AO) staining in five cancer cell lines treated with the most effective fractions (SHP-1.7/HeLa, SHP-1.0/B16, SLP-1.3/A549, SLP-1.3/HepG<sub>2</sub>, and UPP-0.5/SH-SY5Y) for 48 h. The left side represents the control group, and the right side represents the experimental group.</p>
Full article ">Figure 10
<p>The effects of the <span class="html-italic">S. horneri</span> polysaccharide fraction (SHP-1.7) on the expression levels of Caspase-3, Caspase-8, Caspase-9, Bax, Bcl-2, and CDK-2 in HeLa cells. (<b>A</b>) Caspase-3, (<b>B</b>) Caspase-8, (<b>C</b>) Caspase-9, (<b>D</b>) Bax, (<b>E</b>) Bcl-2, and (<b>F</b>) CDK-2. All data are presented as mean ± SD. Significant differences (<span class="html-italic">p</span> &lt; 0.05) between groups are labeled with different letters (a–d).</p>
Full article ">

2022

Jump to: 2024, 2021, 2020, 2019, 2018, 2017, 2016, 2015, 2014, 2013, 2012, 2011, 2010

18 pages, 4078 KiB  
Article
Complexes of Cu–Polysaccharide of a Marine Red Microalga Produce Spikes with Antimicrobial Activity
by Nofar Yehuda, Levi A. Gheber, Ariel Kushmaro and Shoshana (Mails) Arad
Mar. Drugs 2022, 20(12), 787; https://doi.org/10.3390/md20120787 - 19 Dec 2022
Cited by 2 | Viewed by 3970
Abstract
Metal–polysaccharides have recently raised significant interest due to their multifunctional bioactivities. The antimicrobial activity of a complex of Cu2O with the sulfated polysaccharide (PS) of the marine red microalga Porphyridium sp. was previously attributed to spikes formed on the complex surface [...] Read more.
Metal–polysaccharides have recently raised significant interest due to their multifunctional bioactivities. The antimicrobial activity of a complex of Cu2O with the sulfated polysaccharide (PS) of the marine red microalga Porphyridium sp. was previously attributed to spikes formed on the complex surface (roughness). This hypothesis was further examined here using other Cu–PS complexes (i.e., monovalent-Cu2O, CuCl and divalent-CuO, CuCl2). The nanostructure parameters of the monovalent complexes, namely, longer spikes (1000 nm) and greater density (2000–5000 spikes/µm2) were found to be related to the superior inhibition of microbial growth and viability and biofilm formation. When Escherichia coli TV1061, used as a bioluminescent test organism, was exposed to the monovalent Cu–PS complexes, enhanced bioluminescence accumulation was observed, probably due to membrane perforation by the spikes on the surface of the complexes and consequent cytoplasmic leakage. In addition, differences were found in the surface chemistry of the monovalent and divalent Cu–PS complexes, with the monovalent Cu–PS complexes exhibiting greater stability (ζ-potential, FTIR spectra, and leaching out), which could be related to spike formation. This study thus supports our hypothesis that the spikes protruding from the monovalent Cu–PS surfaces, as characterized by their aspect ratio, are responsible for the antimicrobial and antibiofilm activities of the complexes. Full article
Show Figures

Figure 1

Figure 1
<p>FTIR transmission spectra of the divalent Cu–PS complexes (blue), the monovalent Cu–PS complexes (red) and the polysaccharide (black). All Cu–PS complexes contained 0.7% polysaccharide (<span class="html-italic">w/v</span>) and 500 ppm copper. To facilitate ease of viewing, the spectra are displaced with respect to the Y axis.</p>
Full article ">Figure 2
<p>Copper-release profiles from monovalent and divalent Cu–PS complexes. The copper concentration was determined in distilled water. Data represent the average values of three independent experiments. All the Cu–PS complexes contained 0.7% polysaccharide (<span class="html-italic">w/v</span>) and 500 ppm copper. Copper concentration was measured using a SPECTRO ARCOS ICP-OES analyzer.</p>
Full article ">Figure 3
<p>Effect of the monovalent and divalent Cu–PS complexes on (<b>A</b>) the inhibition of growth and (<b>B</b>) the cell viability of a fungus (<span class="html-italic">Candida albicans</span>), Gram-negative bacteria (<span class="html-italic">Acinetobacter baumannii</span>, <span class="html-italic">Pseudomonas aeruginosa</span>, and <span class="html-italic">Escherichia coli</span>), and Gram-positive bacteria (<span class="html-italic">Staphylococcus aureus</span> and <span class="html-italic">Bacillus subtilis</span>). All the Cu−PS complexes contained 0.07% (<span class="html-italic">w/v</span>) polysaccharide and 30 ppm copper. For the growth inhibition experiments, the control was the absorbance of the relevant growth medium with only the bacterium or the fungus (see the experimental section). The microbial cultures were incubated with shaking in 96-well plates at 37 °C for 14 h for each bacterial species or 48 h for <span class="html-italic">C. albicans</span>. Each sample was plated on an agar plate of the relevant medium after serial dilution and incubated overnight at 37 °C. CFUs were counted the following morning and were assessed vs. untreated cells. Values are the means ± standard error of mean (SEM)of three independent experiments performed in triplicate. All of the results were significantly different from their relative controls (ANOVA; <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Effect of monovalent and divalent Cu–PS complexes on the swarming motility of <span class="html-italic">P. aeruginosa</span> PA14. Top row: Motility of <span class="html-italic">P. aeruginosa</span> PA14 in the presence of Cu–PS complexes. Bottom row: Controls. The bacteria were inoculated into the center of each plate consisting of M9 solidified with 0.5% (<span class="html-italic">w/v</span>) Difco agar and containing 0.1% of the relevant Cu–PS complex. Surface coverage was assessed after 24 h of growth at 37 °C. All of the Cu–PS complexes contained 0.7% polysaccharide (<span class="html-italic">w/v</span>) and 500 ppm copper. For the control treatments, the copper concentration in the copper-containing plates was 500 ppm, and the PS plate contained 0.7% <span class="html-italic">Porphyridium</span> sp. polysaccharide (<span class="html-italic">w/v</span>).</p>
Full article ">Figure 5
<p>Effect of monovalent and divalent Cu–PS complexes on <span class="html-italic">P. aeruginosa</span> PA14 biofilm formation. Forty-eight hours after inoculation, biofilm formation was assessed by CLSM on (<b>A</b>) an untreated surface, (<b>B</b>) a surface pre-coated with <span class="html-italic">Porphyridium</span> sp. polysaccharide, (<b>C</b>,<b>D</b>) surfaces pre-coated with monovalent Cu–PS complexes, and (<b>E</b>,<b>F</b>) surfaces pre-coated with divalent Cu–PS complexes. Viable cells stained green, and dead cells stained red with the BacLight<sup>®</sup> DEAD/LIVE Kit for scanned areas of ~318 μm × 318 μm. (<b>G</b>) Cell-layer thickness (μm<sup>3</sup>/μm<sup>2</sup>) for live and dead cells; values are means ± standard error of mean (SEM). The results are presented for three independent sets of flow-cell experiments, each containing 30 measurements. Significant differences between the groups (<span class="html-italic">p</span> &lt; 0.0001) by two-way ANOVA, followed by Tukey’s test. (<b>H</b>) Ratio of dead-to-live cells (calculated from <a href="#marinedrugs-20-00787-f005" class="html-fig">Figure 5</a>G). All of the Cu–PS complexes contained 0.7% polysaccharide (<span class="html-italic">w/v</span>) and 500 ppm copper.</p>
Full article ">Figure 6
<p>Surface topography and spike distribution of monovalent and divalent Cu–PS complexes with and without the gold coating. (<b>A</b>) AFM 3D images and (<b>B</b>) spike parameters. 3D images of the Cu–PS complexes and the polysaccharide were manually analyzed using Gwyddion and ImageJ software. Spike thickness and density were calculated manually using ImageJ and the aspect ratio was calculated from the data collected from the AFM images analyzed by Gwyddion and ImageJ software. All of the Cu–PS complexes contained 0.7% polysaccharide (<span class="html-italic">w/v</span>) and 500 ppm copper.</p>
Full article ">Figure 7
<p>SEM micrographs showing the effect of monovalent and divalent Cu–PS complexes that were surface coated with gold vs. non-coated on <span class="html-italic">P. aeruginosa</span> PA14 biofilm formation. ×10,000, scale bar = 4 µm. All of the Cu–PS complexes contained 0.7% polysaccharide (<span class="html-italic">w/v</span>) and 500 ppm copper. Images of the control (glass surface alone) are presented in <a href="#app1-marinedrugs-20-00787" class="html-app">Figure S3, Supplementary Materials</a>.</p>
Full article ">Figure 8
<p>(<b>A</b>) Bioluminescence signal from <span class="html-italic">E. coli</span> TV1061 induced by monovalent and divalent Cu–PS complexes. The bioluminescence was measured in relative light units (RLU) at 490 nm (i.e., the wavelength attributed to bacterial luciferase). (<b>B</b>) Area under the peak showing the leakage of luciferase from the bacterial cells. All of the Cu−PS complexes contained 0.07% (w/v) polysaccharide and 30 ppm copper. In the copper salts (Cu<sub>2</sub>O, CuCl, CuO, CuCl<sub>2</sub>), the same pattern of bioluminescence as the PS alone was shown (not shown).</p>
Full article ">Figure 9
<p>Microbial cell death as a function of the characteristics of the spikes induced on the Cu–PS complexes. (<b>A</b>) Aspect ratios (•) of the spikes of the various Cu–PS complexes. (<b>B</b>) Luminescence of <span class="html-italic">E. coli</span> TV1061 as a function of the aspect ratio (<tt>■</tt>). (<b>C</b>) Microbial growth inhibition as function luminescence (∆) for five different microorganisms: <span class="html-italic">Candida albicans, Acinetobacter baumannii</span>, <span class="html-italic">Pseudomonas aeruginosa</span>, <span class="html-italic">Escherichia coli</span>, <span class="html-italic">Staphylococcus aureus</span>, and <span class="html-italic">Bacillus subtilis</span>. All the Cu−PS complexes contained 0.07% (<span class="html-italic">w/v</span>) polysaccharide and 30 ppm copper. The results presented are calculated from <a href="#marinedrugs-20-00787-f003" class="html-fig">Figure 3</a> and <a href="#marinedrugs-20-00787-f006" class="html-fig">Figure 6</a>.</p>
Full article ">
11 pages, 6884 KiB  
Article
Polysaccharide from Edible Alga Enteromorpha clathrata Improves Ulcerative Colitis in Association with Increased Abundance of Parabacteroides spp. in the Gut Microbiota of Dextran Sulfate Sodium-Fed Mice
by Mingfeng Ma, Tianyu Fu, Yamin Wang, Aijun Zhang, Puyue Gao, Qingsen Shang and Guangli Yu
Mar. Drugs 2022, 20(12), 764; https://doi.org/10.3390/md20120764 - 4 Dec 2022
Cited by 8 | Viewed by 2318
Abstract
Polysaccharide from the edible alga Enteromorpha clathrata has been demonstrated to exert beneficial effects on human health. However, what effect it has on inflammatory bowel diseases has not been investigated. Here, using a mouse model of dextran sulfate sodium (DSS)-induced ulcerative colitis, we [...] Read more.
Polysaccharide from the edible alga Enteromorpha clathrata has been demonstrated to exert beneficial effects on human health. However, what effect it has on inflammatory bowel diseases has not been investigated. Here, using a mouse model of dextran sulfate sodium (DSS)-induced ulcerative colitis, we illustrate that Enteromorpha clathrata polysaccharide (ECP) could alleviate body weight loss, reduce incidences of colonic bleeding, improve stool consistency and ameliorate mucosal damage in diseased mice. 16S rRNA high-throughput sequencing and bioinformatic analysis indicated that ECP significantly changed the structure of the gut microbiota and increased the abundance of Parabacteroides spp. in DSS-fed mice. In vitro fermentation studies further confirmed that ECP could promote the growth of Parabacteroides distasonis F1-28, a next-generation probiotic bacterium isolated from the human gut, and increase its production of short-chain fatty acids. Additionally, Parabacteroides distasonis F1-28 was also found to have anti-ulcerative colitis effects in DSS-fed mice. Altogether, our study demonstrates for the first time a beneficial effect of ECP on ulcerative colitis and provides a possible basis for understanding its therapeutic mechanisms from the perspective of symbiotic gut bacteria Parabacteroides distasonis. Full article
Show Figures

Figure 1

Figure 1
<p>Dietary ECP improved ulcerative colitis in DSS-fed mice. Changes of the body weight during the experiment (<b>A</b>). Representative morphologies of the colon (<b>B</b>). Colon length (<b>C</b>). Symptom score analysis of ulcerative colitis (<b>D</b>). * <span class="html-italic">p</span> &lt; 0.05 versus NC group; ** <span class="html-italic">p</span> &lt; 0.01 versus NC group; <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05 versus MD group; <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 versus MD group.</p>
Full article ">Figure 2
<p>Dietary ECP ameliorated DSS-induced mucosal damage in the colon. H&amp;E staining (<b>A</b>) and Alcian blue staining (<b>B</b>) of the colon tissues. Alcian blue staining was applied to show the changes of the intestinal acidic mucin <span class="html-italic">O</span>-glycans. Histopathological colon score analysis (<b>C</b>). ** <span class="html-italic">p</span> &lt; 0.01 versus NC group; <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05 versus MD group.</p>
Full article ">Figure 3
<p>Dietary ECP changed the overall structure of the gut microbiota in diseased mice. Venn diagram analysis of the OTUs (<b>A</b>). PCA score plot analysis (<b>B</b>). NMDS score plot analysis (<b>C</b>).</p>
Full article ">Figure 4
<p>Dietary ECP modulated the composition of the gut microbiota at different taxonomic levels. Heatmap analysis of the gut microbiota at the phylum and genus levels.</p>
Full article ">Figure 5
<p>LEfSe LDA score analysis of the gut microbiota in NC versus MD group (<b>A</b>) and MD versus ECP group (<b>B</b>). Only bacterial taxa with an LDA score of above 4.0 and 3.0 are listed.</p>
Full article ">Figure 6
<p>ECP promoted the growth of <span class="html-italic">P. distasonis</span> F1-28 and increased the production of SCFAs. Growth curves (<b>A</b>) and colony forming units (CFUs) (<b>B</b>). Concentrations of total SCFAs (<b>C</b>), acetate (<b>D</b>), propionate (<b>E</b>) and succinate (<b>F</b>). * <span class="html-italic">p</span> &lt; 0.05 versus NC group; ** <span class="html-italic">p</span> &lt; 0.01 versus NC group.</p>
Full article ">Figure 7
<p>Oral administration of live <span class="html-italic">P. distasonis</span> F1-28 alleviated ulcerative colitis in DSS-fed mice. Changes of the body weight during the experiment (<b>A</b>). Representative morphologies of the colon (<b>B</b>). Colon length (<b>C</b>). Symptom score analysis of ulcerative colitis (<b>D</b>). ** <span class="html-italic">p</span> &lt; 0.01 versus NC group; <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05 versus MD group; <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 versus MD group.</p>
Full article ">Figure 8
<p>Oral administration of live <span class="html-italic">P. distasonis</span> F1-28 attenuated DSS-induced mucosal damage in the colon. H&amp;E staining (<b>A</b>) and Alcian blue staining (<b>B</b>) of the colon tissues. Histopathological colon score analysis (<b>C</b>). ** <span class="html-italic">p</span> &lt; 0.01 versus NC group; <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 versus MD group.</p>
Full article ">
22 pages, 6555 KiB  
Article
Anti-Biofilm Activity of a Hyaluronan-like Exopolysaccharide from the Marine Vibrio MO245 against Pathogenic Bacteria
by Marie Champion, Emilie Portier, Karine Vallée-Réhel, Isabelle Linossier, Eric Balnois, Guillaume Vignaud, Xavier Moppert, Claire Hellio and Fabienne Faÿ
Mar. Drugs 2022, 20(11), 728; https://doi.org/10.3390/md20110728 - 21 Nov 2022
Cited by 7 | Viewed by 2430
Abstract
Biofilms, responsible for many serious drawbacks in the medical and marine environment, can grow on abiotic and biotic surfaces. Commercial anti-biofilm solutions, based on the use of biocides, are available but their use increases the risk of antibiotic resistance and environmental pollution in [...] Read more.
Biofilms, responsible for many serious drawbacks in the medical and marine environment, can grow on abiotic and biotic surfaces. Commercial anti-biofilm solutions, based on the use of biocides, are available but their use increases the risk of antibiotic resistance and environmental pollution in marine industries. There is an urgent need to work on the development of ecofriendly solutions, formulated without biocidal agents, that rely on the anti-adhesive physico-chemical properties of their materials. In this context, exopolysaccharides (EPSs) are natural biopolymers with complex properties than may be used as anti-adhesive agents. This study is focused on the effect of the EPS MO245, a hyaluronic acid-like polysaccharide, on the growth, adhesion, biofilm maturation, and dispersion of two pathogenic model strains, Pseudomonas aeruginosa sp. PaO1 and Vibrio harveyi DSM19623. Our results demonstrated that MO245 may limit biofilm formation, with a biofilm inhibition between 20 and 50%, without any biocidal activity. Since EPSs have no significant impact on the bacterial motility and quorum sensing factors, our results indicate that physico-chemical interactions between the bacteria and the surfaces are modified due to the presence of an adsorbed EPS layer acting as a non-adsorbing layer. Full article
Show Figures

Figure 1

Figure 1
<p>Structure of (<b>A</b>) MO245 and (<b>B</b>) HA.</p>
Full article ">Figure 2
<p>Impact of MO245 or HA (125 µg/mL) on the growth of (<b>A</b>) <span class="html-italic">P. aeruginosa</span> in LB medium at 37 °C and (<b>B</b>) <span class="html-italic">V. harveyi</span> in Zobell medium at 28 °C for 30 h under 125 rpm agitation. Average of three independent replicates ± standard deviation. Bactericidal effect and consumption of MO245 or HA (125 µg/mL) or glucose (4 g/L) as a carbon source in M9 medium for (<b>C</b>) <span class="html-italic">P. aeruginosa</span> at 37 °C and (<b>D</b>) <span class="html-italic">V. harveyi</span> at 28 °C for 26 h under 125 rpm agitation.</p>
Full article ">Figure 3
<p>Conditions of use of MO and HA during adhesion. (<b>A</b>) Addition of MO245 or HA within the bacterial suspension and (<b>B</b>) glass slide conditioned with MO245 or HA.</p>
Full article ">Figure 4
<p>Impact of MO245 and HA on a 24 h <span class="html-italic">P. aeruginosa</span> biofilm maturation. (<b>A</b>) Biomass and average thickness quantification after COMSTAT analysis of confocal laser microscopy observations. (<b>B</b>) Confocal laser microscopy observation (Syto9<sup>®</sup>) without or with the addition of MO245 or HA at 125 µg/mL in the LB growth medium. Data represent the mean ± the standard deviation. * represents the significant difference at α 5%: <span class="html-italic">p</span> &lt; 0.05, *** represents the significant difference at α 5%: <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 5
<p>Impact of MO245 and HA on a 24 h <span class="html-italic">V. harveyi</span> biofilm maturation. (<b>A</b>) Biomass and average thickness quantification after COMSTAT analysis of confocal laser microscopy observation. (<b>B</b>) Confocal laser microscopy observation (Syto9<sup>®</sup>) without or with the addition of MO245 or HA at 125 µg/mL in the Zobell growth medium. Data represent the mean ± the standard deviation. ** represents the significant difference at α 5%: <span class="html-italic">p</span> &lt; 0.01; *** represents the significant difference at α 5%: <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 6
<p>Impact of MO245 and HA on the degradation of a 24 h <span class="html-italic">P. aeruginosa</span> biofilm already formed. (<b>A</b>) Biomass and average thickness quantification after COMSTAT analysis of confocal laser microscopy observation. (<b>B</b>) Confocal laser microscopy observation (Syto9<sup>®</sup>) without or with the addition of MO245 or HA at 125 µg/mL on a 24 h biofilm already formed. Data represent the mean ± the standard deviation. * represents the significant difference at α 5%: <span class="html-italic">p</span> &lt; 0.05, ** represents the significant difference at α 5%: <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 7
<p>Impact of MO245 and HA on the degradation of a 24 h <span class="html-italic">V. harveyi</span> biofilm already formed. (<b>A</b>) Biomass and average thickness quantification after COMSTAT analysis of confocal laser microscopy observation. (<b>B</b>) Confocal laser microscopy observation (Syto9<sup>®</sup>) without or with the addition of MO245 or HA at 125 µg/mL on a 24 h biofilm already formed. Data represent the mean ± the standard deviation. ** represents the significant difference at α 5%: <span class="html-italic">p</span> &lt; 0.01, *** represents the significant difference at α 5%: <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 8
<p>Motility assay of (<b>A</b>) <span class="html-italic">P. aeruginosa</span> and (<b>B</b>) <span class="html-italic">V. harveyi</span> incubated with MO245 or HA at 125 µg/mL or nothing for 2 h at room temperature. Diameters were measured after the overnight incubation of agar plates at 37 °C for <span class="html-italic">P. aeruginosa</span> and 28 °C for <span class="html-italic">V. harveyi</span>. Data represent the mean ± the standard deviation. * represents the significant difference at α 5%: <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 9
<p>Anti-quorum sensing effect of MO245, HA, and kojic acid (KA) at 125 µg/mL on (<b>A</b>) the biosensor <span class="html-italic">E. coli</span> pSB401 and (<b>B</b>) <span class="html-italic">V. harveyi</span> for 9 h at 28 °C under 125 rpm agitation. Luminescence and O.D.<sub>600</sub> were measured every hour and RLU ratios were calculated.</p>
Full article ">Figure 10
<p>Percentage hydrophobicity of <span class="html-italic">P. aeruginosa</span> and <span class="html-italic">V. harveyi</span> in the presence of MO245 or HA. A total of 10<sup>9</sup> bacteria were put in contact with MO245 or HA at 125 µg/mL for 2 h and then in contact with toluene. At the appearance of a phase separation, the aqueous phase was recovered and the optical density was measured. The percentage of hydrophobicity was then calculated. * represents the significant difference at α 5%: <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 11
<p>Evolution of the frequency F3 versus time for MO245 and HA on silica-coated quartz crystal (pH 7, 150 mM NaCl). MO245 and HA were injected at t = 0 min. After the adsorption of MO245 on the surface of silica-coated quartz crystal, rinsing was performed to observe if desorption occurred.</p>
Full article ">Figure 12
<p>The water contact angle of a surface conditioned with MO245 or HA at 125 µg/mL measured with a Digidrop. A total of 3 mL of MO245 and HA were deposited on a glass slide and allowed to evaporate in a sterile environment. Contact angles were measured at room temperature with a volume of 3 µL. *** represents the significant difference at α 5%: <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 13
<p>Emulsifying properties of MO245, HA, and Triton X-100 at 0.25% (<span class="html-italic">w</span>/<span class="html-italic">v</span>) in the oil-in-water phase over time. The emulsion index was calculated after 1 h, 24 h, 48 h, and 168 h at 20 °C.</p>
Full article ">Figure 14
<p>Summary scheme of the potential anti-biofilm mode of action of MO245.</p>
Full article ">
18 pages, 2621 KiB  
Article
Neoagaro-Oligosaccharides Ameliorate Chronic Restraint Stress-Induced Depression by Increasing 5-HT and BDNF in the Brain and Remodeling the Gut Microbiota of Mice
by Yan Zhuang, Runying Zeng, Xiao Liu, Longhe Yang and Zhuhua Chan
Mar. Drugs 2022, 20(11), 725; https://doi.org/10.3390/md20110725 - 18 Nov 2022
Cited by 10 | Viewed by 3128
Abstract
Neoagaro-oligosaccharides (NAOs) belong to the algae oligosaccharides. NAOs have been found to have diverse biological activities. However, the effects of NAOs on depression and their underlying mechanism have not been thoroughly studied. A chronic restraint stress (CRS)-induced C57BL/6J mouse model was used to [...] Read more.
Neoagaro-oligosaccharides (NAOs) belong to the algae oligosaccharides. NAOs have been found to have diverse biological activities. However, the effects of NAOs on depression and their underlying mechanism have not been thoroughly studied. A chronic restraint stress (CRS)-induced C57BL/6J mouse model was used to assess the antidepressant effects of NAOs. Anxiety and depression behaviors were assessed by open field tests (OFT) and forced swimming tests (FST), while interleukin 18 (IL-18), 5-hydroxytryptamine (5-HT) and brain-derived neurotrophic factor (BDNF) were the molecular biomarkers of depression. Fecal microbiota transplantation (FMT) was performed. The results showed that NAO treatment significantly improved the body weight of depressed mice and reduced the central area time in the OFT and immobility time in the FST. NAO treatment decreased the levels of IL-18 in the serum and increased the levels of 5-HT in the serum and whole brain and of BDNF in the whole brain. NAO treatment mitigated the gut microbiota dysbiosis in the depressed mice and reversed the decreased levels of short-chain fatty acids (SCFAs) in the cecum of the depressed mice. FMT indicated that the gut microbiota is, indeed, linked to depression, which was reflected in the changes in weight gain and behaviors. In a word, NAOs effectively reversed the CRS-induced mice model of depression, which depended on the changes in the gut microbiota and SCFAs, as well as its modulation of 5-HT and BDNF. Full article
Show Figures

Figure 1

Figure 1
<p>Effects of NAOs on body weight, OFT, and FST in CRS-induced mice. (<b>A</b>) The CRS mice were repeatedly placed in a plastic restrainer customized with a 50 mL centrifuge tube for 4 h (from 10:00 to 14:00) every day for 28 consecutive days. During the restraint stress period, the CRS mice were deprived of food and water. Mice in the control group had free access to water and food without any restraint. After establishing the depression model on day 28, The mice in the low-concentration NAO group (LOW) and HIGH groups were supplied daily with NAOs at doses of 100 and 200 mg/kg based on their body weight, respectively. The mice in the PA group were supplied with paroxetine at a dose of 10 mg/kg. The mice in the CK and CRS groups received sterilized double-distilled water through oral gavage in the same volume as other groups. <span class="html-italic">n</span> = 10–12 per test. (<b>B</b>) Central area time over 4 min in the OFT. <span class="html-italic">n</span> = 4–5 per test. (<b>C</b>) Immobility time in FST. <span class="html-italic">n</span> = 4–5 per test. (<b>D</b>) Representative locomotion traces in OFT. Two-way ANOVA tests were used to analyze the weight gain rate, and unpaired <span class="html-italic">t</span> tests were used to analyze the other data. Data are presented as mean ± SEM. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, **** <span class="html-italic">p</span> &lt; 0.0001 vs. control, # <span class="html-italic">p</span> &lt; 0.05, ## <span class="html-italic">p</span> &lt; 0.01 vs. model group.</p>
Full article ">Figure 2
<p>Weight gain rate 7 days after terminating the restraints and administrations. Data are presented as mean ± SEM. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 vs. control group.</p>
Full article ">Figure 3
<p>Gut microbiota diversity. (<b>A</b>) Relative abundance of the cecal microbiota on the phylum level and mean abundance of marker species in each group. The <span class="html-italic">p</span> values are based on Welch’s <span class="html-italic">t</span> test. (<b>B</b>) Relative abundance of the cecal microbiota on the genus level and mean abundance of marker species in each group. The <span class="html-italic">p</span> values are based on Welch’s <span class="html-italic">t</span> test. (<b>C</b>) The α-diversity analysis and Shannon diversity index. The <span class="html-italic">p</span> values are based on Kruskal–Wallis. (<b>D</b>) Effects of NAOs on β-diversity of intestinal microbes. PCoA and NMDS score plots based on Bray (OTU) (<span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 3 Cont.
<p>Gut microbiota diversity. (<b>A</b>) Relative abundance of the cecal microbiota on the phylum level and mean abundance of marker species in each group. The <span class="html-italic">p</span> values are based on Welch’s <span class="html-italic">t</span> test. (<b>B</b>) Relative abundance of the cecal microbiota on the genus level and mean abundance of marker species in each group. The <span class="html-italic">p</span> values are based on Welch’s <span class="html-italic">t</span> test. (<b>C</b>) The α-diversity analysis and Shannon diversity index. The <span class="html-italic">p</span> values are based on Kruskal–Wallis. (<b>D</b>) Effects of NAOs on β-diversity of intestinal microbes. PCoA and NMDS score plots based on Bray (OTU) (<span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 4
<p>Effects of NAOs on the fecal SCFAs in CRS mice. (<b>A</b>) Total SCFAs, (<b>B</b>) hexanoic acid, (<b>C</b>) valeric acid, (<b>D</b>) butyric acid, (<b>E</b>) isobutyric acid, and (<b>F</b>) acetic acid, <span class="html-italic">n</span> = 4–5 per test. Unpaired <span class="html-italic">t</span> tests were used to analyze the data. Data are presented as mean ± SEM. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 vs. control group; # <span class="html-italic">p</span> &lt; 0.05, ## <span class="html-italic">p</span> &lt; 0.01 vs. model group.</p>
Full article ">Figure 5
<p>Biochemical assays in the serum and whole brain. (<b>A</b>) IL-18 and 5-HT levels were detected by ELISA in the serum. <span class="html-italic">n</span> = 4–5 per test. (<b>B</b>) 5-HT and BDNF in the brain were detected by ELISA. <span class="html-italic">n</span> = 4–5 per test. Unpaired <span class="html-italic">t</span> tests were used to analyze the data. Data are presented as mean ± SEM. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.0 vs. control, # <span class="html-italic">p</span> &lt; 0.05, ## <span class="html-italic">p</span> &lt; 0.01 vs. model group.</p>
Full article ">Figure 5 Cont.
<p>Biochemical assays in the serum and whole brain. (<b>A</b>) IL-18 and 5-HT levels were detected by ELISA in the serum. <span class="html-italic">n</span> = 4–5 per test. (<b>B</b>) 5-HT and BDNF in the brain were detected by ELISA. <span class="html-italic">n</span> = 4–5 per test. Unpaired <span class="html-italic">t</span> tests were used to analyze the data. Data are presented as mean ± SEM. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.0 vs. control, # <span class="html-italic">p</span> &lt; 0.05, ## <span class="html-italic">p</span> &lt; 0.01 vs. model group.</p>
Full article ">Figure 6
<p>Effects of NAOs on body weight, OFT, and FST in FMT mice. (<b>A</b>) Weight measurements of FMT mice. <span class="html-italic">n</span> = 4–5 per test. (<b>B</b>) OFT of FMT mice. <span class="html-italic">n</span> = 4 per test. (<b>C</b>) FST of FMT mice. <span class="html-italic">n</span> = 4 per test. (<b>D</b>) Representative locomotion traces in OFT. # <span class="html-italic">p</span> &lt; 0.05, ## <span class="html-italic">p</span> &lt; 0.01 vs. model group.</p>
Full article ">Figure 7
<p>Experimental procedure. (<b>A</b>) Schematic of the total experimental procedures. (<b>B</b>) Schematic of the FMT experiment.</p>
Full article ">
14 pages, 2582 KiB  
Article
In Vivo Anticoagulant and Antithrombic Activity of Depolymerized Glycosaminoglycan from Apostichopus japonicus and Dynamic Effect–Exposure Relationship in Rat Plasma
by Han Wang, Dandan He, Linlin Duan, Lv Lv, Qun Gao, Yuanhong Wang, Shuang Yang and Zhihua Lv
Mar. Drugs 2022, 20(10), 631; https://doi.org/10.3390/md20100631 - 2 Oct 2022
Cited by 2 | Viewed by 2638
Abstract
Glycosaminoglycan from Apostichopus japonicus (AHG) and its depolymerized fragments (DAHGs) are anticoagulant fucosylated chondroitin sulfate. The aim of this study was to further evaluate the anticoagulant and antithrombic activity of AHG and DAHGs, as well as reveal the dynamic relationship between exposure and [...] Read more.
Glycosaminoglycan from Apostichopus japonicus (AHG) and its depolymerized fragments (DAHGs) are anticoagulant fucosylated chondroitin sulfate. The aim of this study was to further evaluate the anticoagulant and antithrombic activity of AHG and DAHGs, as well as reveal the dynamic relationship between exposure and effect in vivo. The results demonstrated that AHG100 (Mw~100 kDa), DAHG50 (Mw~50 kDa), and DAHG10 (Mw~10 kDa) exhibited potent anticoagulant activity by inhibiting intrinsic factor Xase complex (FXase) as well as antithrombin-dependent factor IIa (FIIa) and factor Xa (FXa). These glycosaminoglycans markedly prevented thrombosis formation and thrombin-induced platelet aggregation in a dose- and molecular weight-dependent manner in vitro and in vivo. The further bleeding time measurement indicated that DAHG10 exhibited obviously lower hemorrhage risks than native AHG100. Following oral administration, DAHG10 could be absorbed into blood, further dose-dependently prolonging activated partial thromboplastin time (APTT) and thrombin time (TT) as well as inhibiting FXa and FIIa partially through FXase. Anticoagulant activity was positively associated with plasma concentration following oral administration of DAHG10. Our study proposed a new point of view to understand the correlation between effects and exposure of fucosylated chondroitin sulfate as an effective and safe oral antithrombotic agent. Full article
Show Figures

Figure 1

Figure 1
<p>The structure of AHG100.</p>
Full article ">Figure 2
<p><sup>1</sup>H NMR and IR spectrum of AHG100 and DAHGs. (<b>A</b>) <sup>1</sup>H NMR spectrum, (<b>B</b>) IR spectrum.</p>
Full article ">Figure 3
<p>The structural information and anticoagulant activity of AHG100 and DAHGs (mean ± SD, <span class="html-italic">n</span> = 3). (<b>A</b>) Anti-FIIa activity, (<b>B</b>) anti-FXa activity, (<b>C</b>) anti-intrinsic FXase complex activity, and (<b>D</b>) platelet aggregation. Compared with normal control (NC) group, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 4
<p>Effects of AHG100 and DAHGs on CT and BT (mean ± SD, <span class="html-italic">n</span> = 10). (<b>A</b>) CT, (<b>B</b>) BT. Compared with NC group * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 5
<p>Effects of AHG100 and DAHGs on antithrombosis (mean ± SD, <span class="html-italic">n</span> = 8). (<b>A</b>) Thrombus formation time, (<b>B</b>) thrombus length, and (<b>C</b>) thrombus weight. Compared with NC group ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 6
<p>Effect of AHG100 and DAHGs on APTT and TT in rats (mean ± SD, <span class="html-italic">n</span> = 3). (<b>A</b>) AHG100, APTT; (<b>B</b>) AHG100, TT; (<b>C</b>) DAHG50, APTT; (<b>D</b>) DAHG50, TT; (<b>E</b>) DAHG10, APTT; (<b>F</b>) DAHG10, TT. Compared with blank time of intravenous administration, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01. Compared with blank time of oral administration, <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 7
<p>Effect of DAHG10 on FXa, FIIa, and FXIIa in rats (mean ± SD, <span class="html-italic">n</span> = 4). (<b>A</b>) Residual activity of FIIa after oral dosing with DAHG10. (<b>B</b>) Residual activity of FIIa after injection of DAHG10. (<b>C</b>) Residual activity of FXa after oral dosing with DAHG10. (<b>D</b>) Residual activity of FXa after injection of DAHG10. (<b>E</b>) Residual activity of factor XIIa (FXIIa) after oral dosing with DAHG10. (<b>F</b>) Residual activity of FXIIa after injection of DAHG10. Compared with blank time after administration, <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05, <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01, <sup>###</sup> <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 8
<p>The concentration–time–effect relationship in rats (mean ± SD, <span class="html-italic">n</span> = 4). (<b>A</b>) Inhibition of FIIa activity after oral administration of DAHG10 at 250 mg/kg. (<b>B</b>) Inhibition of FIIa activity after intravenous injection of DAHG10 at 5 mg/kg [<a href="#B22-marinedrugs-20-00631" class="html-bibr">22</a>]. (<b>C</b>) Inhibition of FXa activity after oral administration of DAHG10 at 250 mg/kg. (<b>D</b>) Inhibition of FXa activity after intravenous injection of DAHG10 at 5 mg/kg [<a href="#B22-marinedrugs-20-00631" class="html-bibr">22</a>].</p>
Full article ">
16 pages, 4370 KiB  
Article
Synthesis, Characterization, and the Antioxidant Activity of Phenolic Acid Chitooligosaccharide Derivatives
by Yan Sun, Xia Ji, Jingmin Cui, Yingqi Mi, Jingjing Zhang and Zhanyong Guo
Mar. Drugs 2022, 20(8), 489; https://doi.org/10.3390/md20080489 - 28 Jul 2022
Cited by 18 | Viewed by 2221
Abstract
A series of phenolic acid chitooligosaccharide (COS) derivatives synthesized by two mild and green methods were illuminated in this paper. Seven phenolic acids were selected to combine two kinds of COS derivatives: the phenolic acid chitooligosaccharide salt derivatives and the phenolic-acid-acylated chitooligosaccharide derivatives. [...] Read more.
A series of phenolic acid chitooligosaccharide (COS) derivatives synthesized by two mild and green methods were illuminated in this paper. Seven phenolic acids were selected to combine two kinds of COS derivatives: the phenolic acid chitooligosaccharide salt derivatives and the phenolic-acid-acylated chitooligosaccharide derivatives. The structures of the derivatives were characterized by FT-IR and 1H NMR spectra. The antioxidant experiment results in vitro (including DPPH-radical scavenging activity, superoxide-radical scavenging activity, hydroxyl-radical scavenging ability, and reducing power) demonstrated that the derivatives exhibited significantly enhanced antioxidant activity compared to COS. Moreover, the study showed that the phenolic acid chitooligosaccharide salts had stronger antioxidant activity than phenolic-acid-acylated chitooligosaccharide. The cytotoxicity assay of L929 cells in vitro indicated that the derivatives had low cytotoxicity and good biocompatibility. In conclusion, this study provides a possible synthetic method for developing novel and nontoxic antioxidant agents which can be used in the food and cosmetics industry. Full article
Show Figures

Figure 1

Figure 1
<p>FT-IR spectra of COS and phenolic acid COS salt derivatives.</p>
Full article ">Figure 2
<p>FT-IR spectra of COS and phenolic-acid-acylated chitooligosaccharide derivatives.</p>
Full article ">Figure 3
<p><sup>1</sup>H NMR spectra of COS and phenolic acid chitooligosaccharide salt derivatives.</p>
Full article ">Figure 4
<p><sup>1</sup>H NMR spectra of COS and phenolic-acid-acylated chitooligosaccharide derivatives.</p>
Full article ">Figure 5
<p>DPPH-radical scavenging activity of COS, ascorbic acid (VC), phenolic acid chitooligosaccharide salt derivatives (<b>a</b>), and phenolic-acid-acylated chitooligosaccharide derivatives (<b>b</b>).</p>
Full article ">Figure 6
<p>Superoxide-radical scavenging activity of COS, ascorbic acid (VC), phenolic acid chitooligosaccharide salt derivatives (<b>a</b>), and phenolic-acid-acylated chitooligosaccharide derivatives (<b>b</b>).</p>
Full article ">Figure 7
<p>Hydroxyl-radical scavenging activity of COS, ascorbic acid (VC), phenolic acid chitooligosaccharide salt derivatives (<b>a</b>), and phenolic-acid-acylated chitooligosaccharide derivatives (<b>b</b>).</p>
Full article ">Figure 8
<p>Reducing power of COS ascorbic acid (VC), phenolic acid chitooligosaccharide salt derivatives (<b>a</b>), and phenolic-acid-acylated chitooligosaccharide derivatives (<b>b</b>).</p>
Full article ">Figure 9
<p>The cytotoxicity of COS and the phenolic acid chitooligosaccharide salt derivatives on L929 cells.</p>
Full article ">Figure 10
<p>The cytotoxicity of COS and the phenolic-acid-acylated chitooligosaccharide derivatives on L929 cells.</p>
Full article ">Figure 11
<p>The L929 cells after the addition of COS and the phenolic acid chitooligosaccharide salt derivatives for 24 h (the sample concentration for the cell growth pictures was 1000 μg/mL).</p>
Full article ">Figure 12
<p>The L929 cells after the addition of COS and phenolic-acid-acylated chitooligosaccharide derivatives for 24 h (the sample concentration for the cell growth pictures was 1000 μg/mL).</p>
Full article ">Scheme 1
<p>Synthesis routes of phenolic acid chitooligosaccharide derivatives.</p>
Full article ">
20 pages, 3981 KiB  
Article
Immunomodulatory Activity In Vitro and In Vivo of a Sulfated Polysaccharide with Novel Structure from the Green Alga Ulvaconglobata Kjellman
by Sujian Cao, Yajing Yang, Shan Liu, Zhuling Shao, Xiao Chu and Wenjun Mao
Mar. Drugs 2022, 20(7), 447; https://doi.org/10.3390/md20070447 - 8 Jul 2022
Cited by 10 | Viewed by 2723
Abstract
Algae accumulate large amounts of polysaccharides in their cell walls or intercellular regions. Polysaccharides from algae possess high potential as promising candidates for marine drug development. In this study, a sulfated polysaccharide, UCP, from the green alga Ulva conglobata Kjellman was obtained by [...] Read more.
Algae accumulate large amounts of polysaccharides in their cell walls or intercellular regions. Polysaccharides from algae possess high potential as promising candidates for marine drug development. In this study, a sulfated polysaccharide, UCP, from the green alga Ulva conglobata Kjellman was obtained by water extraction, anion-exchange, and size-exclusion chromatography purification, and its structure was characterized by a combination of chemical and spectroscopic methods. UCP mainly consisted of →4)-α/β-l-Rhap-(1→, →4)-β-d-Xylp-(1→ and →4)-β-d-GlcAp-(1→ residues. Sulfate ester groups were substituted mainly at C-3 of →4)-l-Rhap-(1→ and C-2 of →4)-β-d-Xylp-(1→. Partial glycosylation was at C-2 of →4)-α-l-Rhap-(1→ residues. UCP possessed a potent immunomodulatory effect in vitro, evaluated by the assays of lymphocyte proliferation and macrophage phagocytosis. The immunomodulatory activity of UCP in vivo was further investigated using immunosuppressive mice induced by cyclophosphamide. The results showed that UCP markedly increased the spleen and thymus indexes and ameliorated the cyclophosphamide-induced damage to the spleen and thymus. UCP could increase the levels of white blood cells, lymphocytes, and platelets, and improve the hematopoietic inhibition caused by cyclophosphamide. Moreover, UCP significantly promoted the secretions of the immunoglobulin (Ig)G, IgE, and IgM. The data demonstrated that UCP is a novel sulfated polysaccharide and may be a promising immunomodulatory agent. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>HPGPC chromatogram, HPLC chromatogram, and IR spectrum of UCP. (<b>A</b>) HPGPC chromatogram of UCP on a Shodex OHpak SB-804 HQ column, and the standard curve of molecular weight; (<b>B</b>) HPLC chromatogram for monosaccharide composition analysis of UCP (Man: <span class="html-small-caps">d</span>-mannose, GlcN: <span class="html-small-caps">d</span>-glucosamine, Rha: <span class="html-small-caps">l</span>-rhamnose, GlcA: <span class="html-small-caps">d</span>-glucuronic acid, GalA: <span class="html-small-caps">d</span>-galacturonic acid, Glc: <span class="html-small-caps">d</span>-glucose, Gal: <span class="html-small-caps">d</span>-galactose, Xyl: <span class="html-small-caps">d</span>-xylose, Ara: <span class="html-small-caps">l</span>-arabinose, Fuc: <span class="html-small-caps">l</span>-fucose); (<b>C</b>) HPLC chromatogram of the sugar configuration determination of UCP (<span class="html-small-caps">d</span>-Rha: D-rhamnose, <span class="html-small-caps">d</span>-GlcA: <span class="html-small-caps">d</span>-glucuronic acid, <span class="html-small-caps">d</span>-Xyl: <span class="html-small-caps">d</span>-xylose, <span class="html-small-caps">l</span>-Rha: <span class="html-small-caps">l</span>-rhamnose, <span class="html-small-caps">l</span>-GlcA: <span class="html-small-caps">l</span>-glucuronic acid, <span class="html-small-caps">l</span>-Xyl: <span class="html-small-caps">l</span>-xylose); (<b>D</b>) IR spectrum of UCP.</p>
Full article ">Figure 1 Cont.
<p>HPGPC chromatogram, HPLC chromatogram, and IR spectrum of UCP. (<b>A</b>) HPGPC chromatogram of UCP on a Shodex OHpak SB-804 HQ column, and the standard curve of molecular weight; (<b>B</b>) HPLC chromatogram for monosaccharide composition analysis of UCP (Man: <span class="html-small-caps">d</span>-mannose, GlcN: <span class="html-small-caps">d</span>-glucosamine, Rha: <span class="html-small-caps">l</span>-rhamnose, GlcA: <span class="html-small-caps">d</span>-glucuronic acid, GalA: <span class="html-small-caps">d</span>-galacturonic acid, Glc: <span class="html-small-caps">d</span>-glucose, Gal: <span class="html-small-caps">d</span>-galactose, Xyl: <span class="html-small-caps">d</span>-xylose, Ara: <span class="html-small-caps">l</span>-arabinose, Fuc: <span class="html-small-caps">l</span>-fucose); (<b>C</b>) HPLC chromatogram of the sugar configuration determination of UCP (<span class="html-small-caps">d</span>-Rha: D-rhamnose, <span class="html-small-caps">d</span>-GlcA: <span class="html-small-caps">d</span>-glucuronic acid, <span class="html-small-caps">d</span>-Xyl: <span class="html-small-caps">d</span>-xylose, <span class="html-small-caps">l</span>-Rha: <span class="html-small-caps">l</span>-rhamnose, <span class="html-small-caps">l</span>-GlcA: <span class="html-small-caps">l</span>-glucuronic acid, <span class="html-small-caps">l</span>-Xyl: <span class="html-small-caps">l</span>-xylose); (<b>D</b>) IR spectrum of UCP.</p>
Full article ">Figure 2
<p>Structures of the possible main repeating disaccharides in UCP. (<b>a</b>): →2,4)-α-<span class="html-small-caps">l</span>-Rha<span class="html-italic">p</span>-(1→4)-β-<span class="html-small-caps">d</span>-Xyl<span class="html-italic">p</span>-(1→, (<b>b</b>): →4)-α-<span class="html-small-caps">l</span>-Rha<span class="html-italic">p</span>(3SO<sub>4</sub>)-(1→4)-β-<span class="html-small-caps">d</span>-Xyl<span class="html-italic">p</span>-(2SO<sub>4</sub>)-(1→, (<b>c</b>): →4)-β-<span class="html-small-caps">d</span>-Xyl<span class="html-italic">p</span>-(2SO<sub>4</sub>)-(1→4)-α-<span class="html-small-caps">l</span>-Rha<span class="html-italic">p</span>(3SO<sub>4</sub>)-(1→, (<b>d</b>): →4)-β-<span class="html-small-caps">l</span>-Rha<span class="html-italic">p</span>(3SO<sub>4</sub>)-(1→2,4)-α-<span class="html-small-caps">l</span>-Rha<span class="html-italic">p</span>-(1→, (<b>e</b>): →4)-β-<span class="html-small-caps">l</span>-Rha<span class="html-italic">p</span>(3SO<sub>4</sub>)-(1→4)-β-<span class="html-small-caps">d</span>-GlcA<span class="html-italic">p</span>-(1→, (<b>f</b>): →4)-β-<span class="html-small-caps">l</span>-Rha<span class="html-italic">p</span>-(1→4)-β-<span class="html-small-caps">d</span>-Xyl<span class="html-italic">p</span>-(2SO<sub>4</sub>)-(1→, (<b>g</b>): →4)-β-<span class="html-small-caps">d</span>-GlcA<span class="html-italic">p</span>-(1→4)-α-<span class="html-small-caps">l</span>-Rha<span class="html-italic">p</span>(3SO<sub>4</sub>)-(1→, (<b>h</b>): →4)-β-<span class="html-small-caps">d</span>-Xyl<span class="html-italic">p</span>-(1→4)-α-<span class="html-small-caps">l</span>-Rha<span class="html-italic">p</span>(3SO<sub>4</sub>)-(1→. These are the possible main repeating disaccharides, and some of them may not be present.</p>
Full article ">Figure 3
<p>Effects of UCP on lymphocyte proliferation and macrophage phagocytosis in vitro. (<b>A</b>) Spleen cell proliferation was treated with 5 μg/mL Con A; (<b>B</b>) spleen cell proliferation was treated with 20 μg/mL LPS; and (<b>C</b>) macrophage phagocytosis. Values are mean ± standard deviation (SD) (<span class="html-italic">n</span> = 3). Significant: * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05; <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 vs. control group. OD: optical density.</p>
Full article ">Figure 3 Cont.
<p>Effects of UCP on lymphocyte proliferation and macrophage phagocytosis in vitro. (<b>A</b>) Spleen cell proliferation was treated with 5 μg/mL Con A; (<b>B</b>) spleen cell proliferation was treated with 20 μg/mL LPS; and (<b>C</b>) macrophage phagocytosis. Values are mean ± standard deviation (SD) (<span class="html-italic">n</span> = 3). Significant: * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05; <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 vs. control group. OD: optical density.</p>
Full article ">Figure 4
<p>Influences of UCP on the spleen index (<b>A</b>) and thymus index (<b>B</b>) in the immunosuppressive mice induced by cyclophosphamide. The data were represented as the means ± SD (<span class="html-italic">n</span> = 10). Significant: ** <span class="html-italic">p</span> &lt; 0.01 vs. normal group; <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05, <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 vs. model group.</p>
Full article ">Figure 4 Cont.
<p>Influences of UCP on the spleen index (<b>A</b>) and thymus index (<b>B</b>) in the immunosuppressive mice induced by cyclophosphamide. The data were represented as the means ± SD (<span class="html-italic">n</span> = 10). Significant: ** <span class="html-italic">p</span> &lt; 0.01 vs. normal group; <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05, <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 vs. model group.</p>
Full article ">Figure 5
<p>Effects of UCP on the white blood cell content (<b>A</b>), lymphocyte content (<b>B</b>), and platelet content (<b>C</b>) in immunosuppressive mice induced by cyclophosphamide. The data were represented as the means ± SD (<span class="html-italic">n</span> = 10). Significant: ** <span class="html-italic">p</span> &lt; 0.01 vs. normal group; <sup>##</sup> <span class="html-italic">p</span> &lt; 0.05 vs. model group.</p>
Full article ">Figure 5 Cont.
<p>Effects of UCP on the white blood cell content (<b>A</b>), lymphocyte content (<b>B</b>), and platelet content (<b>C</b>) in immunosuppressive mice induced by cyclophosphamide. The data were represented as the means ± SD (<span class="html-italic">n</span> = 10). Significant: ** <span class="html-italic">p</span> &lt; 0.01 vs. normal group; <sup>##</sup> <span class="html-italic">p</span> &lt; 0.05 vs. model group.</p>
Full article ">Figure 6
<p>The effects of UCP on the IgG (<b>A</b>), IgE (<b>B</b>), and IgM (<b>C</b>) levels in immunosuppressive mice induced by cyclophosphamide. The data were represented as the means ± SD (<span class="html-italic">n</span> = 10). Significant: ** <span class="html-italic">p</span> &lt; 0.01 vs. normal group; <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05, <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 vs. model group.</p>
Full article ">Figure 6 Cont.
<p>The effects of UCP on the IgG (<b>A</b>), IgE (<b>B</b>), and IgM (<b>C</b>) levels in immunosuppressive mice induced by cyclophosphamide. The data were represented as the means ± SD (<span class="html-italic">n</span> = 10). Significant: ** <span class="html-italic">p</span> &lt; 0.01 vs. normal group; <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05, <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 vs. model group.</p>
Full article ">
18 pages, 2839 KiB  
Article
Stimulating the Hematopoietic Effect of Simulated Digestive Product of Fucoidan from Sargassum fusiforme on Cyclophosphamide-Induced Hematopoietic Damage in Mice and Its Protective Mechanisms Based on Serum Lipidomics
by Wei-Ping Ma, Shi-Ning Yin, Jia-Peng Chen, Xi-Cheng Geng, Ming-Fei Liu, Hai-Hua Li, Ming Liu and Hong-Bing Liu
Mar. Drugs 2022, 20(3), 201; https://doi.org/10.3390/md20030201 - 9 Mar 2022
Cited by 7 | Viewed by 2967
Abstract
Hematopoietic damage is a serious side effect of cytotoxic drugs, and agents promoting hematopoiesis are quite important for decreasing the death rate in cancer patients. In our previous work, we prepared the simulated digestive product of fucoidan from Sargassum fusiforme, DSFF, and [...] Read more.
Hematopoietic damage is a serious side effect of cytotoxic drugs, and agents promoting hematopoiesis are quite important for decreasing the death rate in cancer patients. In our previous work, we prepared the simulated digestive product of fucoidan from Sargassum fusiforme, DSFF, and found that DSFF could activate macrophages. However, more investigations are needed to further evaluate whether DSFF could promote hematopoiesis in the chemotherapy process. In this study, the protective effect of DSFF (1.8–7.2 mg/kg, i.p.) on cyclophosphamide-induced hematopoietic damage in mice and the underlying mechanisms were investigated. Our results show that DSFF could restore the numbers of white blood cells, neutrophils, and platelets in the peripheral blood, and could also retard bone marrow cell decrease in mice with cyclophosphamide-induced hematopoietic damage. UPLC/Q-Extraction Orbitrap/MS/MS-based lipidomics results reveal 16 potential lipid biomarkers in a serum that responded to hematopoietic damage in mice. Among them, PC (20:1/14:0) and SM (18:0/22:0) were the key lipid molecules through which DSFF exerted protective actions. In a validation experiment, DSFF (6.25–100 μg/mL) could also promote K562 cell proliferation and differentiation in vitro. The current findings indicated that DSFF could affect the blood cells and bone marrow cells in vivo and thus showed good potential and application value in alleviating the hematopoietic damage caused by cyclophosphamide. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Effects of simulated digestive product of fucoidan from <span class="html-italic">Sargassum fusiforme</span> (DSFF) on the (<b>A</b>) body weight gain, (<b>B</b>) splenic index, and (<b>C</b>) thymic index of mice with cyclophosphamide-induced hematopoietic damage. ** <span class="html-italic">p</span> &lt; 0.01, vs. control; <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05, <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01, vs. model.</p>
Full article ">Figure 2
<p>Effects of DSFF on bone marrow DNA content in cyclophosphamide-induced mice. Data are the mean ± standard deviation (<span class="html-italic">n</span> = 10). ** <span class="html-italic">p</span> &lt; 0.01, vs. control; <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01, vs. model.</p>
Full article ">Figure 3
<p>PCA score plots of serum samples: (<b>A</b>) control vs. model in positive ion mode; (<b>B</b>) control vs. model in negative ion mode; (<b>C</b>) control vs. model vs. rhG-CSF in positive ion mode; (<b>D</b>) control vs. model vs. rhG-CSF in negative ion mode.</p>
Full article ">Figure 4
<p>OPLS-DA score plot of serum samples from control group and model group in (<b>A</b>) positive ion mode and (<b>B</b>) negative ion mode.</p>
Full article ">Figure 5
<p>Differential metabolite pathway analysis of serum samples in the control group and the model group. A: arachidonic acid metabolism; B: glycerophospholipid metabolism; C: alpha-linolenic acid metabolism; D: linoleic acid metabolism.</p>
Full article ">Figure 6
<p>PCA and OPLS-DA score plots of serum samples between the control group, model group, and 3.6 μg/mL DSFF group under positive ion mode. (<b>A</b>): PCA score plot; (<b>B</b>): OPLS-DA score plot.</p>
Full article ">Figure 7
<p>Effects of simulated digestive product of fucoidan from <span class="html-italic">Sargassum fusiforme</span> (DSFF) on the proliferation and differentiation of K562 cells. (<b>A</b>) Proliferative effect of DSFF on K562 cells using CCK-8 assay; (<b>B</b>) effects of DSFF on the differentiation of K562 cells using a benzidine-staining assay (10×, scale bar: 100 μm); (<b>C</b>) effects of DSFF on the erythroid differentiation of K562 cells using flow cytometry; (<b>D</b>) effects of DSFF on the megakaryocyte differentiation of K562 cells using flow cytometry; (<b>E</b>) effects of DSFF on the differentiation-related protein expression of K562 cells using Western blotting. Data are the mean ± standard deviation (<span class="html-italic">n</span> = 10). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, vs. control.</p>
Full article ">

2021

Jump to: 2024, 2022, 2020, 2019, 2018, 2017, 2016, 2015, 2014, 2013, 2012, 2011, 2010

24 pages, 4568 KiB  
Article
Agarose Stearate-Carbomer940 as Stabilizer and Rheology Modifier for Surfactant-Free Cosmetic Formulations
by Qiong Xiao, Guo Chen, Yong-Hui Zhang, Fu-Quan Chen, Hui-Fen Weng and An-Feng Xiao
Mar. Drugs 2021, 19(6), 344; https://doi.org/10.3390/md19060344 - 16 Jun 2021
Cited by 16 | Viewed by 3536
Abstract
Some commonly used surfactants in cosmetic products raise concerns due to their skin-irritating effects and environmental contamination. Multifunctional, high-performance polymers are good alternatives to overcome these problems. In this study, agarose stearate (AS) with emulsifying, thickening, and gel properties was synthesized. Surfactant-free cosmetic [...] Read more.
Some commonly used surfactants in cosmetic products raise concerns due to their skin-irritating effects and environmental contamination. Multifunctional, high-performance polymers are good alternatives to overcome these problems. In this study, agarose stearate (AS) with emulsifying, thickening, and gel properties was synthesized. Surfactant-free cosmetic formulations were successfully prepared from AS and carbomer940 (CBM940) mixed systems. The correlation of rheological parameter with skin feeling was determined to study the usability of the mixed systems in cosmetics. Based on rheological analysis, the surfactant-free cosmetic cream (SFC) stabilized by AS-carbomer940 showed shear-thinning behavior and strongly synergistic action. The SFC exhibited a gel-like behavior and had rheological properties similar to commercial cosmetic creams. Scanning electron microscope images proved that the AS-CBM940 network played an important role in SFC’s stability. Oil content could reinforce the elastic characteristics of the AS-CBM940 matrix. The SFCs showed a good appearance and sensation during and after rubbing into skin. The knowledge gained from this study may be useful for designing surfactant-free cosmetic cream with rheological properties that can be tailored for particular commercial cosmetic applications. They may also be useful for producing medicine products with highly viscous or gel-like textures, such as some ointments and wound dressings. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Characterization of AG and AS, (<b>A</b>) FT-IR spectra of AG, AS/DS 0.25, AS/DS 0.79, AS/DS1.01 and AS/uncleaned; (<b>B</b>) TGA curves of AG, AS/DS 0.25, AS/DS 0.79, AS/DS 1.01; (<b>C</b>) <sup>13</sup>C-NMR of AG, AS/DS 0.25, AS/DS 0.79, AS/DS 1.01.</p>
Full article ">Figure 2
<p>Emulsifying properties of AG and AS; (<b>A</b>) effect of AG and AS concentration on ST; (<b>B</b>) effect of AG and AS concentration on IT; (<b>C</b>) effect of AG and AS concentration on EA; (<b>D</b>) effect of AG and AS concentration on ESI.</p>
Full article ">Figure 3
<p>Viscosity curves and thixotropy curves for cosmetics and SFC and FTIR spectra of AS−CBM<sub>940</sub>. (<b>A</b>) Viscosity curves of SFC with different AS concentration; (<b>B</b>) viscosity curves of commercial moisturizers (Kiehl’s and SPDC) and lotions (MUJI and OSM); (<b>C</b>) thixotropy curves of emulsion gels with different AS concentration; (<b>D</b>) thixotropy curves of commercial moisturizers and lotions; (<b>E</b>) FTIR spectra of AS, CBM<sub>940</sub>, and AS−CBM<sub>940</sub>.</p>
Full article ">Figure 3 Cont.
<p>Viscosity curves and thixotropy curves for cosmetics and SFC and FTIR spectra of AS−CBM<sub>940</sub>. (<b>A</b>) Viscosity curves of SFC with different AS concentration; (<b>B</b>) viscosity curves of commercial moisturizers (Kiehl’s and SPDC) and lotions (MUJI and OSM); (<b>C</b>) thixotropy curves of emulsion gels with different AS concentration; (<b>D</b>) thixotropy curves of commercial moisturizers and lotions; (<b>E</b>) FTIR spectra of AS, CBM<sub>940</sub>, and AS−CBM<sub>940</sub>.</p>
Full article ">Figure 4
<p>Dynamical oscillatory frequency sweep test curves and creep–recovery curves of SFCs with various AS concentrations and commercial moisturizers and lotions. (<b>A</b>) G′ and G″ of SFCs with various AS concentrations; (<b>B</b>) tanδ of SFCs with various AS concentrations; (<b>C</b>) G′ and G″ of commercial moisturizers and lotions; (<b>D</b>) tanδ of commercial moisturizers and lotions; (<b>E</b>) creep−recovery curves of the moisturizers; (<b>F</b>) creep–recovery curves of the lotions; (<b>G</b>) creep–recovery curves of the SFCs.</p>
Full article ">Figure 4 Cont.
<p>Dynamical oscillatory frequency sweep test curves and creep–recovery curves of SFCs with various AS concentrations and commercial moisturizers and lotions. (<b>A</b>) G′ and G″ of SFCs with various AS concentrations; (<b>B</b>) tanδ of SFCs with various AS concentrations; (<b>C</b>) G′ and G″ of commercial moisturizers and lotions; (<b>D</b>) tanδ of commercial moisturizers and lotions; (<b>E</b>) creep−recovery curves of the moisturizers; (<b>F</b>) creep–recovery curves of the lotions; (<b>G</b>) creep–recovery curves of the SFCs.</p>
Full article ">Figure 5
<p>Flow behavior (<b>A</b>), thixotropy (<b>B</b>), dynamical viscoelasticity (<b>C</b>), and creep−recovery behaviors (<b>D</b>) of emulsion at different CBM<sub>940</sub> concentrations.</p>
Full article ">Figure 6
<p>Flow behavior (<b>A</b>), thixotropy (<b>B</b>), dynamical viscoelasticity (<b>C</b>), and creep−recovery behaviors (<b>D</b>) of SFCs at different pH levels.</p>
Full article ">Figure 6 Cont.
<p>Flow behavior (<b>A</b>), thixotropy (<b>B</b>), dynamical viscoelasticity (<b>C</b>), and creep−recovery behaviors (<b>D</b>) of SFCs at different pH levels.</p>
Full article ">Figure 7
<p>Flow behavior (<b>A</b>), thixotropy (<b>B</b>), dynamical viscoelasticity (<b>C</b>), creep–recovery behavior (<b>D</b>) and temperature ramp (<b>E1</b>): G′ curves of SFC for the gel melting process (up, 20−90 °C); (<b>E2</b>): G″ curves of SFC for the gel melting process; (<b>E3</b>): G′ curves of SFC for the gelation process (down, 90−20 °C); (<b>E4</b>): G″ curves of SFC for the gelation process).</p>
Full article ">Figure 7 Cont.
<p>Flow behavior (<b>A</b>), thixotropy (<b>B</b>), dynamical viscoelasticity (<b>C</b>), creep–recovery behavior (<b>D</b>) and temperature ramp (<b>E1</b>): G′ curves of SFC for the gel melting process (up, 20−90 °C); (<b>E2</b>): G″ curves of SFC for the gel melting process; (<b>E3</b>): G′ curves of SFC for the gelation process (down, 90−20 °C); (<b>E4</b>): G″ curves of SFC for the gelation process).</p>
Full article ">Figure 8
<p>Optical microscopy of SFC and SEM micrographs of the SFC network after deoiling and freeze-dried. (<b>A</b>) Oil volume fraction of 72%; (<b>B</b>) oil volume fraction of 54%; (<b>C</b>) oil volume fraction of 36%; (<b>D</b>) oil volume fraction of 18%; (<b>E</b>) oil volume fraction of 18%; (<b>F</b>) oil volume fraction of 72%.</p>
Full article ">Figure 9
<p>Quantitative results of the sensory evaluation of the investigated SFC (<b>A<sub>1</sub></b>,<b>A<sub>2</sub></b>,<b>B<sub>1</sub></b>,<b>C<sub>1</sub></b>) with various oil phase ratio and commercial cosmetic moisturizers (<b>B<sub>2</sub></b>,<b>C<sub>2</sub></b>).</p>
Full article ">Figure 9 Cont.
<p>Quantitative results of the sensory evaluation of the investigated SFC (<b>A<sub>1</sub></b>,<b>A<sub>2</sub></b>,<b>B<sub>1</sub></b>,<b>C<sub>1</sub></b>) with various oil phase ratio and commercial cosmetic moisturizers (<b>B<sub>2</sub></b>,<b>C<sub>2</sub></b>).</p>
Full article ">

2020

Jump to: 2024, 2022, 2021, 2019, 2018, 2017, 2016, 2015, 2014, 2013, 2012, 2011, 2010

14 pages, 4128 KiB  
Article
Isolation, Characterization and Bioactive Properties of Alkali-Extracted Polysaccharides from Enteromorpha prolifera
by Shifeng Zhao, Yuan He, Chungu Wang, Israa Assani, Peilei Hou, Yan Feng, Juanjuan Yang, Yehua Wang, Zhixin Liao and Songdong Shen
Mar. Drugs 2020, 18(11), 552; https://doi.org/10.3390/md18110552 - 6 Nov 2020
Cited by 23 | Viewed by 2973
Abstract
Four new purified polysaccharides (PAP) were isolated and purified from the Enteromorpha prolifera by alkali extraction, and further characterization was investigated. Their average molecular weights of PAP-1, PAP-2, PAP-3, and PAP-4 were estimated as 3.44 × 104, 6.42 × 104 [...] Read more.
Four new purified polysaccharides (PAP) were isolated and purified from the Enteromorpha prolifera by alkali extraction, and further characterization was investigated. Their average molecular weights of PAP-1, PAP-2, PAP-3, and PAP-4 were estimated as 3.44 × 104, 6.42 × 104, 1.20 × 105, and 4.82 × 104 Da, respectively. The results from monosaccharide analysis indicated that PAP-1, PAP-2, PAP-3 were acidic polysaccharides and PAP-4 was a neutral polysaccharide. PAP-1 and PAP-2 mainly consist of galacturonic acid, while PAP-3 and PAP-4 mainly contained rhamnose. Congo red test showed that no triple helical structure was detected in the four polysaccharides. The antioxidant activities were investigated using 1,1-diphenyl-2-picrylhydrazyl (DPPH), Superoxide, and 2, 2′-azino-bis(3-ethylbenzothiazoline-6-sulfonic acid) (ABTS) radical assay. In vitro antitumor activities were evaluated by 3-(4,5-Dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide (MTT) assay. PAP-1 exhibited relatively stronger antioxidant activities among the four polysaccharides in a dose-dependent manner. At a concentration of 1.00 mg/mL, the antioxidant activities of PAP-1 on the DPPH radical scavenging rate, superoxide anion radical scavenging rate, and ABTS radical rate at 1.00 mg/mL were 56.40%, 54.27%, and 42.07%, respectively. They also showed no significant inhibitory activity against MGC-803, HepG2, T24, and Bel-7402 cells. These investigations of polysaccharides provide a scientific basis for the use of E. prolifera as an ingredient in functional foods and medicines. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Elution profile of AP by DEAE-52 cellulose column chromatography.</p>
Full article ">Figure 2
<p>Elution profile of PAP by gel filtration chromatography on Sephadex G-100 (<b>a</b>): PAP-1, (<b>b</b>): PAP-2, (<b>c</b>): PAP-3, (<b>d</b>): PAP-4.</p>
Full article ">Figure 3
<p>HPGPC chromatograms of purified polysaccharides (<b>a</b>): PAP-1, (<b>b</b>): PAP-2, (<b>c</b>): PAP-3, (<b>d</b>): PAP-4.</p>
Full article ">Figure 4
<p>The HPLC chromatogram of standard monosaccharides and purified polysaccharides.</p>
Full article ">Figure 5
<p>UV (<b>a</b>) and FT-IR (<b>b</b>) spectra of the purified polysaccharides.</p>
Full article ">Figure 6
<p>The UV absorbance spectrum of purified polysaccharides in Congo red assay. Each value represents the mean ± SD (<span class="html-italic">n</span> = 3); * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 7
<p>SEM micrographs (×6000) of polysaccharides, (<b>a</b>): PAP-1, (<b>b</b>): PAP-2, (<b>c</b>): PAP-3, (<b>d</b>): PAP-4.</p>
Full article ">Figure 8
<p>Scavenging effects of polysaccharides: (<b>a</b>) DPPH assay in 15 min, (<b>b</b>) DPPH assay in 30 min, (<b>c</b>) Superoxide radicals assay, (<b>d</b>) ABTS radicals assay, (<b>e</b>) Inhibitory on H<sub>2</sub>O<sub>2</sub> Induces Red Blood Cell Hemolysis assay. Each value represents the mean ± SD (<span class="html-italic">n</span> = 3), * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">
13 pages, 3091 KiB  
Article
Sargassum fusiforme Polysaccharides Prevent High-Fat Diet-Induced Early Fasting Hypoglycemia and Regulate the Gut Microbiota Composition
by Bin Wei, Qi-Wu Zhong, Song-Ze Ke, Tao-Shun Zhou, Qiao-Li Xu, Si-Jia Wang, Jian-Wei Chen, Hua-Wei Zhang, Wei-Hua Jin and Hong Wang
Mar. Drugs 2020, 18(9), 444; https://doi.org/10.3390/md18090444 - 27 Aug 2020
Cited by 19 | Viewed by 3411
Abstract
A low fasting blood glucose level is a common symptom in diabetes patients and can be induced by high-fat diet (HFD) feeding at an early stage, which may play important roles in the development of diabetes, but has received little attention. In this [...] Read more.
A low fasting blood glucose level is a common symptom in diabetes patients and can be induced by high-fat diet (HFD) feeding at an early stage, which may play important roles in the development of diabetes, but has received little attention. In this study, five polysaccharides were prepared from Sargassumfusiforme and their effects on HFD-induced fasting hypoglycemia and gut microbiota dysbiosis were investigated. The results indicated that C57BL/6J male mice fed an HFD for 4 weeks developed severe hypoglycemia and four Sargassumfusiforme polysaccharides (SFPs), consisting of Sf-2, Sf-3, Sf-3-1, and Sf-A, significantly prevented early fasting hypoglycemia without inducing hyperglycemia. Sf-1 and Sf-A could also significantly prevent HFD-induced weight gain. Sf-2, Sf-3, Sf-3-1, and Sf-A mainly attenuated the HFD-induced decrease in Bacteroidetes, and all five SFPs had a considerable influence on the relative abundance of Oscillospira, Mucispirillum, and Clostridiales. Correlation analysis revealed that the fasting blood glucose level was associated with the relative abundance of Mucispinllum and Oscillospira. Receiver operating characteristic analysis indicated that Mucispinllum and Oscillospira exhibited good discriminatory power (AUC = 0.745–0.833) in the prediction of fasting hypoglycemia. Our findings highlight the novel application of SFPs (especially Sf-A) in glucose homeostasis and the potential roles of Mucispinllum and Oscillospira in the biological activity of SFPs. Full article
Show Figures

Figure 1

Figure 1
<p>Effects of <span class="html-italic">S. fusiforme</span> polysaccharides on the (<b>A</b>) body weight, (<b>B</b>) fasting blood glucose, (<b>C</b>) oral glucose tolerance test (OGTT), (<b>D</b>) liver, (<b>E</b>) epididymal fat, and (<b>F</b>) pancreas weight in high-fat diet (HFD)-treated mice. Values are the mean ± SD (n = 10). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001 vs. blank. <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05, <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01, and <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 vs. control.</p>
Full article ">Figure 2
<p>Effects of <span class="html-italic">S. fusiforme</span> polysaccharides on the composition and α-diversity of the gut microbiota in HFD-treated mice. (<b>A</b>) Gut microbiota composition at the phylum level, (<b>B</b>) Abundance-based Coverage Estimator (ACE) metric, (<b>C</b>) Chao1 diversity index, (<b>D</b>) Simpson’s diversity index, and (<b>E</b>) Shannon’s diversity index of the gut microbiota. Values are the mean ± SD (n = 10). *** <span class="html-italic">p</span> &lt; 0.001 vs. blank. <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05 vs. control.</p>
Full article ">Figure 3
<p>Effects of <span class="html-italic">S. fusiforme</span> polysaccharides on the relative abundance of (<b>A</b>) Bacteroidetes, (<b>B</b>) Firmicutes, (<b>C</b>) Proteobacteria, (<b>D</b>) Actinobacteria, (<b>E</b>) Deferribacteres, and (<b>F</b>) Verrucomicrobia in the gut microbiota in HFD-treated mice. Values are the mean ± SD (n = 10). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001 vs. blank. <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05, <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 vs. control.</p>
Full article ">Figure 4
<p>Effects of <span class="html-italic">S. fusiforme</span> polysaccharides on the relative abundance of (<b>A</b>) o_<span class="html-italic">Clostridiales</span>._._, (<b>B</b>) g_<span class="html-italic">Oscillospira</span>, (<b>C</b>) g_<span class="html-italic">Mucispirillum</span>, (<b>D</b>) f_<span class="html-italic">Coriobacteriaceae</span>.g_, (<b>E</b>) g_<span class="html-italic">Moryella</span>, and (<b>F</b>) g_<span class="html-italic">Bifidobacterium</span> in the gut microbiota in HFD-treated mice. Values are the mean ± SD (n = 10). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001 vs. blank. <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05, <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01, and <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 vs. control.</p>
Full article ">Figure 5
<p>Effects of <span class="html-italic">S. fusiforme</span> polysaccharides on the (<b>A</b>) gut microbiota composition at the genus level, (<b>B</b>) predicted functional gene, (<b>C</b>) metabolic pathway coverage, and (<b>D</b>) abundance in HFD-treated mice.</p>
Full article ">Figure 6
<p>(<b>A</b>) Correlations between the fasting blood glucose level and the relative abundance of gut microbiota. A color key is shown at the bottom right of the heatmap to demonstrate the size of the correlation coefficient. Values in each lattice represent the correlation coefficients. * False discovery rate (FDR)-corrected <span class="html-italic">p</span>-value &lt; 0.05; ** FDR-corrected <span class="html-italic">p</span>-value &lt; 0.01; *** FDR-corrected <span class="html-italic">p</span>-value &lt; 0.05. NA, not applicable. (<b>B</b>) Fitting receiver operating characteristic (ROC) curves of g_<span class="html-italic">Oscillospira</span>, g_<span class="html-italic">Bifidobacterium</span>, f_<span class="html-italic">Coriobacteriaceae</span>.g_, and g_<span class="html-italic">Mucispinllum</span> for the prediction of fasting hypoglycemia.</p>
Full article ">
16 pages, 4520 KiB  
Article
Immunomodulatory Effects of N-Acetyl Chitooligosaccharides on RAW264.7 Macrophages
by Jun-Jin Deng, Zong-Qiu Li, Ze-Quan Mo, Shun Xu, He-Hua Mao, Dan Shi, Zhi-Wei Li, Xue-Ming Dan and Xiao-Chun Luo
Mar. Drugs 2020, 18(8), 421; https://doi.org/10.3390/md18080421 - 12 Aug 2020
Cited by 35 | Viewed by 4426
Abstract
The ongoing development of new production methods may lead to the commercialization of N-acetyl chitooligosaccharides (NACOS), such as chitosan oligosaccharides (COS). The bioactivity of NACOS, although not well detailed, differs from that of COS, as they have more acetyl groups than COS. [...] Read more.
The ongoing development of new production methods may lead to the commercialization of N-acetyl chitooligosaccharides (NACOS), such as chitosan oligosaccharides (COS). The bioactivity of NACOS, although not well detailed, differs from that of COS, as they have more acetyl groups than COS. We used two enzymatically produced NACOS with different molecular compositions and six NACOS (NACOS1–6) with a single degree of polymerization to verify their immunomodulatory effects on the RAW264.7 macrophage cell line. We aimed to identify any differences between COS and various NACOS with a single degree of polymerization. The results showed that NACOS had similar immune enhancement effects on RAW264.7 cells as COS, including the generation of reactive oxygen species (ROS), phagocytotic activity, and the production of pro-inflammation cytokines (IL-1β, IL-6, and TNF-α). However, unlike COS and lipopolysaccharide (LPS), NACOS1 and NACOS6 significantly inhibited nitric oxide (NO) production. Besides their immune enhancement effects, NACOS also significantly inhibited the LPS-induced RAW264.7 inflammatory response with some differences between various polymerization degrees. We confirmed that the NF-κB pathway is associated with the immunomodulatory effects of NACOS on RAW264.7 cells. This study could inform the application of NACOS with varying different degrees of polymerization in human health. Full article
Show Figures

Figure 1

Figure 1
<p>Effects of <span class="html-italic">N</span>-acetyl chitooligosaccharides (NACOS) on production of nitric oxide (NO), reactive oxygen species (ROS), gene expression of <span class="html-italic">iNOS</span> and <span class="html-italic">NOX2</span>, and phagocytosis of RAW264.7 cells. Cells were treated with NACOS (100 µg/mL) or lipopolysaccharide (LPS) (1 µg/mL). (<b>A</b>) NO production following treatment with NACOS for 24 h. (<b>B</b>) Expression of the <span class="html-italic">iNOS</span> gene following treatment with NACOS for 8 h. (<b>C</b>) ROS production following treatment with NACOS for 4 h. (<b>D</b>) Expression of the <span class="html-italic">NOX2</span> gene following treatment with NACOS for 4 h. (<b>E</b>) Phagocytosis of beads by RAW 264.7 cells, as detected by flow cytometry, following treatment with NACOS for 24 h. Only the control (Ctrl), LPS, and NACOS6 treatments are shown (chitosan oligosaccharides (COS) and NACOS1–5 are presented in <a href="#app1-marinedrugs-18-00421" class="html-app">Supplementary Figure S3</a>). (<b>F</b>) Histogram of phagocytosis rates following treatment with NACOS. Data were normalized to that of β-actin and presented as the fold induction compared with the control group. Values are presented as the mean ± SD (<span class="html-italic">n</span> = 3). (* <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001, compared with the control group).</p>
Full article ">Figure 2
<p>Effects of <span class="html-italic">N</span>-acetyl chitooligosaccharides (NACOS) on gene expression of <span class="html-italic">TNF-α</span>, <span class="html-italic">IL-6</span>, <span class="html-italic">IL-1β</span>, and protein level of IL-1β in RAW264.7 macrophages. The RAW264.7 cells were treated with NACOS (100 µg/mL) or LPS (1 µg/mL) for 8 h. Gene expressions were measured by RT-qPCR. Gene expressions of (<b>A</b>) <span class="html-italic">TNF-α</span>; (<b>B</b>) <span class="html-italic">IL-6</span>; and (<b>C</b>) <span class="html-italic">IL-1β</span>, following treatment with NACOS. (<b>D</b>) Western blot analysis of IL-1β following treatment with NACOS for 8 h. (<b>E</b>) IL-1β protein levels following treatment with NACOS for 8 h, evaluated by densitometry analysis of the blot. Data were normalized to that of β-actin and are presented as the fold induction compared with the control group. Values are presented as the mean ± SD (<span class="html-italic">n</span> = 3). * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001, compared with the control group.</p>
Full article ">Figure 3
<p>Effects of <span class="html-italic">N</span>-acetyl chitooligosaccharides (NACOS) on the phosphorylation of IκBα and NF-κB p65 in RAW264.7 macrophages. The RAW 264.7 cells were treated with similar concentrations of the various NACOS (100 µg/mL), chitosan oligosaccharides (COS) (100 µg/mL), or lipopolysaccharide (LPS) (1 µg/mL) for 8 h. (<b>A</b>) Western blot analysis of cellular IκBα and phosphorylated IκBα. (<b>B</b>) Cellular phosphorylated IκBα protein levels based on densitometry analysis of the blot. (<b>C</b>) Western blot analysis of nucleic phosphorylated NF-κB p65. (<b>D</b>) Nucleic phosphorylated NF-κB p65 protein levels based on densitometry analysis of the blot. Data were normalized to β-actin or Lamin B1 and presented as the induction fold, compared with the control group. Values are presented as the mean ± SD (<span class="html-italic">n</span> = 3). ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001, compared with the control group.</p>
Full article ">Figure 4
<p>Effects of <span class="html-italic">N</span>-acetyl chitooligosaccharides (NACOS) pretreatment on the production of NO, reactive oxygen species (ROS), gene expression of <span class="html-italic">iNOS</span> and <span class="html-italic">NOX2</span>, and phagocytosis in lipopolysaccharide (LPS)-treated RAW264.7 macrophages. Cells were pretreated with NACOS (100 µg/mL), chitosan oligosaccharides (COS) (100 µg/mL), or lipopolysaccharide (LPS) (1 µg/mL) for 24 h and then co-incubated with LPS. Gene expressions were measured by RT-qPCR. (<b>A</b>) Production of NO following treatment with NACOS for 24 h and LPS for 24 h. (<b>B</b>) Expression of the <span class="html-italic">iNOS</span> gene following treatment with NACOS for 24 h and LPS for 8 h. (<b>C</b>) ROS production following treatment with NACOS for 24 h and LPS for 4 h. (<b>D</b>) Expression of the <span class="html-italic">NOX2</span> gene following treatment with NACOS for 24 h and LPS for 4 h. (<b>E</b>) Phagocytosis by RAW 264.7 cells, as detected by flow cytometry following treatment with NACOS for 24 h and LPS for 4 h. Phagocytosis of beads was determined by flow cytometry. Only the control (Ctrl), LPS, and NACOS6 + LPS treatments are shown (COS + LPS and NACOS1 to NACOS5 + LPS are presented in <a href="#app1-marinedrugs-18-00421" class="html-app">Supplementary Figure S4</a>). (<b>F</b>) Histogram of phagocytosis rates following pretreatment of RAW264.7 cells with NACOS. Data were normalized to β-actin and presented as the fold induction, compared with the control group. Values are presented as the mean ± SD (<span class="html-italic">n</span> = 3). * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01, compared with the only LPS-treated group.</p>
Full article ">Figure 5
<p>Effects of <span class="html-italic">N</span>-acetyl chitooligosaccharides (NACOS) pretreatment on gene expression of <span class="html-italic">TNF-α</span>, <span class="html-italic">IL-6</span>, <span class="html-italic">IL-1β</span>, and <span class="html-italic">IL-1β</span> and protein levels in lipopolysaccharide (LPS)-induced RAW264.7 macrophages. RAW-264.7 cells were pretreated with NACOS (100 µg/mL), chitosan oligosaccharides (COS) (100 µg/mL), or LPS (1 µg/mL) for 24 h, and then further co-incubated with LPS (1 µg/mL). Gene expressions were measured by RT-qPCR. Gene expressions of (<b>A</b>) <span class="html-italic">TNF-α</span>; (<b>B</b>) <span class="html-italic">IL-6</span>; and (<b>C</b>) <span class="html-italic">IL-1β</span> following treatment with NACOS for 24 h and LPS for 8 h. (<b>D</b>) Western blot analysis of IL-1β following treatment with NACOS for 24 h and LPS for 8 h. (<b>E</b>) The IL-1β levels based on densitometry analysis of the blot. Values are presented as the mean ± SD (<span class="html-italic">n</span> = 3). * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001, compared with the only LPS-treated group.</p>
Full article ">Figure 6
<p>Effects of <span class="html-italic">N</span>-acetyl chitooligosaccharides (NACOS) pretreatment on the phosphorylation of IκBα and NF-κB p65 in lipopolysaccharide (LPS)-induced RAW264.7 macrophages. RAW 264.7 cells were pretreated with the similar concentrations of NACOS (100 µg/mL), chitosan oligosaccharides (COS) (100 µg/mL), or LPS (1 µg/mL) for 24 h and then co-incubated with LPS for 8 h. (<b>A</b>) Western blot analysis of cellular IκBα and phosphorylated IκBα. (<b>B</b>) Cellular phosphorylated IκBα protein levels based on densitometry analysis of the blot. (<b>C</b>) Western blot analysis of nucleic phosphorylated NF-κB p65. (<b>D</b>) Nucleic phosphorylated NF-κB p65 protein levels based on densitometry analysis of the blot. Data were normalized to β-actin or Lamin B1, and presented as the induction fold, compared with the control group. Values are presented as the mean ± SD (<span class="html-italic">n</span> = 3). * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001, compared with the only LPS-treated group.</p>
Full article ">
26 pages, 1315 KiB  
Review
Potential Beneficial Actions of Fucoidan in Brain and Liver Injury, Disease, and Intoxication—Potential Implication of Sirtuins
by Jasmina Dimitrova-Shumkovska, Ljupcho Krstanoski and Leo Veenman
Mar. Drugs 2020, 18(5), 242; https://doi.org/10.3390/md18050242 - 5 May 2020
Cited by 37 | Viewed by 6590
Abstract
Increased interest in natural antioxidants has brought to light the fucoidans (sulfated polysaccharides present in brown marine algae) as highly valued nutrients as well as effective and safe therapeutics against several diseases. Based on their satisfactory in vitro antioxidant potency, researchers have identified [...] Read more.
Increased interest in natural antioxidants has brought to light the fucoidans (sulfated polysaccharides present in brown marine algae) as highly valued nutrients as well as effective and safe therapeutics against several diseases. Based on their satisfactory in vitro antioxidant potency, researchers have identified this molecule as an efficient remedy for neuropathological as well as metabolic disorders. Some of this therapeutic activity is accomplished by upregulation of cytoprotective molecular pathways capable of restoring the enzymatic antioxidant activity and normal mitochondrial functions. Sirtuin-3 has been discovered as a key player for achieving the neuroprotective role of fucoidan by managing these pathways, whose ultimate goal is retrieving the entirety of the antioxidant response and preventing apoptosis of neurons, thereby averting neurodegeneration and brain injuries. Another pathway whereby fucoidan exerts neuroprotective capabilities is by interactions with P-selectin on endothelial cells, thereby preventing macrophages from entering the brain proper. Furthermore, beneficial influences of fucoidan have been established in hepatocytes after xenobiotic induced liver injury by decreasing transaminase leakage and autophagy as well as obtaining optimal levels of intracellular fiber, which ultimately prevents fibrosis. The hepatoprotective role of this marine polysaccharide also includes a sirtuin, namely sirtuin-1 overexpression, which alleviates obesity and insulin resistance through suppression of hyperglycemia, reducing inflammation and stimulation of enzymatic antioxidant response. While fucoidan is very effective in animal models for brain injury and neuronal degeneration, in general, it is accepted that fucoidan shows somewhat limited potency in liver. Thus far, it has been used in large doses for treatment of acute liver injuries. Thus, it appears that further optimization of fucoidan derivatives may establish enhanced versatility for treatments of various disorders, in addition to brain injury and disease. Full article
Show Figures

Figure 1

Figure 1
<p>Mechanism of action of fucoidan in traumatic brain injury (TBI). Fucoidan alleviates brain injury through upregulation of sirtuin, which decreases reactive oxygen species (ROS) overproduction by inhibiting the mitochondrial permeability transition pore (mPTP) opening, and restores normal mitochondrial function via stimulation of ATP synthesis, and attenuates mitochondria-initiated apoptosis by decreasing leakage of cytochrome c from the mitochondria into the cytosol. Additionally, fucoidan stimulates expression of FOXO3A and Nrf-2-ARE genes, thus increasing glutathione (GSH) production and Mn-SOD and Cat activity.</p>
Full article ">Figure 2
<p>Effects of fucoidan on brain disease. Fucoidan reduces inflammatory response in brain diseases by inhibiting microglial activation, thus resulting in significantly decreased neuronal and astrocyte degeneration due to diminishing production of pro-apoptotic agents and improving antioxidant responses of the cell. Furthermore, fucoidan prevents leukocyte adhesion to the brain by blocking P-selectin.</p>
Full article ">Figure 3
<p>Effects of fucoidan on liver injury. Damaging agents at the top left side, APAP and carbon tetrachloride (CCl<sub>4</sub>), and protective fucoidan at the top right side. Fucoidan averts liver fibrosis by inhibiting HSCs production through optimal synthesis of collagen and alpha smooth muscle actin and prevents tissue damage by reducing transaminase release and restoring antioxidant potentials of cells. It decreases CYP2E1 activity, which reduces levels of toxic metabolites and inhibits TGF-β/Smad pathway, thereby hindering the occurrence of autophagosomes. Fucoidan also stimulates expression of sirtuin-1 in the liver, which activates AMPK and alleviates insulin resistance.</p>
Full article ">
14 pages, 1935 KiB  
Article
Anti-Inflammatory Effects of Fucoxanthinol in LPS-Induced RAW264.7 Cells through the NAAA-PEA Pathway
by Wenhui Jin, Longhe Yang, Zhiwei Yi, Hua Fang, Weizhu Chen, Zhuan Hong, Yiping Zhang, Guangya Zhang and Long Li
Mar. Drugs 2020, 18(4), 222; https://doi.org/10.3390/md18040222 - 21 Apr 2020
Cited by 23 | Viewed by 4344
Abstract
Palmitoylethanolamide (PEA) is an endogenous lipid mediator with powerful anti-inflammatory and analgesic functions. PEA can be hydrolyzed by a lysosomal enzyme N-acylethanolamine acid amidase (NAAA), which is highly expressed in macrophages and other immune cells. The pharmacological inhibition of NAAA activity is a [...] Read more.
Palmitoylethanolamide (PEA) is an endogenous lipid mediator with powerful anti-inflammatory and analgesic functions. PEA can be hydrolyzed by a lysosomal enzyme N-acylethanolamine acid amidase (NAAA), which is highly expressed in macrophages and other immune cells. The pharmacological inhibition of NAAA activity is a potential therapeutic strategy for inflammation-related diseases. Fucoxanthinol (FXOH) is a marine carotenoid from brown seaweeds with various beneficial effects. However, the anti-inflammatory effects and mechanism of action of FXOH in lipopolysaccharide (LPS)-stimulated macrophages remain unclear. This study aimed to explore the role of FXOH in the NAAA–PEA pathway and the anti-inflammatory effects based on this mechanism. In vitro results showed that FXOH can directly bind to the active site of NAAA protein and specifically inhibit the activity of NAAA enzyme. In an LPS-induced inflammatory model in macrophages, FXOH pretreatment significantly reversed the LPS-induced downregulation of PEA levels. FXOH also substantially attenuated the mRNA expression of inflammatory factors, including inducible nitric oxide synthase (iNOS), interleukin-6 (IL-6) and tumor necrosis factor-alpha (TNF-α), and markedly reduced the production of TNF-α, IL-6, IL-1β, and nitric oxide (NO). Moreover, the inhibitory effect of FXOH on NO induction was significantly abolished by the peroxisome proliferator-activated receptor α (PPAR-α) inhibitor GW6471. All these findings demonstrated that FXOH can prevent LPS-induced inflammation in macrophages, and its mechanisms may be associated with the regulation of the NAAA-PEA-PPAR-α pathway. Full article
Show Figures

Figure 1

Figure 1
<p>Inhibitory effects of fucoxanthinol (FXOH) on N-acylethanolamine acid amidase (NAAA) and fatty acid amide hydrolase (FAAH) activity. (<b>A</b>) Dose-dependent effects of FXOH on NAAA (filled circles) and FAAH activity (filled triangles). (<b>B</b>) Dose-dependent effects of FX on NAAA activity (filled circles). (<b>C</b>) Effects of FXOH on cell viability.</p>
Full article ">Figure 2
<p>Interactions of FXOH with NAAA. (<b>A</b>) Chemical structure of FXOH; (<b>B</b>) interaction mode between FXOH and the active site of NAAA; (<b>C</b>) planar view of interactions between FXOH and NAAA; and (<b>D</b>) planar view of interactions between Arn19702 and NAAA.</p>
Full article ">Figure 3
<p>Effects of FXOH on fatty acid ethanolamide levels in LPS-activated macrophage. (<b>A</b>) LC–MS/MS chromatogram (<b>B</b>–<b>F</b>) palmitoylethanolamide (PEA), oleoylethanolamide (OEA), stearoylethanolamide (SEA), anandamide (AEA), and 2-AG levels were measured by LC–MS/MS assay. RAW264.7 cells were pre-treated with FXOH and then stimulated with lipopolysaccharide (LPS) for 24 h. Data represent the mean ± SEM. ** <span class="html-italic">P</span> &lt; 0.01 compared with the control group, # <span class="html-italic">P</span> &lt; 0.05 compared with the LPS + vehicle group.</p>
Full article ">Figure 4
<p>Effects of FXOH on the mRNA expression of inflammatory cytokines. (<b>A</b>) iNOS, (<b>B</b>) IL-6, and (<b>C</b>) TNF-α mRNA expression levels were determined by real-time PCR. RAW264.7 cells were pre-treated with FXOH (20, 10, and 5 µM) and then stimulated with LPS (100 ng/mL) for 24 h. Data represent the mean ± SEM. *** <span class="html-italic">P</span> &lt; 0.001 compared with the control group, # <span class="html-italic">P</span> &lt; 0.05 compared with the LPS group (received the treatment of LPS and the vehicle of FXOH), ## <span class="html-italic">P</span> &lt; 0.01 compared with the LPS group, ### <span class="html-italic">P</span> &lt; 0.001 compared with the LPS group.</p>
Full article ">Figure 5
<p>Effects of FXOH on cytokines protein levels and NO production in macrophage. (<b>A</b>) TNF-α, (<b>B</b>) IL-6, and (<b>C</b>) IL-1β protein levels were measured by ELISA; (<b>D</b>) nitrate expression in culture media was detected using a Griess Reagent kit. RAW264.7 cells were treated with 20, 10, and 5 µM FXOH 2 h prior to LPS (100 ng/ml) stimulation. Data are expressed as mean ± SEM. *** <span class="html-italic">P</span> &lt; 0.001 compared with the control group, # <span class="html-italic">P</span> &lt; 0.05 compared with the LPS group (received the treatment of LPS and the vehicle of FXOH), ## <span class="html-italic">P</span> &lt; 0.01 compared with the LPS group, ### <span class="html-italic">P</span> &lt; 0.001 compared with the LPS group.</p>
Full article ">Figure 6
<p>Effects of FXOH on LPS-induced macrophage activation were mediated by PPAR-α. (<b>A</b>) PPAR-α protein levels were measured by Western blot analysis and normalized against β-actin expression levels. (<b>B</b>) Nitrate concentrations in the culture media were evaluated using a Griess reagent kit. Data represent the mean ± SEM. *** <span class="html-italic">P</span> &lt; 0.001 compared with the control group, ## <span class="html-italic">P</span> &lt; 0.01 compared with the LPS + vehicle group.</p>
Full article ">
13 pages, 1892 KiB  
Article
Identification of a Key Enzyme for the Hydrolysis of β-(1→3)-Xylosyl Linkage in Red Alga Dulse Xylooligosaccharide from Bifidobacterium Adolescentis
by Manami Kobayashi, Yuya Kumagai, Yohei Yamamoto, Hajime Yasui and Hideki Kishimura
Mar. Drugs 2020, 18(3), 174; https://doi.org/10.3390/md18030174 - 20 Mar 2020
Cited by 21 | Viewed by 4308
Abstract
Red alga dulse possesses a unique xylan, which is composed of a linear β-(1→3)/β-(1→4)-xylosyl linkage. We previously prepared characteristic xylooligosaccharide (DX3, (β-(1→3)-xylosyl-xylobiose)) from dulse. In this study, we evaluated the prebiotic effect of DX3 on enteric bacterium. Although DX3 was utilized by Bacteroides [...] Read more.
Red alga dulse possesses a unique xylan, which is composed of a linear β-(1→3)/β-(1→4)-xylosyl linkage. We previously prepared characteristic xylooligosaccharide (DX3, (β-(1→3)-xylosyl-xylobiose)) from dulse. In this study, we evaluated the prebiotic effect of DX3 on enteric bacterium. Although DX3 was utilized by Bacteroides sp. and Bifidobacterium adolescentis, Bacteroides Ksp. grew slowly as compared with β-(1→4)-xylotriose (X3) but B. adolescentis grew similar to X3. Therefore, we aimed to find the key DX3 hydrolysis enzymes in B. adolescentis. From bioinformatics analysis, two enzymes from the glycoside hydrolase family 43 (BAD0423: subfamily 12 and BAD0428: subfamily 11) were selected and expressed in Escherichia coli. BAD0423 hydrolyzed β-(1→3)-xylosyl linkage in DX3 with the specific activity of 2988 mU/mg producing xylose (X1) and xylobiose (X2), and showed low activity on X2 and X3. BAD0428 showed high activity on X2 and X3 producing X1, and the activity of BAD0428 on DX3 was 1298 mU/mg producing X1. Cooperative hydrolysis of DX3 was found in the combination of BAD0423 and BAD0428 producing X1 as the main product. From enzymatic character, hydrolysis of X3 was completed by one enzyme BAD0428, whereas hydrolysis of DX3 needed more than two enzymes. Full article
Show Figures

Figure 1

Figure 1
<p>Effect of carbohydrate on bacterial growth and pH. The bacterial strains were cultured at 37 °C for 96 h in PYF (Peptone−Yeast extract−Fildes) medium containing 0.5% carbohydrate samples under anaerobic conditions. The data were obtained as <span class="html-italic">Δ</span>OD<sub>600</sub> and <span class="html-italic">Δ</span>pH by the subtraction of each bacterial growth without carbohydrate samples. Error bars indicate SD (n = 3). Growth data of <span class="html-italic">C</span>. <span class="html-italic">paraputrificum</span> and <span class="html-italic">E</span>. <span class="html-italic">limosum</span> for β-(1→4)-xylotriose (X3) were not determined. Symbols: black bar, glucose; gray bar, xylose; red bar, X3; and blue bar, β-(1→3)/β-(1→4)-xylotriose (DX3).</p>
Full article ">Figure 2
<p>Time course growth rate of <span class="html-italic">B</span>. <span class="html-italic">vulgatus</span> (<b>a</b>) and <span class="html-italic">B</span>. <span class="html-italic">adolescentis</span> (<b>b</b>). Bacteria grow in PYF medium containing 0.5% carbohydrates at 37 °C in anaerobic condition. Symbols: ●, glucose (G1); ◯, xylose (X1); ▲, X3; △, DX3. Mean  ±  standard deviation of three replicate determinations and the error bars were within the symbols.</p>
Full article ">Figure 3
<p>Relationship between bacteria and the number of enzymes containing β-xylosidase (EC 3.2.1.37) in glycoside hydrolase families (GHs). The colors are related to the number of enzymes from pink (low) to red (high). The bold and underlined number means that GHs contain β-xylosidase.</p>
Full article ">Figure 4
<p>Relationship between bacteria and the number of enzymes in the GH43 subfamily. The colors are related to the number of enzymes from pink (low) to purple (high). The bold and underlined number shows subfamilies contain β-xylosidase.</p>
Full article ">Figure 5
<p>Hydrolysis products of oligosaccharides (XOS). Ten mM XOS was hydrolyzed by 50 μg/mL BAD0423, BAD0428, and BAD1527 at pH 6.5 and 37 °C for 1 h. The products were analyzed by HPLC. Black and gray indicate the amount of X1 and X2, respectively.</p>
Full article ">Figure 6
<p>The putative hydrolysis mechanism of DX3 and X3. (<b>a</b>) DX3; (<b>b</b>) X3.</p>
Full article ">
29 pages, 2151 KiB  
Review
Advanced Technologies for the Extraction of Marine Brown Algal Polysaccharides
by Ana Dobrinčić, Sandra Balbino, Zoran Zorić, Sandra Pedisić, Danijela Bursać Kovačević, Ivona Elez Garofulić and Verica Dragović-Uzelac
Mar. Drugs 2020, 18(3), 168; https://doi.org/10.3390/md18030168 - 18 Mar 2020
Cited by 154 | Viewed by 11590
Abstract
Over the years, brown algae bioactive polysaccharides laminarin, alginate and fucoidan have been isolated and used in functional foods, cosmeceutical and pharmaceutical industries. The extraction process of these polysaccharides includes several complex and time-consuming steps and the correct adjustment of extraction parameters (e.g., [...] Read more.
Over the years, brown algae bioactive polysaccharides laminarin, alginate and fucoidan have been isolated and used in functional foods, cosmeceutical and pharmaceutical industries. The extraction process of these polysaccharides includes several complex and time-consuming steps and the correct adjustment of extraction parameters (e.g., time, temperature, power, pressure, solvent and sample to solvent ratio) greatly influences the yield, physical, chemical and biochemical properties as well as their biological activities. This review includes the most recent conventional procedures for brown algae polysaccharides extraction along with advanced extraction techniques (microwave-assisted extraction, ultrasound assisted extraction, pressurized liquid extraction and enzymes assisted extraction) which can effectively improve extraction process. The influence of these extraction techniques and their individual parameters on yield, chemical structure and biological activities from the most current literature is discussed, along with their potential for commercial applications as bioactive compounds and drug delivery systems. Full article
Show Figures

Figure 1

Figure 1
<p>Structure of laminarin [<a href="#B26-marinedrugs-18-00168" class="html-bibr">26</a>].</p>
Full article ">Figure 2
<p>Structure of alginates [<a href="#B35-marinedrugs-18-00168" class="html-bibr">35</a>].</p>
Full article ">Figure 3
<p>Structure of fucoidan from <span class="html-italic">Fucus vesiculosus</span>, with a backbone of alternating (1→3)-linked α-L-fucopyranose and (1→4)-linked α-L-fucopyranose residues and the presence of sulfate groups on both <span class="html-italic">O</span>-2 and <span class="html-italic">O</span>-3 [<a href="#B41-marinedrugs-18-00168" class="html-bibr">41</a>].</p>
Full article ">Figure 4
<p>Schematic overview of essential steps for extraction of brown algae polysaccharides.</p>
Full article ">Figure 5
<p>Schematic diagram of process parameters, chemical structure properties, biological activity and potential industrial uses of brown algae polysaccharides.</p>
Full article ">
26 pages, 1286 KiB  
Review
Advances in Research on the Bioactivity of Alginate Oligosaccharides
by Maochen Xing, Qi Cao, Yu Wang, Han Xiao, Jiarui Zhao, Qing Zhang, Aiguo Ji and Shuliang Song
Mar. Drugs 2020, 18(3), 144; https://doi.org/10.3390/md18030144 - 28 Feb 2020
Cited by 153 | Viewed by 9055
Abstract
Alginate is a natural polysaccharide present in various marine brown seaweeds. Alginate oligosaccharide (AOS) is a degradation product of alginate, which has received increasing attention due to its low molecular weight and promising biological activity. The wide-ranging biological activity of AOS is closely [...] Read more.
Alginate is a natural polysaccharide present in various marine brown seaweeds. Alginate oligosaccharide (AOS) is a degradation product of alginate, which has received increasing attention due to its low molecular weight and promising biological activity. The wide-ranging biological activity of AOS is closely related to the diversity of their structures. AOS with a specific structure and distinct applications can be obtained by different methods of alginate degradation. This review focuses on recent advances in the biological activity of alginate and its derivatives, including their anti-tumor, anti-oxidative, immunoregulatory, anti-inflammatory, neuroprotective, antibacterial, hypolipidemic, antihypertensive, and hypoglycemic properties, as well as the ability to suppress obesity and promote cell proliferation and regulate plant growth. We hope that this review will provide theoretical basis and inspiration for the high-value research developments and utilization of AOS-related products. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic representation of the molecular structure of alginate oligosaccharide prepared by enzymatic degradation [<a href="#B50-marinedrugs-18-00144" class="html-bibr">50</a>].</p>
Full article ">Figure 2
<p>Signaling pathways involved in the macrophage activation effect of alginate-derived guluronate oligosaccharide [<a href="#B18-marinedrugs-18-00144" class="html-bibr">18</a>].</p>
Full article ">Figure 3
<p>Schematic representation of the molecular structure of saturated mannuronate oligomers prepared by acid hydrolysis [<a href="#B61-marinedrugs-18-00144" class="html-bibr">61</a>].</p>
Full article ">Figure 4
<p>Schematic representation of chemical structures of alginate-derived oligosaccharide prepared by oxidative degradation [<a href="#B19-marinedrugs-18-00144" class="html-bibr">19</a>]. (The average molecular weight of this AOS is about 1500 Da).</p>
Full article ">
13 pages, 4649 KiB  
Article
Impact of Prevalence Ratios of Chondroitin Sulfate (CS)- 4 and -6 Isomers Derived from Marine Sources in Cell Proliferation and Chondrogenic Differentiation Processes
by Estefanía López-Senra, Paula Casal-Beiroa, Miriam López-Álvarez, Julia Serra, Pío González, Jesus Valcarcel, José Antonio Vázquez, Elena F. Burguera, Francisco J. Blanco and Joana Magalhães
Mar. Drugs 2020, 18(2), 94; https://doi.org/10.3390/md18020094 - 31 Jan 2020
Cited by 17 | Viewed by 4544
Abstract
Osteoarthritis is the most prevalent rheumatic disease. During disease progression, differences have been described in the prevalence of chondroitin sulfate (CS) isomers. Marine derived-CS present a higher proportion of the 6S isomer, offering therapeutic potential. Accordingly, we evaluated the effect of exogenous supplementation [...] Read more.
Osteoarthritis is the most prevalent rheumatic disease. During disease progression, differences have been described in the prevalence of chondroitin sulfate (CS) isomers. Marine derived-CS present a higher proportion of the 6S isomer, offering therapeutic potential. Accordingly, we evaluated the effect of exogenous supplementation of CS, derived from the small spotted catshark (Scyliorhinus canicula), blue shark (Prionace glauca), thornback skate (Raja clavata) and bovine CS (reference), on the proliferation of osteochondral cell lines (MG-63 and T/C-28a2) and the chondrogenic differentiation of mesenchymal stromal cells (MSCs). MG-G3 proliferation was comparable between R. clavata (CS-6 intermediate ratio) and bovine CS (CS-4 enrichment), for concentrations below 0.5 mg/mL, defined as a toxicity threshold. T/C-28a2 proliferation was significantly improved by intermediate ratios of CS-6 and -4 isomers (S. canicula and R. clavata). A dose-dependent response was observed for S. canicula (200 µg/mL vs 50 and 10 µg/mL) and bovine CS (200 and 100 µg/mL vs 10 µg/mL). CS sulfation patterns discretely affected MSCs chondrogenesis; even though S. canicula and R. clavata CS up-regulated chondrogenic markers expression (aggrecan and collagen type II) these were not statistically significant. We demonstrate that intermediate values of CS-4 and -6 isomers improve cell proliferation and offer potential for chondrogenic promotion, although more studies are needed to elucidate its mechanism of action. Full article
Show Figures

Figure 1

Figure 1
<p>MG-63 osteoblast cell line proliferation (MTT assay), after 72 h incubation in minimum essential medium-Eagle with Earle’s balanced salt solution (EMEM) supplemented with different concentrations of chondroitin sulfate (CS) (50 µg/mL, 100 µg/mL, 0.5 mg/mL, 1 mg/mL and 10 mg/mL), from fish (<span class="html-italic">Prionace glauca</span>, <span class="html-italic">Raja clavata</span> and <span class="html-italic">Scyliorhinus canicula</span>) and bovine sources, normalized against CS-free condition which was assigned the value of 1. Significant statistical differences for <span class="html-italic">p</span> &lt; 0.05 (*) are shown.</p>
Full article ">Figure 2
<p>T/C-28a2 chondrocytic cell line viability and proliferation (directly proportional to alamar blue dye reduction) after 24, 48 and 72 h incubation in basal medium (Dulbecco’s modified Eagle’s medium (DMEM) 10%) supplemented with different concentrations of CS (10 µg/mL, 50 µg/mL, 100 µg/mL and 200 µg/mL) or CS-free, derived from fish (<span class="html-italic">Prionace glauca</span>, <span class="html-italic">Raja clavata</span> and <span class="html-italic">Scyliorhinus canicula</span>) and bovine sources. Significant statistical differences for <span class="html-italic">p</span> &lt; 0.05 (*) and <span class="html-italic">p</span> &lt; 0.01 (**) are shown. No significant differences were found amongst the different CS sources and CS-free condition.</p>
Full article ">Figure 3
<p>Histological analysis for morphology (haematoxylin and eosin, HE), proteoglycans (toluidine blue, TB) and sulfated glycosaminoglycans (GAGs) synthesis (safranin-O, SO), and immunolocalization of collagen-type II (Col-II) and aggrecan (scale bar = 200 µm), was performed in osteoarthritic bone marrow mesenchymal stromal cells (BM-MSCs) pellets, after 14 days, in chondrogenic medium supplemented with 100 µg/mL CS extracted from fish (<span class="html-italic">Prionace glauca</span>, <span class="html-italic">Raja clavata</span> and <span class="html-italic">Scyliorhinus canicula</span>) and bovine sources or CS-free. Immunopositive aggrecan (bottom, left) and collagen type-II (bottom, right) percentage area were normalized against cell-pellets total area (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 4
<p>mRNA relative expression of <span class="html-italic">SOX9, ACAN, COL2A1, COL1A1, COL10A1</span> and <span class="html-italic">IL6</span> from osteoarthritic BM-MSCs pellets, after 14 days, in chondrogenic medium supplemented with 100 µg/mL CS extracted from fish (<span class="html-italic">Prionace glauca, Raja clavata</span> and <span class="html-italic">Scyliorhinus canicula</span>) and bovine sources. Data were normalized against chondrogenic medium in the absence of CS (CS-free) condition (positive control for chondrogenesis), represented as <span class="html-italic">n</span> = 1. Values are given as the mean of 4 donors with standard deviation. No significant differences were found.</p>
Full article ">
16 pages, 7952 KiB  
Article
Pre-Treatment with Laminarin Protects Hippocampal CA1 Pyramidal Neurons and Attenuates Reactive Gliosis Following Transient Forebrain Ischemia in Gerbils
by Tae-Kyeong Lee, Ji Hyeon Ahn, Cheol Woo Park, Bora Kim, Young Eun Park, Jae-Chul Lee, Joon Ha Park, Go Eun Yang, Myoung Cheol Shin, Jun Hwi Cho, Il-Jun Kang and Moo-Ho Won
Mar. Drugs 2020, 18(1), 52; https://doi.org/10.3390/md18010052 - 12 Jan 2020
Cited by 23 | Viewed by 3651
Abstract
Transient brain ischemia triggers selective neuronal death/loss, especially in vulnerable regions of the brain including the hippocampus. Laminarin, a polysaccharide originating from brown seaweed, has various pharmaceutical properties including an antioxidant function. To the best of our knowledge, few studies have been conducted [...] Read more.
Transient brain ischemia triggers selective neuronal death/loss, especially in vulnerable regions of the brain including the hippocampus. Laminarin, a polysaccharide originating from brown seaweed, has various pharmaceutical properties including an antioxidant function. To the best of our knowledge, few studies have been conducted on the protective effects of laminarin against ischemic injury induced by ischemic insults. In this study, we histopathologically investigated the neuroprotective effects of laminarin in the Cornu Ammonis 1 (CA1) field of the hippocampus, which is very vulnerable to ischemia-reperfusion injury, following transient forebrain ischemia (TFI) for five minutes in gerbils. The neuroprotective effect was examined by cresyl violet staining, Fluoro-Jade B histofluorescence staining and immunohistochemistry for neuronal-specific nuclear protein. Additionally, to study gliosis (glial changes), we performed immunohistochemistry for glial fibrillary acidic protein to examine astrocytes, and ionized calcium-binding adaptor molecule 1 to examine microglia. Furthermore, we examined alterations in pro-inflammatory M1 microglia by using double immunofluorescence. Pretreatment with 10 mg/kg laminarin failed to protect neurons in the hippocampal CA1 field and did not attenuate reactive gliosis in the field following TFI. In contrast, pretreatment with 50 or 100 mg/kg laminarin protected neurons, attenuated reactive gliosis and reduced pro-inflammatory M1 microglia in the CA1 field following TFI. Based on these results, we firmly propose that 50 mg/kg laminarin can be strategically applied to develop a preventative against injuries following cerebral ischemic insults. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Cresyl Violet (CV) staining in the hippocampus (<b>A</b>–<b>H</b>) and its Cornu Ammonis 1 (CA1) field (<b>a</b>–<b>h</b>) of the vehicle/sham (<b>A</b>,<b>a</b>), 10, 50 and 100 mg/kg laminarin (LA)/sham (<b>C</b>,<b>c</b>,<b>E</b>,<b>e</b>,<b>G</b>,<b>g</b>), vehicle/ischemia (<b>B</b>,<b>b</b>) and 10, 50 and 100 mg/kg LA/ischemia (<b>D</b>,<b>d</b>,<b>F</b>,<b>f</b>,<b>H</b>,<b>h</b>) groups at 5 days after sham or transient forebrain ischemia (TFI) operation. In the vehicle/ischemia group, CV dyeability is remarkably reduced in the stratum pyramidale (SP, arrows) of the CA1 field (asterisks). In the 10 mg/kg LA/ischemia group, the distribution pattern of CV stained cells is similar to that in the vehicle/ischemia group. However, in the 50 mg/kg and 100 mg/kg LA/ischemia groups, CV stainability is conserved. DG, dentate gyrus; SO, stratum oriens; SR stratum radiatum. Scale bars = 400 μm (<b>A</b>–<b>H</b>) and 100 μm (<b>a</b>–<b>h</b>).</p>
Full article ">Figure 2
<p>NeuN immunohistochemistry in the CA1 field of the vehicle/sham (<b>A</b>), 10, 50 and 100 mg/kg LA/sham (<b>C</b>,<b>E</b>,<b>G</b>), vehicle/ischemia (<b>B</b>) and 10, 50 and 100 mg/kg LA/ischemia (<b>D</b>,<b>F</b>,<b>H</b>) groups at 5 days after sham or TFI operation. Numerous NeuN immunoreactive CA1 pyramidal neurons can be observed in the vehicle/sham group. In the vehicle/ischemia and 10 mg/kg LA/ischemia groups, significant decreases in NeuN immunoreactive CA1 pyramidal neurons were detected. In the 50 mg/kg and 100 mg/kg LA/ischemia groups, CA1 pyramidal neurons show strong NeuN immunoreactivity. Scale bar = 100 μm. (<b>I</b>) Mean number of NeuN immunoreactive pyramidal cells in the CA1 field at 5 days after TFI (<span class="html-italic">n</span> = 7 in each group, * <span class="html-italic">p</span> &lt; 0.05 versus vehicle/sham group, † <span class="html-italic">p</span> &lt; 0.05 versus vehicle/ischemia group). The bars indicate the means ± SEM.</p>
Full article ">Figure 3
<p>F-J B histofluorescence staining in the CA1 field of the vehicle/sham (<b>A</b>), 10, 50 and 100 mg/kg LA/sham (<b>C</b>,<b>E</b>,<b>G</b>), vehicle-ischemia (<b>B</b>) and 10, 50 and 100 mg/kg LA/ischemia (<b>D</b>,<b>F</b>,<b>H</b>) groups at 5 days after sham or TFI operation. In all the sham groups, no F-J B positive cells are found in the CA1 field; numerous F-J B positive cells are shown in the SP (asterisks) in the vehicle/ and 10 mg/kg LA/ischemia groups. However, in the 50 mg/kg and 100 mg/kg LA/ischemia groups, F-J B positive cells (arrows) are significantly decreased. Scale bar = 100 μm. (<b>I</b>) Mean number of F-J B positive pyramidal cells in the CA1 field at 5 days after TFI (<span class="html-italic">n</span> = 7 in each group, * <span class="html-italic">p</span> &lt; 0.05 versus vehicle/sham group, † <span class="html-italic">p</span> &lt; 0.05 versus vehicle/ischemia group). The bars indicate the means ± SEM.</p>
Full article ">Figure 4
<p>Glial fibrillary acidic protein (GFAP) immunohistochemistry in the CA1 field of the vehicle/sham (<b>A</b>), 10, 50 and 100 mg/kg LA/sham (<b>C</b>,<b>E</b>,<b>G</b>), vehicle/ischemia (<b>B</b>) and 10, 50 and 100 mg/kg LA/ischemia (<b>D</b>,<b>F</b>,<b>H</b>) groups at 5 days after sham or TFI operation. In all the sham groups, typical GFAP immunoreactive astrocytes are generally distributed in the stratum oriens (SO) and radiatum (SR). In the vehicle/ischemia group, GFAP immunoreactive astrocytes are hypertrophied. In the 10 mg/kg LA/ischemia group, GFAP immunoreactive astrocytes are similar to those in the vehicle/ischemia group. In the 50 mg/kg and 100 mg/kg LA/ischemia groups, hypertrophy of GFAP immunoreactive astrocytes is apparently attenuated. Scale bar = 100 μm. (<b>I</b>) ROD (percentage) of GFAP immunoreactive structures in the CA1 field at 5 days after TFI (<span class="html-italic">n</span> = 7 in each group, * <span class="html-italic">p</span> &lt; 0.05 versus vehicle/sham group, † <span class="html-italic">p</span> &lt; 0.05 versus vehicle/ischemia group). The bars indicate the means ± SEM.</p>
Full article ">Figure 5
<p>Ionized calcium-binding adapter molecule 1 (Iba-1) immunohistochemistry in the CA1 field of the vehicle/sham (<b>A</b>), 10, 50 and 100 mg/kg LA/sham (<b>C</b>,<b>E</b>,<b>G</b>), vehicle/ischemia (<b>B</b>) and 10, 50 and 100 mg/kg LA/ischemia (<b>D</b>,<b>F</b>,<b>H</b>) groups at 5 days after sham or TFI operation. Iba-1 immunoreactive microglia are in a resting state in all the sham groups. In the vehicle/ischemia and 10 mg/kg LA/ischemia groups, Iba-1 immunoreactive microglia are hypertrophied, showing that many activated microglia gather in the SP (arrows). In the 50 mg/kg and 100 mg/kg LA/ischemia groups, activation of Iba-1 immunoreactive microglia is markedly attenuated, showing that they are evenly distributed in the CA1 field. Scale bar = 100 μm. (<b>I</b>) ROD (percentage) of Iba-1 immunoreactive structures in the CA1 field at 5 days after TFI (<span class="html-italic">n</span> = 7 in each group, * <span class="html-italic">p</span> &lt; 0.05 versus vehicle/sham group, † <span class="html-italic">p</span> &lt; 0.05 versus vehicle/ischemia group). The bars indicate the means ± SEM.</p>
Full article ">Figure 6
<p>Double immunofluorescence staining for Iba-1 (red), interleukin 2 (IL-2) (green) and merged images in the hippocampal CA1 field of the vehicle/ischemia (<b>A</b>–<b>C</b>) and 50 mg/kg LA/ischemia (<b>D</b>–<b>F</b>) groups at 5 days after TFI. Many IL-2 immunoreactive microglia (arrows) are shown in the vehicle/ischemia group. However, in the 50 mg/kg LA/ischemia group, a few IL-2 immunoreactive microglia are detected. Scale bar = 40 μm (<span class="html-italic">n</span> = 7 in each group).</p>
Full article ">

2019

Jump to: 2024, 2022, 2021, 2020, 2018, 2017, 2016, 2015, 2014, 2013, 2012, 2011, 2010

15 pages, 4412 KiB  
Article
Fucoidan from Undaria pinnatifida Ameliorates Epidermal Barrier Disruption via Keratinocyte Differentiation and CaSR Level Regulation
by Yu Chen, Xuenan Li, Xiaoshuang Gan, Junmei Qi, Biao Che, Meiling Tai, Shuang Gao, Wengang Zhao, Nuo Xu and Zhenlin Hu
Mar. Drugs 2019, 17(12), 660; https://doi.org/10.3390/md17120660 - 24 Nov 2019
Cited by 10 | Viewed by 3942
Abstract
The epidermal barrier acts as a line of defense against external agents as well as helps to maintain body homeostasis. The calcium concentration gradient across the epidermal barrier is closely related to the proliferation and differentiation of keratinocytes (KCs), and the regulation of [...] Read more.
The epidermal barrier acts as a line of defense against external agents as well as helps to maintain body homeostasis. The calcium concentration gradient across the epidermal barrier is closely related to the proliferation and differentiation of keratinocytes (KCs), and the regulation of these two processes is the key to the repair of epidermal barrier disruption. In the present study, we found that fucoidan from Undaria pinnatifida (UPF) could promote the repair of epidermal barrier disruption in mice. The mechanistic study demonstrated that UPF could promote HaCaT cell differentiation under low calcium condition by up-regulating the expression of calcium-sensing receptor (CaSR), which could then lead to the activation of the Catenin/PLCγ1 pathway. Further, UPF could increase the expression of CaSR through activate the ERK and p38 pathway. These findings reveal the molecular mechanism of UPF in the repair of the epidermal barrier and provide a basis for the development of UPF into an agent for the repair of epidermal barrier repair. Full article
Show Figures

Figure 1

Figure 1
<p>Molecular weight and monosaccharide composition of UPF. (<b>A</b>) The molecular weight and molecular mass distributions of UPF were determined by GPC-MALLS consisting of a refractive index detector Waters 2414 (RI) and a Wyatt DAWN EOS MALLS detector (<b>B</b>) The UPF were dissolved in ammonia, mixed with PMP, and neutralized with 200 µL of formic acid. The derivatization chromatomap was collected by UPLC/Q-TOF-MS. (<b>C</b>) Derivatization chromatomap of standard monosaccharides (Man, Rib, Rha, GluUA, GalUA, Glc, Gal, Xyl, Ara, Fuc).</p>
Full article ">Figure 2
<p>The effect of UPF on the recovery of epidermal barrier. Epidermal barrier disruption of ICR mice was induced by tape stripping on their shaved back skin until the TEWL reached 40 mg/cm<sup>2</sup>/hour. UPF hydrogel was administrated topically on the dorsal skin. (<b>A</b>) The photos were taken every 12 h after disruption. (<b>B</b>) The values of TEWL were measured at 0 h, 12 h, 24 h, 48 h, 72 h, 84 h. Data are presented as means ± SEM obtained from three independent experiments, * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 versus the vehicle control. (<b>C</b>) The back skin were harvested at 84 h and the skin sections were prepared and stained with hematoxylin–eosin. (<b>D</b>) The back skin were harvested at 84 h and the skin sections were prepared and stained with immunohistochemistry.</p>
Full article ">Figure 3
<p>The effect of UPF on proliferation, differentiation and sustained Ca<sup>2+</sup> concentration in HaCaT cells. HaCaT cells were treated with or without indicated concentrations (10, 20, 50 μg/mL of UPF. (<b>A</b>) The involucrin and filaggrin mRNA levels in HaCaT cells were assayed after 72 h treatment by qPCR. (<b>B</b>) The involucrin and filaggrin proteins levels in HaCaT cells were assayed after 72 h treatment by Western blot. (<b>C</b>) The involucrin and filaggrin mRNA levels in NHEK cells were assayed after 72 h treatment by qPCR. (<b>D</b>) The involucrin and filaggrin proteins levels in NHEK cells were assayed after 72 h treatment by Western blot. (<b>E</b>) The proliferation of HaCaT cells were assayed after 24 h, 48 h or 72 h treatment by BrdU assay. (<b>F</b>) HaCaT cells were treated with or without indicated concentrations (10, 20, 50 µg/mL) of UPF for 24 h, loaded with Fura-4 AM, and the fluorescence was recorded using a confocal microscope. Data are presented as means ± SEM obtained from three independent experiments, * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 versus Control.</p>
Full article ">Figure 4
<p>The effect of UPF on the expression of CaSR and CaSR mediated signaling pathway. (<b>A</b>) HaCaT cells were treated with or without indicated concentrations (10, 20, 50 µg/mL) of UPF, the CaSR mRNA levels in HaCaT cells were assayed at 12 h, 24 h, and 48 h by qPCR. (<b>B</b>) HaCaT cells were treated with or without indicated concentrations (10, 20, 50 µg/mL) of UPF, the CaSR protein levels in HaCaT cells were assayed at 12 h, 24 h, and 48 h by Western blot. (<b>C</b>) CaSR mediated signaling pathway. (<b>D</b>) HaCaT cells were treated with or without indicated concentrations (10, 20, 50 µg/mL) of UPF, the levels of PLC-γ1, p120-catenin, phospho-PLCγ1 and phospho- p120-catenin in cytoplasm and β-catenin in the nucleus were detected by Western blot. (<b>E</b>) HaCaT cells were exposed to indicate concentrations of UPF with or without NPS-2143 (CaSR inhibitor, 150 nM), the protein levels of PLC-γ1, p120-catenin, phospho-PLCγ1, phospho-p120-catenin, the involucrin and filaggrin were assayed by Western blot. (<b>F</b>) HaCaT cells were exposed to indicate concentrations of UPF with or without NPS-2143 (CaSR inhibitor) for 72 h, the involucrin and filaggrin mRNA levels were assayed by qPCR.</p>
Full article ">Figure 5
<p>UPF increases the expression of CaSR via the ERK and p38 signaling pathway. (<b>A</b>) HaCaT cells were treated with or without indicated concentrations (10, 20, 50 μg/mL) of UPF, the levels of ERK, p38, phospho-ERK and phospho-p38 in cells were detected by Western blotting. (<b>B</b>) HaCaT cells were exposed to indicated concentrations of UPF with or without LY3214996 (ERK inhibitor, 100 nM) and SB203580 (p38 inhibitor, 20 µM) for 12 h, the levels of PLC-γ1, p120-catenin, phospho-PLCγ1 and phospho-p120-catenin in cells were detected by Western blotting. (<b>C</b>) The relative expression levels of phosphorylation of PLCγ1 and p120-catenin were detected by gray analysis. (<b>D</b>) HaCaT cells were exposed to indicated concentrations of UPF with or without LY3214996 and SB203580 for 72 h, the CaSR mRNA levels in HaCaT cells were assayed by qPCR. (<b>E</b>) HaCaT cells were exposed to indicated concentrations of UPF with or without LY3214996 and SB203580 for 72 h, the involucrin and filaggrin mRNA levels in HaCaT cells were assayed by qPCR.</p>
Full article ">
12 pages, 1899 KiB  
Article
Effect of Carboxymethylation and Phosphorylation on the Properties of Polysaccharides from Sepia esculenta Ink: Antioxidation and Anticoagulation in Vitro
by Huazhong Liu, Fangping Li and Ping Luo
Mar. Drugs 2019, 17(11), 626; https://doi.org/10.3390/md17110626 - 1 Nov 2019
Cited by 28 | Viewed by 2964
Abstract
To investigate the effect of carboxymethylation and phosphorylation modification on Sepia esculenta ink polysaccharide (SIP) properties, this study prepared carboxymethyl SIP (CSIP) with the chloracetic acid method, and phosphorylated SIP (PSIP) with the sodium trimetaphosphate (STMP)/sodium tripolyphosphate (STPP) method, on the basis of [...] Read more.
To investigate the effect of carboxymethylation and phosphorylation modification on Sepia esculenta ink polysaccharide (SIP) properties, this study prepared carboxymethyl SIP (CSIP) with the chloracetic acid method, and phosphorylated SIP (PSIP) with the sodium trimetaphosphate (STMP)/sodium tripolyphosphate (STPP) method, on the basis of an orthogonal experiment. The in vitro antioxidant and anticoagulant activities of the derivatives were determined by assessing the scavenging capacity of the 1,1-diphenyl-2-picrylhydrazyl (DPPH) and hydroxyl radicals, which activated the partial thromboplastin time (APTT), prothrombin time (PT), and thrombin time (TT). The results showed that SIP was modified successfully to be CSIP and PSIP, and degrees of substitution (DSs) of the two products were 0.9913 and 0.0828, respectively. Phosphorylation efficiently improved the antioxidant property of SIP, and the IC50 values of PSIP on DPPH and hydroxyl radicals decreased by 63.25% and 13.77%, respectively. But carboxymethylation reduced antioxidant activity of the native polysaccharide, IC50 values of CSIP on the DPPH and hydroxyl radicals increased by 16.74% and 6.89%, respectively. SIP significantly prolonged the APTT, PT, and TT in a dose-dependent fashion, suggesting that SIP played an anticoagulant action through intrinsic, extrinsic, and common coagulation pathways. CSIP and PSIP both possessed a stronger anticoagulant capacity than SIP via the same pathways; moreover, CSIP was observed to be more effective in prolonging APTT and PT than PSIP. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Effect of different reaction conditions on degree of substitution (DS) of carboxymethylated <span class="html-italic">Sepia esculenta</span> ink polysaccharides (CSIP). Under the scheduled different reaction conditions, the SIP was carboxymethylated by chloroacetic acid. The degree of substitution of CSIP was determined by the neutralization titration method and calculation according to the formula.</p>
Full article ">Figure 2
<p>IR spectra of SIP and CSIP. IR spectra of the SIP and CSIP were recorded with KBr pellets on a Bruker Tensor 27 Fourier infrared spectrophotometer between 400–4000 cm<sup>−1</sup>.</p>
Full article ">Figure 3
<p>Effect of different reaction conditions on the content of phosphate in phosphorylated SIP (PSIP). Under the scheduled different reaction conditions, the SIP was phosphorylated with the STMP/STPP method. The content of phosphate in PSIP was determined according to the equation, which was subjected to calculating the degree of substitution of PSIP.</p>
Full article ">Figure 4
<p>IR spectra of SIP and PSIP. The IR spectra of SIP and PSIP were recorded with KBr pellets on a BRUKER TENSOR 27 Fourier infrared spectrophotometer between 400–4000 cm<sup>−1</sup>.</p>
Full article ">Figure 5
<p>In vitro antioxidant activity of SIP, CSIP, and PSIP. The antioxidant capacity of SIP and its derivatives was assessed by the scavenging activities of DPPH and hydroxyl radicals.</p>
Full article ">Figure 6
<p>In vitro anticoagulant property of SIP, CSIP, and PSIP. The anticoagulant capacity of SIP and its derivatives was assessed by determining the clotting time, including the activated partial thromboplastin time (APTT), prothrombin time (PT), and thrombin time (TT). Different capital or lowercase letters express significant differences among various dosages (0, 6.25, 12.5, 25, 50, and 100) of polysaccharide (SIP, CSIP, or PSIP), <sup>ABCDE</sup> <span class="html-italic">p</span> &lt; 0.01 or <sup>abcdef</sup> <span class="html-italic">p</span> &lt; 0.05. Asterisk, * or **, means <span class="html-italic">p</span> &lt; 0.05 or <span class="html-italic">p</span> &lt; 0.01 between the same dosage of SIP and modified SIP (CSIP or PSIP), respectively. <sup>#</sup> or <sup>##</sup> represent the difference, <span class="html-italic">p</span> &lt; 0.05 or <span class="html-italic">p</span> &lt; 0.01, between the same dosage of CSIP and PSIP, respectively.</p>
Full article ">Figure 6 Cont.
<p>In vitro anticoagulant property of SIP, CSIP, and PSIP. The anticoagulant capacity of SIP and its derivatives was assessed by determining the clotting time, including the activated partial thromboplastin time (APTT), prothrombin time (PT), and thrombin time (TT). Different capital or lowercase letters express significant differences among various dosages (0, 6.25, 12.5, 25, 50, and 100) of polysaccharide (SIP, CSIP, or PSIP), <sup>ABCDE</sup> <span class="html-italic">p</span> &lt; 0.01 or <sup>abcdef</sup> <span class="html-italic">p</span> &lt; 0.05. Asterisk, * or **, means <span class="html-italic">p</span> &lt; 0.05 or <span class="html-italic">p</span> &lt; 0.01 between the same dosage of SIP and modified SIP (CSIP or PSIP), respectively. <sup>#</sup> or <sup>##</sup> represent the difference, <span class="html-italic">p</span> &lt; 0.05 or <span class="html-italic">p</span> &lt; 0.01, between the same dosage of CSIP and PSIP, respectively.</p>
Full article ">
14 pages, 3132 KiB  
Article
Anti-Inflammatory and Anti-Aging Evaluation of Pigment–Protein Complex Extracted from Chlorella Pyrenoidosa
by Ruilin Zhang, Jian Chen, Xinwu Mao, Ping Qi and Xuewu Zhang
Mar. Drugs 2019, 17(10), 586; https://doi.org/10.3390/md17100586 - 16 Oct 2019
Cited by 23 | Viewed by 5472
Abstract
Oxidative stress contributes to chronic inflammatory processes implicated in aging, referred to as “inflamm-aging.” In this study, the potential anti-inflammatory and anti-aging effects of a pigment–protein complex (PPC) from Chlorella pyrenoidosa were investigated using lipopolysaccharide (LPS)-stimulated RAW 264.7 macrophages and D-galactose (D-gal)-induced aging [...] Read more.
Oxidative stress contributes to chronic inflammatory processes implicated in aging, referred to as “inflamm-aging.” In this study, the potential anti-inflammatory and anti-aging effects of a pigment–protein complex (PPC) from Chlorella pyrenoidosa were investigated using lipopolysaccharide (LPS)-stimulated RAW 264.7 macrophages and D-galactose (D-gal)-induced aging in a murine model. Results indicated that PPC inhibits the production of the inflammatory cytokines TNF-α and IL-6, and the inflammatory mediator nitric oxide (NO) in LPS-stimulated RAW 264.7 cells. It also protected mice from D-gal induced informatory aging by increasing the activity of the antioxidant enzyme, such as superoxide dismutase (SOD), inhibiting D-gal-induced NF-κB upregulation, and increasing PPARs expression in the brain and gut. The findings indicated that PPC has favorable anti-inflammatory and anti-aging properties, and could be useful in the treatment of acute inflammation and senescence diseases. Full article
Show Figures

Figure 1

Figure 1
<p>Purification and characterization of the pigment–protein mixture from <span class="html-italic">Chlorella pyrenoidosa</span>. (<b>a</b>) Sephadex G-25 gel filtration chromatography; (<b>b</b>) Absorption spectrum of the pigment–protein complex (PPC); (<b>c</b>) HPLC analysis.</p>
Full article ">Figure 2
<p>Effect of PPC on RAW 264.7 cells. (<b>a</b>) The viability was measured by 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide (MTT) assay. The control group consisted of untreated cells and was considered as 100% of viable cells. Results were expressed as a percentage of viable cells when compared with the control group. (<b>b</b>) Nitric oxide (NO) production levels; (<b>c</b>) Expression levels of tumor necrosis factor (TNF)-α; (<b>d</b>) The levels of interleukin (IL)-6 production; (<b>e</b>) The levels of IL-10 production; (<b>f</b>) Phagocytosis rate. All the experiments were performed in triplicate. Duncan’s new multiple range test was performed to determine the significance differences. The values are expressed as the as mean ± SD. <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05 vs. the blank group, <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 vs. the blank group, * <span class="html-italic">p</span> &lt; 0.05 vs. the LPS group, and ** <span class="html-italic">p</span> &lt; 0.05 vs. the lipopolysaccharide (LPS) group.</p>
Full article ">Figure 3
<p>D-galactose (D-gal) induced KMB17 cells premature senescence in vitro. (<b>A</b>) Cells morphology changes: normal group (<b>a</b>), Model group D-gal-stimulated (<b>b</b>), Ascorbic acid group (<b>c</b>), 200 μg/mL PPC (<b>d</b>), 400 μg/mL PPC (<b>e</b>); (<b>B</b>) The effects of PPC on KMB17 cell viability. The experiments were performed in triplicate, and the values are expressed as the as mean ± SD. Results were expressed as a percentage of viable cells when compared with the control group.</p>
Full article ">Figure 4
<p>Effect of PPC on superoxide dismutase (SOD), malondialdehyde (MDA), NO, IL-6, TNF-α levels, and the morphological features of gut in D-gal-induced mice. (<b>a</b>) Changes in MDA, (<b>b</b>) SOD, and (<b>c</b>) NO production levels; (<b>d</b>) The levels of IL-6 production; (<b>e</b>) Expression levels of TNF-α; (<b>f</b>) The morphological features of hematoxylin and eosin (H&amp;E) stained gut sections, Scale bar = 50 μm. A one-way ANOVA was performed to compare the three experimental groups. All the data are mean ± SD of three independent experiments. <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05 vs. the control group, <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 vs. the control group, * <span class="html-italic">p</span> &lt; 0.05 vs. the model group, and ** <span class="html-italic">p</span> &lt; 0.01 vs. the model group.</p>
Full article ">Figure 4 Cont.
<p>Effect of PPC on superoxide dismutase (SOD), malondialdehyde (MDA), NO, IL-6, TNF-α levels, and the morphological features of gut in D-gal-induced mice. (<b>a</b>) Changes in MDA, (<b>b</b>) SOD, and (<b>c</b>) NO production levels; (<b>d</b>) The levels of IL-6 production; (<b>e</b>) Expression levels of TNF-α; (<b>f</b>) The morphological features of hematoxylin and eosin (H&amp;E) stained gut sections, Scale bar = 50 μm. A one-way ANOVA was performed to compare the three experimental groups. All the data are mean ± SD of three independent experiments. <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05 vs. the control group, <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 vs. the control group, * <span class="html-italic">p</span> &lt; 0.05 vs. the model group, and ** <span class="html-italic">p</span> &lt; 0.01 vs. the model group.</p>
Full article ">Figure 5
<p>(<b>a</b>) Effects of PPC on the expression levels of nuclear factor κB (NF-κB), PPARα, PPARγ, and p53 in mice, the proteins of brain and gut in each group were processed by Western blotting, and all data performed in triplicate; (<b>b</b>) Potential regulation pathways by Search Tool for the Retrieval of Interacting Genes/Proteins (STRING) software (<a href="http://string.embl.de/);" target="_blank">http://string.embl.de/);</a> (<b>c</b>) Proposed mechanism of action on brain and intestine.</p>
Full article ">
17 pages, 3723 KiB  
Article
Efficiently Anti-Obesity Effects of Unsaturated Alginate Oligosaccharides (UAOS) in High-Fat Diet (HFD)-Fed Mice
by Shangyong Li, Ningning He and Linna Wang
Mar. Drugs 2019, 17(9), 540; https://doi.org/10.3390/md17090540 - 17 Sep 2019
Cited by 59 | Viewed by 4680
Abstract
Obesity and its related complications have become one of the leading problems affecting human health. However, current anti-obesity treatments are limited by high cost and numerous adverse effects. In this study, we investigated the use of a non-toxic green food additive, known as [...] Read more.
Obesity and its related complications have become one of the leading problems affecting human health. However, current anti-obesity treatments are limited by high cost and numerous adverse effects. In this study, we investigated the use of a non-toxic green food additive, known as unsaturated alginate oligosaccharides (UAOS) from the enzymatic degradation of Laminaria japonicais, which showed effective anti-obesity effects in a high-fat diet (HFD) mouse model. Compared with acid hydrolyzed saturated alginate oligosaccharides (SAOS), UAOS significantly reduced body weight, serum lipid, including triacylglycerol (TG), total cholesterol (TC) and free fatty acids (FFA), liver weight, liver TG and TC, serum alanine aminotransferase (ALT), and aspartate aminotransferase (AST) levels, adipose mass, reactive oxygen species (ROS) formation, and accumulation induced in HFD mice. Moreover, the structural differences in β-d-mannuronate (M) and its C5 epimer α-l-guluronate (G) did not cause significant functional differences. Meanwhile, UAOS significantly increased both AMP-activated protein kinase α (AMPKα) and acetyl-CoA carboxylase (ACC) phosphorylation in adipocytes, which indicated that UAOS had an anti-obesity effect mainly through AMPK signaling. Our results indicate that UAOS has the potential for further development as an adjuvant treatment for many metabolic diseases such as fatty liver, hypertriglyceridemia, and possibly diabetes. Full article
Show Figures

Figure 1

Figure 1
<p>Scheme (<b>A</b>), TLC analysis (<b>B</b>), SE-HPLC analysis (<b>C</b>,<b>D</b>), and ESI-MS analysis (<b>E</b>,<b>F</b>) of UAOS and SAOS.</p>
Full article ">Figure 2
<p>Changes in body weight on AOS treatment. Changes in body weight of treatment of UAOS (<b>A</b>) and SAOS (<b>B</b>), body weight gain (<b>C</b>) for 4 weeks. The data are represented as means ± standard deviation (SD, <span class="html-italic">n</span> = 6). Changes in the energy intake (<b>D</b>), body weight (<b>E</b>), body weight gain (<b>F</b>), and Lee’s index (<b>G</b>) during the eight-week treatment are shown. The body weight was recorded weekly. The data are represented as means ± standard deviation (SD, <span class="html-italic">n</span> = 12). Compare with HFD group, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001. Compare with indicated groups, <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 3
<p>Changes of AOS treatment on serum lipid. TG levels (<b>A</b>), TC levels (<b>B</b>), HDL-c levels (<b>C</b>), LDL-c levels (<b>D</b>), and FFA levels (<b>E</b>). The data are represented as means ± standard deviation (SD, <span class="html-italic">n</span> = 12). Compare with the HFD group, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001. Compare with indicated groups, <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05, <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 4
<p>Changes of AOS treatment on the hepatic lipid. The whole liver weight (<b>A</b>), liver TC (<b>B</b>), and TG (<b>C</b>) are displayed. The whole liver morphology (<b>D</b>) and H&amp;E staining for liver tissue (<b>E</b>) (200×) are shown. The concentration of serum AST (<b>F</b>) and ALT (<b>G</b>) are also shown. The data are represented as means ± standard deviation (SD, <span class="html-italic">n</span> = 12). Compare with the HFD group, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001. Compare with indicated groups, <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 5
<p>Changes of UAOS treatment on adipocytes in adipose mass. The fat mass of epididymal (<b>A</b>), mesenteric (<b>B</b>), perirenal tissues (<b>C</b>), adipocyte size (<b>D</b>), the whole WAT (<b>E</b>), and WAT tissue stained by H&amp;E staining (200×) (<b>F</b>) are shown in the figure. The data are represented as means ± standard deviation (SD, <span class="html-italic">n</span> = 12). Compare with the HFD group, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001. Compare with indicated groups, <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05, <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 6
<p>The effect of AOS treatment on hepatic H<sub>2</sub>O<sub>2</sub> level (<b>A</b>) and liver MDA level (<b>B</b>) in mice fed a high fat diet for 8 weeks. The data are represented as means ± standard deviation (SD, <span class="html-italic">n</span> = 12). Compare with the HFD group, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001. Compare with indicated groups, <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05, <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 7
<p>UAOS activated the AMPK signaling pathway in WAT. (<b>A</b>) The protein expression level of AMPK, p-AMPK, ACC, and p-ACC in WAT were detected by Western blotting. (<b>B</b>) The relative protein levels were quantified. (<b>C</b>) The proposed mechanism by which AOS causes effects of hepatoprotection and anti-obesity. The data are represented as means ± standard deviation (SD, <span class="html-italic">n</span> = 3). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">
18 pages, 1998 KiB  
Article
In-Depth Characterization of Bioactive Extracts from Posidonia oceanica Waste Biomass
by Isaac Benito-González, Amparo López-Rubio, Antonio Martínez-Abad, Ana-Rosa Ballester, Irene Falcó, Luis González-Candelas, Gloria Sánchez, Jesús Lozano-Sánchez, Isabel Borrás-Linares, Antonio Segura-Carretero and Marta Martínez-Sanz
Mar. Drugs 2019, 17(7), 409; https://doi.org/10.3390/md17070409 - 9 Jul 2019
Cited by 23 | Viewed by 5519
Abstract
Posidonia oceanica waste biomass has been valorised to produce extracts by means of different methodologies and their bioactive properties have been evaluated. Water-based extracts were produced using ultrasound-assisted and hot water methods and classified according to their ethanol-affinity (E1: ethanol soluble; E2: non-soluble). [...] Read more.
Posidonia oceanica waste biomass has been valorised to produce extracts by means of different methodologies and their bioactive properties have been evaluated. Water-based extracts were produced using ultrasound-assisted and hot water methods and classified according to their ethanol-affinity (E1: ethanol soluble; E2: non-soluble). Moreover, a conventional protocol with organic solvents was applied, yielding E3 extracts. Compositional and structural characterization confirmed that while E1 and E3 extracts were mainly composed of minerals and lipids, respectively, E2 extracts were a mixture of minerals, proteins and carbohydrates. All the extracts showed remarkably high antioxidant capacity, which was not only related to phenolic compounds but also to the presence of proteins and polysaccharides. All E2 and E3 extracts inhibited the growth of several foodborne fungi, while only E3 extracts decreased substantially the infectivity of feline calicivirus and murine norovirus. These results show the potential of P. oceanica waste biomass for the production of bioactive extracts. Full article
Show Figures

Figure 1

Figure 1
<p>Carbohydrate composition of the aqueous <span class="html-italic">P. oceanica</span> extracts. (<b>A</b>) Absolute concentration values for E1 and E2 extracts and (<b>B</b>) their relative abundance in the E2 extracts. Different letters denote significant monosaccharide content differences between extracts (<span class="html-italic">p</span> ≤ 0.05).</p>
Full article ">Figure 2
<p>FT-IR spectra of the different <span class="html-italic">P. oceanica</span> extracts obtained by means of water-based extractions (<b>A</b>,<b>B</b>) and organic solvent-based extractions (<b>C</b>). E1 US, E2 US and E3 US spectra have been offset for clarity.</p>
Full article ">Figure 3
<p>XRD patterns of the freeze-dried extracts. (<b>A</b>,<b>B</b>) Water-soluble extracts and (<b>C</b>) organic-soluble extracts. The spectra from E1 US and E2 US have been offset for clarity.</p>
Full article ">Figure 4
<p>Reduction of (<b>A</b>) feline calicivirus (FCV) titres (log TCID<sub>50</sub>/mL) and (<b>B</b>) murine norovirus (MNV) titres (log TCID<sub>50</sub>/mL) treated with <span class="html-italic">P. oceanica</span> extracts at different concentrations (0.5 or 0.05%) after 25 °C and 37 °C ON incubations, respectively. * Each bar represents the average of triplicates. Within each column, different letters denote significant differences between treatments. ** Horizontal line depicts the detection limit.</p>
Full article ">
12 pages, 734 KiB  
Communication
Different Antifungal Activity of Anabaena sp., Ecklonia sp., and Jania sp. against Botrytis cinerea
by Hillary Righini, Elena Baraldi, Yolanda García Fernández, Antera Martel Quintana and Roberta Roberti
Mar. Drugs 2019, 17(5), 299; https://doi.org/10.3390/md17050299 - 20 May 2019
Cited by 32 | Viewed by 4638
Abstract
Water extracts and polysaccharides from Anabaena sp., Ecklonia sp., and Jania sp. were tested for their activity against the fungal plant pathogen Botrytis cinerea. Water extracts at 2.5, 5.0, and 10.0 mg/mL inhibited B. cinerea growth in vitro. Antifungal activity of polysaccharides [...] Read more.
Water extracts and polysaccharides from Anabaena sp., Ecklonia sp., and Jania sp. were tested for their activity against the fungal plant pathogen Botrytis cinerea. Water extracts at 2.5, 5.0, and 10.0 mg/mL inhibited B. cinerea growth in vitro. Antifungal activity of polysaccharides obtained by N-cetylpyridinium bromide precipitation in water extracts was evaluated in vitro and in vitro at 0.5, 2.0, and 3.5 mg/mL. These concentrations were tested against fungal colony growth, spore germination, colony forming units (CFUs), CFU growth, and on strawberry fruits against B. cinerea infection with pre- and post-harvest application. In in vitro experiments, polysaccharides from Anabaena sp. and from Ecklonia sp. inhibited B. cinerea colony growth, CFUs, and CFU growth, while those extracted from Jania sp. reduced only the pathogen spore germination. In in vitro experiments, all concentrations of polysaccharides from Anabaena sp., Ecklonia sp., and Jania sp. reduced both the strawberry fruits infected area and the pathogen sporulation in the pre-harvest treatment, suggesting that they might be good candidates as preventive products in crop protection. Full article
Show Figures

Figure 1

Figure 1
<p>Effect of different POL concentrations on <span class="html-italic">Botrytis cinerea</span> colony growth rate. Treatment and dose factors and their interaction are significant, according to factorial ANOVA. F<sub>(2,36)</sub> = 18.2, <span class="html-italic">p</span> &lt; 0.0001 (for treatment factor), F<sub>(3,36)</sub> = 27.8, <span class="html-italic">p</span> &lt; 0.0001 (for dose factor), F<sub>(6,36)</sub> = 2.7, <span class="html-italic">p</span> &lt; 0.05 (for interaction). Columns are mean values ± SD. The same uppercase letter within each POL treatment and the same lowercase letter among concentrations indicates no significant differences according to Student–Newman–Keuls test (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>Infected area of strawberry fruit caused by <span class="html-italic">Botrytis cinerea</span> (<b>a</b>) and its sporulation (<b>b</b>) after pre-harvest treatment with different concentrations of polysaccharides from <span class="html-italic">Anabaena</span> sp. (AN), <span class="html-italic">Ecklonia</span> sp. (ECK), and <span class="html-italic">Jania</span> sp. (JAN). Polysaccharides and concentration factors and their interaction are significant, according to factorial ANOVA. (<b>a</b>) F<sub>(2,240)</sub> = 270.0, <span class="html-italic">p</span> &lt; 0.0001 (for treatment factor), F<sub>(3,240)</sub> = 266.3, <span class="html-italic">p</span> &lt; 0.0001 (for dose factor), F<sub>(6,240)</sub> = 45.0, <span class="html-italic">p</span> &lt; 0.05 (for interaction). (<b>b</b>) F<sub>(2,96)</sub> = 23.0, <span class="html-italic">p</span> &lt; 0.0001 (for treatment factor), F<sub>(3,96)</sub> = 370.0, <span class="html-italic">p</span> &lt; 0.0001 (for dose factor), F<sub>(6,96)</sub> = 18.5, <span class="html-italic">p</span> &lt; 0.05 (for interaction). Columns are mean values ± SD. The same uppercase letter within each POL treatment and the same lowercase letter within each concentration indicates no significant differences according to Student–Newman–Keuls test (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
18 pages, 2351 KiB  
Article
Broad-Spectrum Anti-Adhesive Coating Based on an Extracellular Polymer from a Marine Cyanobacterium
by Bruna Costa, Rita Mota, Paula Parreira, Paula Tamagnini, M. Cristina L. Martins and Fabíola Costa
Mar. Drugs 2019, 17(4), 243; https://doi.org/10.3390/md17040243 - 24 Apr 2019
Cited by 15 | Viewed by 4440
Abstract
Medical device-associated infections are a major health threat, representing about half of all hospital-acquired infections. Current strategies to prevent this problem based on device coatings with antimicrobial compounds (antibiotics or antiseptics) have proven to be insufficient, often toxic, and even promoting bacterial resistance. [...] Read more.
Medical device-associated infections are a major health threat, representing about half of all hospital-acquired infections. Current strategies to prevent this problem based on device coatings with antimicrobial compounds (antibiotics or antiseptics) have proven to be insufficient, often toxic, and even promoting bacterial resistance. Herein, we report the development of an infection-preventive coating (CyanoCoating) produced with an extracellular polymer released by the marine cyanobacterium Cyanothece sp. CCY 0110. CyanoCoating was prepared by spin-coating and its bacterial anti-adhesive efficiency was evaluated against relevant etiological agents (Staphylococcus aureus, S. epidermidis, Pseudomonas aeruginosa and Escherichia coli) and platelets, both in the presence or absence of human plasma proteins. CyanoCoating cytotoxicity was assessed using the L929 fibroblasts cell line. CyanoCoating exhibited a smooth topography, low thickness and high hydrophilic properties with mild negative charge. The non-cytotoxic CyanoCoating prevented adhesion of all the bacteria tested (≤80%) and platelets (<87%), without inducing platelet activation (even in the presence of plasma proteins). The significant reduction in protein adsorption (<77%) confirmed its anti-adhesive properties. The development of this anti-adhesive coating is an important step towards the establishment of a new technological platform capable of preventing medical device-associated infections, without inducing thrombus formation in blood-contacting applications. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Surface characterization of Au substrates coated with a polydopamine (pDA) layer, a pDA layer plus polyurethane (PU), and a pDA layer plus the CyanoCoating by (<b>A</b>) ellipsometry (n = 9) and (<b>B</b>) water contact angle (captive bubble method) (n = 9). Statistical analysis was performed by non-parametric Kruskal–Wallis analysis and statistic differences are indicated with *** (<span class="html-italic">p</span> &lt; 0.005) and **** (<span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 2
<p>Fourier transform–infrared reflection absorption spectroscopy (FT–IRRAS) spectra of Au substrates coated with a polydopamine layer (pDA), a pDA layer plus polyurethane (PU), and a pDA layer plus the CyanoCoating. Typical peaks of each spectrum are highlighted.</p>
Full article ">Figure 3
<p>Adsorption of bovine serum albumin (BSA) onto polyurethane (PU), or CyanoCoating (Voigt modulation). Statistical significance between surfaces is indicated by * (<span class="html-italic">p</span> &lt; 0.05, Unpaired t-test).</p>
Full article ">Figure 4
<p>CyanoCoating anti-adhesive performance compared to medical grade polyurethane (PU), in the absence and presence of human plasma proteins. The coatings were tested against the four relevant etiological agents mentioned above each graph. The assay was performed according to ISO 22196. Statistical analysis was performed by Non-parametric Kruskal-Wallis analysis and statistic differences are indicated with ** (<span class="html-italic">p</span> &lt; 0.01) and **** (<span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 5
<p>Scanning electron micrographs of polyurethane (PU) and CyanoCoating after 24 h incubation at 37 °C with four relevant etiological agents (scale bars—20 µm).</p>
Full article ">Figure 6
<p>Micrographs of <span class="html-italic">Staphylococcus epidermidis</span> cells adhered to polyurethane (PU) and CyanoCoating after 24 h incubation at 37 °C and stained with Draq5 and propidium iodide (PI). In the right panel, the live bacteria are marked in green and the dead bacteria are marked in red. The percentage of dead bacterial cells is indicated (scale bars—60 μm).</p>
Full article ">Figure 7
<p>Number of adhered platelets to polyurethane (PU) and CyanoCoating per µm<sup>2</sup> after 30 min of incubation at 37 °C in the presence or absence of human plasma proteins (n = 9). Statistical analysis was performed by non-parametric Kruskal–Wallis analysis and differences are indicated with **** (<span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 8
<p>Metabolic activity of L929 mouse fibroblasts after 24 h incubation with polyurethane (PU) and CyanoCoating extracts diluted 1:1 (dil) or non-diluted, assessed using the MTT cytotoxicity test (n = 12). Statistical analysis was performed by a non-parametric Kruskal–Wallis analysis.</p>
Full article ">
18 pages, 481 KiB  
Review
Biological Activities of Fucoidan and the Factors Mediating Its Therapeutic Effects: A Review of Recent Studies
by Yu Wang, Maochen Xing, Qi Cao, Aiguo Ji, Hao Liang and Shuliang Song
Mar. Drugs 2019, 17(3), 183; https://doi.org/10.3390/md17030183 - 20 Mar 2019
Cited by 325 | Viewed by 15145
Abstract
The marine acid polysaccharide fucoidan has attracted attention from both the food and pharmaceutical industries due to its promising therapeutic effects. Fucoidan is a polysaccharide that mainly consists of L-fucose and sulphate groups. Its excellent biological function is attributed to its unique biological [...] Read more.
The marine acid polysaccharide fucoidan has attracted attention from both the food and pharmaceutical industries due to its promising therapeutic effects. Fucoidan is a polysaccharide that mainly consists of L-fucose and sulphate groups. Its excellent biological function is attributed to its unique biological structure. Classical activities include antitumor, antioxidant, anticoagulant, antithrombotic, immunoregulatory, antiviral and anti-inflammatory effects. More recently, fucoidan has been shown to alleviate metabolic syndrome, protect the gastrointestinal tract, benefit angiogenesis and bone health. This review focuses on the progress in our understanding of the biological activities of fucoidan, highlighting its benefits for the treatment of human disease. We hope that this review can provide some theoretical basis and inspiration for the product development of fucoidan. Full article
Show Figures

Figure 1

Figure 1
<p>Biological activities of fucoidan.</p>
Full article ">
14 pages, 2095 KiB  
Article
Metabolomic and Transcriptomic Analyses of Escherichia coli for Efficient Fermentation of L-Fucose
by Jungyeon Kim, Yu Eun Cheong, Inho Jung and Kyoung Heon Kim
Mar. Drugs 2019, 17(2), 82; https://doi.org/10.3390/md17020082 - 29 Jan 2019
Cited by 19 | Viewed by 5952
Abstract
L-Fucose, one of the major monomeric sugars in brown algae, possesses high potential for use in the large-scale production of bio-based products. Although fucose catabolic pathways have been enzymatically evaluated, the effects of fucose as a carbon source on intracellular metabolism in industrial [...] Read more.
L-Fucose, one of the major monomeric sugars in brown algae, possesses high potential for use in the large-scale production of bio-based products. Although fucose catabolic pathways have been enzymatically evaluated, the effects of fucose as a carbon source on intracellular metabolism in industrial microorganisms such as Escherichia coli are still not identified. To elucidate the effects of fucose on cellular metabolism and to find clues for efficient conversion of fucose into bio-based products, comparative metabolomic and transcriptomic analyses were performed on E. coli on L-fucose and on D-glucose as a control. When fucose was the carbon source for E. coli, integration of the two omics analyses revealed that excess gluconeogenesis and quorum sensing led to severe depletion of ATP, resulting in accumulation and export of fucose extracellularly. Therefore, metabolic engineering and optimization are needed for E. coil to more efficiently ferment fucose. This is the first multi-omics study investigating the effects of fucose on cellular metabolism in E. coli. These omics data and their biological interpretation could be used to assist metabolic engineering of E. coli producing bio-based products using fucose-containing brown macroalgae. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Comparison of the growth and fermentation product profiles of <span class="html-italic">E. coli</span> cultured on fucose and glucose as the carbon source (mean ± SD): (<b>A</b>) cell density recorded as optical density at 600 nm (OD<sub>600</sub>); (<b>B</b>) concentrations of substrate; and (<b>C</b>) by-products (three independent replicates).</p>
Full article ">Figure 2
<p>PCA (<b>A</b>) score and (<b>B</b>) loading plots of 102 intracellular metabolites of <span class="html-italic">E. coli</span> in the lag, exponential, and stationary phases cultured on fucose and glucose as the carbon source (six replicates; three independent replicates × two technical replicates).</p>
Full article ">Figure 3
<p>MetaMapp analysis of 102 intracellular metabolites of <span class="html-italic">E. coli</span> cultured on fucose and glucose as the carbon source in: (<b>A</b>) the lag phase; (<b>B</b>) exponential phase; and (<b>C</b>) stationary phase. Classes of metabolites are represented by shapes. Significant increases and decreases in metabolite abundance are represented by color (<span class="html-italic">p</span> &lt; 0.05). Magnitudes of fold changes are represented by the size of symbols and labels. Biochemical and structural similarities are represented by the orange and gray edges, respectively (six replicates; three independent replicates × two technical replicates).</p>
Full article ">Figure 4
<p>Hierarchical clustering analysis (FDR adjusted <span class="html-italic">p</span>-value &lt; 0.01, ANOVA) of significantly increased or decreased (<b>A</b>) intracellular and (<b>B</b>) extracellular metabolites of <span class="html-italic">E. coli</span> in the lag, exponential, and stationary phases cultured on fucose as the carbon source. Clustering of the model was based on Pearson’s correlation coefficient and average linkage methods (six replicates; three independent replicates × two technical replicates).</p>
Full article ">Figure 5
<p>Comparison of transcription levels in (<b>A</b>) central carbon metabolism and (<b>B</b>) abundance of cofactors of <span class="html-italic">E. coli</span> in the exponential phase cultured on fucose and glucose. Significant changes to transcript levels are represented by color (<span class="html-italic">p</span>-value &lt; 0.05, and fold changes &gt; 2.0). Significant changes in cofactors are represented by * (three independent replicates; <span class="html-italic">p</span>-value &lt; 0.05).</p>
Full article ">Figure 6
<p>Heat map of 284 transcripts significantly changed (<span class="html-italic">p</span>-value &lt; 0.05) in <span class="html-italic">E. coli</span> in the exponential phase cultured on fucose and glucose (three independent replicates).</p>
Full article ">
10 pages, 1727 KiB  
Article
Activation of Human Dendritic Cells by Ascophyllan Purified from Ascophyllum nodosum
by Wei Zhang, Minseok Kwak, Hae-Bin Park, Takasi Okimura, Tatsuya Oda, Peter Chang-Whan Lee and Jun-O Jin
Mar. Drugs 2019, 17(1), 66; https://doi.org/10.3390/md17010066 - 19 Jan 2019
Cited by 14 | Viewed by 4374
Abstract
In our previous study, we showed that ascophyllan purified from Ascophyllum nodosum treatment promotes mouse dendritic cell (DC) activation in vivo, further induces an antigen-specific immune response and has anticancer effects in mice. However, the effect of ascophyllan has not been studied in [...] Read more.
In our previous study, we showed that ascophyllan purified from Ascophyllum nodosum treatment promotes mouse dendritic cell (DC) activation in vivo, further induces an antigen-specific immune response and has anticancer effects in mice. However, the effect of ascophyllan has not been studied in human immune cells, specifically in terms of activation of human monocyte-derived DCs (MDDCs) and human peripheral blood DCs (PBDCs). We found that the treatment with ascophyllan induced morphological changes in MDDCs and upregulated co-stimulatory molecules and major histocompatibility complex class I (MHC I) and MHC II expression. In addition, pro-inflammatory cytokine levels in culture medium was also dramatically increased following ascophyllan treatment of MDDCs. Moreover, ascophyllan promoted phosphorylation of ERK, p38 and JNK signaling pathways, and inhibition of p38 almost completely suppressed the ascophyllan-induced activation of MDDCs. Finally, treatment with ascophyllan induced activation of BDCA1 and BDCA3 PBDCs. Thus, these data suggest that ascophyllan could be used as an immune stimulator in humans. Full article
Show Figures

Figure 1

Figure 1
<p>Activation of human monocyte-derived dendritic cells (MDDCs) by ascophyllan. CD14<sup>+</sup> monocytes were differentiated to MDDCs by culturing with 50 ng/mL of granulocyte-macrophage colony-stimulating factor (GM-CSF) and 50 ng/mL of interleukin-4 (IL-4) for 6 days. (<b>A</b>) Changes in morphology are shown 24 h after treatment with PBS, ascophyllan (asco) or fucoidan (fuco). (<b>B</b>) Expression levels of CD80 (left panel) and CD83 (right panel) in MDDCs 24 h after treatment with different doses of ascophyllan. (<b>C</b>) Expression levels of co-stimulatory molecules were measured 24 h after treatment of PBS, ascophyllan (50 μg/mL) and fucoidan (50 μg/mL). Mean fluorescence intensity (MFI) levels of CD80, CD83, CD86, MHC class I and MHC class II in MDDCs are shown. Data are representative of or the average of analyses of 6 independent samples. ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 2
<p>Phosphorylation of mitogen-activated protein kinase (MAPK) signaling pathway in MDDCs by ascophyllan. (<b>A</b>) Phosphorylation of ERK, p38 and JNK was analyzed in MDDCs by western blotting after 1 h of ascophyllan (50 μg/mL) or fucoidan (50 μg/mL) treatment. (B and C) MDDCs were pre-treated with PD98059 (ERK inhibitor; 10 μM), SB203580 (p38 inhibitor; 2μM) or SP600125 (JNK inhibitor; 10 μM) for 1 h and then stimulated with ascophyllan for 24 h. (<b>B</b>) MFI levels of CD80 and (<b>C</b>) CD86 are shown. All data are representative of or the average of analyses of 6 independent samples. ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 3
<p>Production of pro-inflammatory cytokines from MDDCs by ascophyllan. MDDCs were treated with PBS, 50 μg/mL ascophyllan (asco) or 50 μg/mL fucoidan (fuco) for 24 h and culture medium was harvested. (<b>A</b>) The concentration of interleukin-6 (IL-6), IL-12 and TNF-α in a culture medium. (<b>B</b>) MDDCs were pre-treated with inhibitors as shown in <a href="#marinedrugs-17-00066-f002" class="html-fig">Figure 2</a>B,C, then cultured with PBS, ascophyllan or fucoidan for 24 h. IL-6, IL-12 and TNF-α levels were measured using ELISA. All data are the average of analyses of 6 independent samples. ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 4
<p>Activation of peripheral blood dendritic cell (PBDC) subsets following treatment with ascophyllan. Peripheral blood mononuclear cells (PBMCs) were cultured with PBS, 50 μg/mL ascophyllan (asco) or 50 μg/mL fucoidan (fuco) for 24 h. (<b>A</b>) BDCA1<sup>+</sup> and BDCA3<sup>+</sup> PBDCs were defined by flow cytometry from CD11c<sup>+</sup>lineage<sup>−</sup> live leukocytes of PBMCs. (<b>B</b>) Co-stimulatory molecules and MHC class I and II expression in BDCA1<sup>+</sup> and BDCA3<sup>+</sup> PBDCs are shown. (<b>C</b>) PD98059 (ERK inhibitor; 10 μM), SB203580 (p38 inhibitor; 2 μM) or SP600125 (JNK inhibitor; 10 μM) were pre-treated in PBDCs for 1 h and the PBDCs were stimulated with 50 μg/mL ascophyllan for 24 h. Expression levels of CD83 were measured in BDCA1<sup>+</sup> and BDCA3<sup>+</sup> PBDCs. (<b>D</b>) Concentration of IL-6, IL-12 and TNF-α in a culture medium of PBMCs. (<b>E</b>) Concentrations of IL-6, IL-12, TNF-α in culture medium from <a href="#marinedrugs-17-00066-f004" class="html-fig">Figure 4</a>C were measured using ELISA. All data are representative of or the average of analyses of 6 independent samples. ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">
14 pages, 3073 KiB  
Article
Protective Effect of Low Molecular Weight Seleno-Aminopolysaccharide on the Intestinal Mucosal Oxidative Damage
by Zheng-Shun Wen, Zhen Tang, Li Ma, Tian-Long Zhu, You-Ming Wang, Xing-Wei Xiang and Bin Zheng
Mar. Drugs 2019, 17(1), 64; https://doi.org/10.3390/md17010064 - 18 Jan 2019
Cited by 15 | Viewed by 3855
Abstract
Low molecular weight seleno-aminopolysaccharide (LSA) is an organic selenium compound comprising selenium and low molecular weight aminopolysaccharide (LA), a low molecular weight natural linear polysaccharide derived from chitosan. LSA has been found to exert strong pharmacological activity. In this study, we aimed to [...] Read more.
Low molecular weight seleno-aminopolysaccharide (LSA) is an organic selenium compound comprising selenium and low molecular weight aminopolysaccharide (LA), a low molecular weight natural linear polysaccharide derived from chitosan. LSA has been found to exert strong pharmacological activity. In this study, we aimed to investigate the protective effect of LSA on intestinal mucosal oxidative stress in a weaning piglet model by detecting the growth performance, intestinal mucosal structure, antioxidant indices, and expression level of intracellular transcription factor nuclear factor erythroid 2-related factor 2 (Nrf2) and its related factors. Our results indicated that LSA significantly increased the average daily gain and feed/gain (p < 0.05), suggesting that LSA can effectively promote the growth of weaning piglets. The results of scanning electron microscope (SEM) microscopy showed that LSA effectively reduced intestinal damage, indicating that LSA improved the intestinal stress response and protected the intestinal structure integrity. In addition, diamine oxidase (DAO) and d-lactic acid (d-LA) levels remarkably decreased in LSA group compared with control group (p < 0.05), suggesting that LSA alleviated the damage and permeability of weaning piglets. LSA significantly increased superoxide dismutase (SOD), catalase (CAT), glutathione peroxidase (GSH-Px), and total antioxidant capacity (T-AOC) levels, but decreased malondialdehyde (MDA) level, indicating that LSA significantly enhanced the antioxidant capacity and reduced oxidative stress in weaning piglets. RT-PCR results showed that LSA significantly increased GSH-Px1, GSH-Px2, SOD-1, SOD-2, CAT, Nrf2, HO-1, and NQO1 gene expression (p < 0.05). Western blot analysis revealed that LSA activated the Nrf2 signaling pathway by downregulating the expression of Keap1 and upregulating the expression of Nrf2 to protect intestinal mucosa against oxidative stress. Collectively, LSA reduced intestinal mucosal damage induced by oxidative stress via Nrf2-Keap1 pathway in weaning stress of infants. Full article
Show Figures

Figure 1

Figure 1
<p>Effects of LSA on diarrhea rate. Bars labeled with different letters (a–d) were significantly different (<span class="html-italic">p</span> &lt; 0.05) from each other.</p>
Full article ">Figure 2
<p>Effects of LSA on intestinal morphology (SEM, 200×).</p>
Full article ">Figure 3
<p>Effects of LSA on the level of diamine oxidase (DAO) and <span class="html-small-caps">d</span>-lactic acid (<span class="html-small-caps">d</span>-LA) in serum. Bars labeled with different letters (a–d) were significantly different (<span class="html-italic">p</span> &lt; 0.05) from each other.</p>
Full article ">Figure 4
<p>Effects of LSA on the antioxidant indices in serum. Bars labeled with different letters (a–d) were significantly different (<span class="html-italic">p</span> &lt; 0.05) from each other.</p>
Full article ">Figure 5
<p>Effect of LSA on the gene expression levels of tight junction proteins in ileum. Bars labeled with different letters (a, b) were significantly different (<span class="html-italic">p</span> &lt; 0.05) from each other.</p>
Full article ">Figure 6
<p>Effects of LSA on the expression of antioxidant genes in ileum. The mRNA expression level of antioxidant genes (GSH-Px1 (<b>a</b>), GSH-Px2 (<b>b</b>), SOD1 (<b>c</b>), SOD2 (<b>d</b>), CAT (<b>e</b>), Nrf2 (<b>f</b>), NQO1 (<b>g</b>), HO-1 (<b>h</b>). Bars labeled with different letters (a–d) were significantly different (<span class="html-italic">p</span> &lt; 0.05) from each other.</p>
Full article ">Figure 7
<p>Effects of LSA on the Keap1–Nrf2 signaling pathway in ileum. Bars labeled with different letters (a-d) were significantly different (<span class="html-italic">p</span> &lt; 0.05) from each other.</p>
Full article ">
15 pages, 2811 KiB  
Communication
Pathway Analysis of Fucoidan Activity Using a Yeast Gene Deletion Library Screen
by Monika Corban, Mark Ambrose, Joanne Pagnon, Damien Stringer, Sam Karpiniec, Ahyoung Park, Raj Eri, J Helen Fitton and Nuri Gueven
Mar. Drugs 2019, 17(1), 54; https://doi.org/10.3390/md17010054 - 14 Jan 2019
Cited by 10 | Viewed by 4946
Abstract
Fucoidan, the sulfated fucose-rich polysaccharide derived from brown macroalgae, was reported to display some anti-cancer effects in in vitro and in vivo models that included apoptosis and cell cycle arrest. The proposed mechanisms of action involve enhanced immune surveillance and direct pro-apoptotic effects [...] Read more.
Fucoidan, the sulfated fucose-rich polysaccharide derived from brown macroalgae, was reported to display some anti-cancer effects in in vitro and in vivo models that included apoptosis and cell cycle arrest. The proposed mechanisms of action involve enhanced immune surveillance and direct pro-apoptotic effects via the activation of cell signaling pathways that remain largely uncharacterized. This study aimed to identify cellular pathways influenced by fucoidan using an unbiased genetic approach to generate additional insights into the anti-cancer effects of fucoidan. Drug–gene interactions of Undaria pinnatifida fucoidan were assessed by a systematic screen of the entire set of 4,733 halpoid Saccharomyces cerevsiae gene deletion strains. Some of the findings were confirmed using cell cycle analysis and DNA damage detection in non-immortalized human dermal fibroblasts and colon cancer cells. The yeast deletion library screen and subsequent pathway and interactome analysis identified global effects of fucoidan on a wide range of eukaryotic cellular processes, including RNA metabolism, protein synthesis, sorting, targeting and transport, carbohydrate metabolism, mitochondrial maintenance, cell cycle regulation, and DNA damage repair-related pathways. Fucoidan also reduced clonogenic survival, induced DNA damage and G1-arrest in colon cancer cells, while these effects were not observed in non-immortalized human fibroblasts. Our results demonstrate global effects of fucoidan in diverse cellular processes in eukaryotic cells and further our understanding about the inhibitory effect of Undaria pinnatifida fucoidan on the growth of human cancer cells. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Pathway analysis of identified genes that resulted in reduced growth of <span class="html-italic">S. cerevisiae</span> in the presence of <span class="html-italic">UPF</span>. Pathway analysis was performed with String software (Version 10.5) using an interaction score of 0.9 (‘highest confidence’). The figure only shows connected genes with disconnected nodes hidden. (<b>B</b>) Pathway analysis of identified genes that resulted in increased growth of <span class="html-italic">S. cerevisiae</span> in the presence of <span class="html-italic">UPF</span>. Pathway analysis was performed with String software (Version 10.5) using an interaction score of 0.9 (‘highest confidence’). The figure only shows connected genes with disconnected nodes hidden. (<b>C</b>) Pathway analysis of all identified genes that interacted with <span class="html-italic">UPF</span>. Pathway analysis was performed with String software (Version 10.5) using an interaction score of 0.9 (‘highest confidence’). The figure only shows connected genes with disconnected nodes hidden.</p>
Full article ">Figure 1 Cont.
<p>(<b>A</b>) Pathway analysis of identified genes that resulted in reduced growth of <span class="html-italic">S. cerevisiae</span> in the presence of <span class="html-italic">UPF</span>. Pathway analysis was performed with String software (Version 10.5) using an interaction score of 0.9 (‘highest confidence’). The figure only shows connected genes with disconnected nodes hidden. (<b>B</b>) Pathway analysis of identified genes that resulted in increased growth of <span class="html-italic">S. cerevisiae</span> in the presence of <span class="html-italic">UPF</span>. Pathway analysis was performed with String software (Version 10.5) using an interaction score of 0.9 (‘highest confidence’). The figure only shows connected genes with disconnected nodes hidden. (<b>C</b>) Pathway analysis of all identified genes that interacted with <span class="html-italic">UPF</span>. Pathway analysis was performed with String software (Version 10.5) using an interaction score of 0.9 (‘highest confidence’). The figure only shows connected genes with disconnected nodes hidden.</p>
Full article ">Figure 2
<p>Effect of <span class="html-italic">UPF</span> on metabolic activity/viability. Human colon carcinoma cells (HCT-116) were exposed to <span class="html-italic">UPF</span> concentrations of up to 100 µg/mL for 24 h before viability was assessed using WST-1 reagent. Data represent one typical experiment out of up to four independent experiments. WST-1 absorption data were standardized on protein content for each well, represent the mean of six individual wells per experiment and are expressed as % viability compared to the untreated control cells. Error bars represent SD with *: <span class="html-italic">p</span> &lt; 0.05. Hydrogen peroxide (H, 100 µM) was used as a positive control for toxicity.</p>
Full article ">Figure 3
<p>Effect of <span class="html-italic">UPF</span> on colony formation. Human colon carcinoma cells (HCT-116, <b>A</b>) and human non-immortalized dermal fibroblasts (HDF, <b>B</b>) were exposed to <span class="html-italic">UPF</span> concentrations up to 100 µg/mL in a colony formation assay. Data represent one typical experiment out of up to four independent experiments. Data is expressed as the mean +/− SD of four plates and expressed as % colony formation compared to the untreated control cells. Error bars represent SD with #: <span class="html-italic">p</span> &lt; 0.05 and * <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 4
<p>DNA damage induction by <span class="html-italic">UPF</span>. Human colon carcinoma cells (HCT-116, <b>A</b>, <b>C</b>) and human non-immortalized dermal fibroblasts (HDF, <b>A</b>,<b>B</b>) were exposed to 100 µg/mL <span class="html-italic">UPF</span> before γH<sub>2</sub>AX immunostaining was performed (<b>A</b>) Data represents the mean of 3 experiments and expressed as % γH<sub>2</sub>AX-positive cells. Error bars represent SD with *: <span class="html-italic">p</span> &lt; 0.001. (<b>B</b>) Data represent one typical experiment out of three experiments for HDF. n/a: cells excluded from analysis due to unquantifiable staining pattern. (<b>C</b>) Representative images for <span class="html-italic">UPF</span>-induced induction of nuclear γH<sub>2</sub>AX foci in HCT-116 cells. H<sub>2</sub>O<sub>2</sub>-treatment (100 µM, 30 min) was used as a positive control. Merged images represent software generated false color overlays of DAPI and γH<sub>2</sub>AX signals.</p>
Full article ">Figure 4 Cont.
<p>DNA damage induction by <span class="html-italic">UPF</span>. Human colon carcinoma cells (HCT-116, <b>A</b>, <b>C</b>) and human non-immortalized dermal fibroblasts (HDF, <b>A</b>,<b>B</b>) were exposed to 100 µg/mL <span class="html-italic">UPF</span> before γH<sub>2</sub>AX immunostaining was performed (<b>A</b>) Data represents the mean of 3 experiments and expressed as % γH<sub>2</sub>AX-positive cells. Error bars represent SD with *: <span class="html-italic">p</span> &lt; 0.001. (<b>B</b>) Data represent one typical experiment out of three experiments for HDF. n/a: cells excluded from analysis due to unquantifiable staining pattern. (<b>C</b>) Representative images for <span class="html-italic">UPF</span>-induced induction of nuclear γH<sub>2</sub>AX foci in HCT-116 cells. H<sub>2</sub>O<sub>2</sub>-treatment (100 µM, 30 min) was used as a positive control. Merged images represent software generated false color overlays of DAPI and γH<sub>2</sub>AX signals.</p>
Full article ">Figure 5
<p>Effect of <span class="html-italic">UPF</span> on cell cycle distribution. HCT 116 cells were treated with 100 µg/mL <span class="html-italic">UPF</span> for up to 72 h and cell cycle distribution was assessed by flow cytometry. Data represents the mean of four independent experiments performed over a two-month time period. Error bars represent SD with *: <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">
18 pages, 3327 KiB  
Article
Structure Analysis and Anti-Tumor and Anti-Angiogenic Activities of Sulfated Galactofucan Extracted from Sargassum thunbergii
by Weihua Jin, Wanli Wu, Hong Tang, Bin Wei, Hong Wang, Jiadong Sun, Wenjing Zhang and Weihong Zhong
Mar. Drugs 2019, 17(1), 52; https://doi.org/10.3390/md17010052 - 11 Jan 2019
Cited by 34 | Viewed by 4543
Abstract
Sulfated galactofucan (ST-2) was obtained from Sargassum thunbergii. It was then desulfated to obtain ST-2-DS, and autohydrolyzed and precipitated by ethanol to obtain the supernatant (ST-2-S) and precipitate (ST-2-C). ST-2-C was further fractionated by gel chromatography into two fractions, ST-2-H (high molecular [...] Read more.
Sulfated galactofucan (ST-2) was obtained from Sargassum thunbergii. It was then desulfated to obtain ST-2-DS, and autohydrolyzed and precipitated by ethanol to obtain the supernatant (ST-2-S) and precipitate (ST-2-C). ST-2-C was further fractionated by gel chromatography into two fractions, ST-2-H (high molecular weight) and ST-2-L (low molecular weight). Mass spectrometry (MS) of ST-2-DS was performed to elucidate the backbone of ST-2. It was shown that ST-2-DS contained a backbone of alternating galactopyranose residues (Gal)n (n ≤ 3) and fucopyranose residues (Fuc)n. In addition, ST-2-S was also determined by MS to elucidate the branches of ST-2. It was suggested that sulfated fuco-oligomers might be the branches of ST-2. Compared to the NMR spectra of ST-2-H, the spectra of ST-2-L was more recognizable. It was shown that ST-2-L contain a backbone of (Gal)n and (Fuc)n, sulfated mainly at C4 of Fuc, and interspersed with galactose (the linkages were likely to be 1→2 and 1→6). Therefore, ST-2 might contain a backbone of (Gal)n (n ≤ 3) and (Fuc)n. The sulfation pattern was mainly at C4 of fucopyranose and partially at C4 of galactopyranose, and the branches were mainly sulfated fuco-oligomers. Finally, the anti-tumor and anti-angiogenic activities of ST-2 and its derivates were determined. It was shown that the low molecular-weight sulfated galactofucan, with higher fucose content, had better anti-angiogenic and anti-tumor activities. Full article
Show Figures

Figure 1

Figure 1
<p>IR spectra of ST-2 (<b>A</b>), ST-2-DS (<b>B</b>), ST-2-H (<b>C</b>), and ST-2-L (<b>D</b>).</p>
Full article ">Figure 1 Cont.
<p>IR spectra of ST-2 (<b>A</b>), ST-2-DS (<b>B</b>), ST-2-H (<b>C</b>), and ST-2-L (<b>D</b>).</p>
Full article ">Figure 2
<p>Negative-ion mode ESI-MS spectrum of ST-2-DS.</p>
Full article ">Figure 3
<p>Negative-ion mode electrospray mass spectrometry in tandem with collision-induced dissociation tandem mass spectrometry (ESI-CID-MS/MS) spectra of the ion at <span class="html-italic">m</span>/<span class="html-italic">z</span> 403.089 (−1) (<b>A</b>), 565.141 (−1) (<b>B</b>), 330.024 (−2) (<b>C</b>), 581.136 (−1) (<b>D</b>), and 889.244 (−1) (<b>E</b>).</p>
Full article ">Figure 3 Cont.
<p>Negative-ion mode electrospray mass spectrometry in tandem with collision-induced dissociation tandem mass spectrometry (ESI-CID-MS/MS) spectra of the ion at <span class="html-italic">m</span>/<span class="html-italic">z</span> 403.089 (−1) (<b>A</b>), 565.141 (−1) (<b>B</b>), 330.024 (−2) (<b>C</b>), 581.136 (−1) (<b>D</b>), and 889.244 (−1) (<b>E</b>).</p>
Full article ">Figure 3 Cont.
<p>Negative-ion mode electrospray mass spectrometry in tandem with collision-induced dissociation tandem mass spectrometry (ESI-CID-MS/MS) spectra of the ion at <span class="html-italic">m</span>/<span class="html-italic">z</span> 403.089 (−1) (<b>A</b>), 565.141 (−1) (<b>B</b>), 330.024 (−2) (<b>C</b>), 581.136 (−1) (<b>D</b>), and 889.244 (−1) (<b>E</b>).</p>
Full article ">Figure 4
<p>Negative-ion mode ESI-MS spectrum of ST-2-S.</p>
Full article ">Figure 5
<p>The DEPTQ NMR spectrum (A), the <sup>13</sup>C NMR spectrum (B), the <sup>1</sup>H NMR spectrum (C), and the HSQC spectrum (D) of ST-2-L.</p>
Full article ">Figure 5 Cont.
<p>The DEPTQ NMR spectrum (A), the <sup>13</sup>C NMR spectrum (B), the <sup>1</sup>H NMR spectrum (C), and the HSQC spectrum (D) of ST-2-L.</p>
Full article ">Figure 6
<p>The proposed structure scheme of ST-2-L.</p>
Full article ">Figure 7
<p>Anti-tumor activities of polysaccharides (ST-2, ST-2-H, and ST-2-L) against human lung cancer A549 cells (<b>A</b>) and anti-angiogenic activities against human umbilical vein endothelial cells (HUVEC) (<b>B</b>). The results are expressed as percent inhibition. Data are the mean of three determinations +/− SEM.</p>
Full article ">

2018

Jump to: 2024, 2022, 2021, 2020, 2019, 2017, 2016, 2015, 2014, 2013, 2012, 2011, 2010

13 pages, 3621 KiB  
Article
Isolation, Characterization, and Pharmaceutical Applications of an Exopolysaccharide from Aerococcus Uriaeequi
by Chunlei Wang, Qiuping Fan, Xiaofei Zhang, Xiaoping Lu, Yanrui Xu, Wenxing Zhu, Jie Zhang, Wen Hao and Lujiang Hao
Mar. Drugs 2018, 16(9), 337; https://doi.org/10.3390/md16090337 - 16 Sep 2018
Cited by 34 | Viewed by 4977
Abstract
Many marine bacteria secrete exopolysaccharides (EPSs), which are made up of a substantial component of the macro-molecules surrounding cells. Recently, the wide demand for EPSs for food, cosmetics, pharmaceutical and other applications has led to great interest in them. In this study, an [...] Read more.
Many marine bacteria secrete exopolysaccharides (EPSs), which are made up of a substantial component of the macro-molecules surrounding cells. Recently, the wide demand for EPSs for food, cosmetics, pharmaceutical and other applications has led to great interest in them. In this study, an EPS produced by marine bacteria Aerococcus uriaeequi HZ strains (EPS-A) was isolated and purified to examine its structure and biological function. The molecular weight of EPS-A analyzed by high-performance liquid gel filtration chromatography (HPGFC) is found to have a number average of 2.22 × 105 and weight average of 2.84 × 105, respectively. High-performance liquid chromatography (HPLC) and Fourier-transform–infrared (FT–IR) analysis indicate that EPS-A was a polysaccharide composed of glucose and a little mannose. In addition, the flocculating rate of sewage of EPS-A was 79.90%. The hygroscopicity studies showed that hygroscopicity of EPS-A was higher than chitosan but lower than that of sodium hyaluronate. The moisture retention of EPS-A showed similar retention activity to both chitosan and sodium hyaluronate. EPS-A also can scavenge free radicals including both OH• free radical and O2 free radical and the activity to O2 free radical is similar to vitamin C. Safety assessment on mice indicated that the EPS-A is safe for external use and oral administration. EPS-A has great potential for applications in medicine due to its characteristics mentioned above. Full article
Show Figures

Figure 1

Figure 1
<p>Purification of EPS-A. (<b>A</b>) Mw distribution of EPS-A determination by high-performance liquid gel permeation chromatography (HPGFC). HPGFC was performed using a Shodex SB-806HQ column in 0.2 M NaCl solution with 0.5 mL/min flow rate. (<b>B</b>) Ultraviolet (UV)-visible spectrum of EPS-A. The absorbance from 200–500 nm was measured in H<sub>2</sub>O at room temperature.</p>
Full article ">Figure 2
<p>High-performance liquid chromatography (HPLC) chromatograms of seven PMP-labeled standard monosaccharides (<b>A</b>) and PMP-labeled monosaccharides released from EPS-A (<b>B</b>). Peaks: 1. D-Mannose; 2. D-Rhamnose; 3. D-Glucuronic acid; 4. D-Galacturonic acid; 5. D-Glucose; 6. D-Galactose;7. d-Xylose.</p>
Full article ">Figure 2 Cont.
<p>High-performance liquid chromatography (HPLC) chromatograms of seven PMP-labeled standard monosaccharides (<b>A</b>) and PMP-labeled monosaccharides released from EPS-A (<b>B</b>). Peaks: 1. D-Mannose; 2. D-Rhamnose; 3. D-Glucuronic acid; 4. D-Galacturonic acid; 5. D-Glucose; 6. D-Galactose;7. d-Xylose.</p>
Full article ">Figure 3
<p>FT-IR spectra of EPS-A. Dried polysaccharides were ground and pelletized with KBr. Ultraviolet-visible spectrum of EPS-A was recorded with a spectrophotometer from 500–4000 cm.</p>
Full article ">Figure 4
<p>Sewage flocculation assay of EPS-A. 0.2 g EPS was added l00 mL sewage and incubated for 1 h. The function of sewage flocculation was measured at 550 nm by a spectrophotometer.</p>
Full article ">Figure 5
<p>Moisture-absorption and retention activity of EPS-A. (<b>A</b>) Hygroscopic activity assay of EPS-A. Hygroscopic activity of EPS-A was determined by measuring the increased weight of absorbing H<sub>2</sub>O by EPS-A. Chitosan and sodium hyaluronate were used as controls. The value obtained at 50 h by sodium hyaluronate was set 100%. (<b>B</b>) Moisture retention activity assay of EPS-A. Moisture retention of EPS-A was determined by measuring the reserved weight of H<sub>2</sub>O by EPS-A. Chitosan and sodium hyaluronate were used as controls. And the value at the beginning was set 100%.</p>
Full article ">Figure 6
<p>Free radical scavenging activity of EPS-A. <b>A</b> OH• free radical scavenging activity of EPS-A. The scavenging activity to OH• of different concentration EPS was determined by removing OH• generated by FeSO<sub>4</sub> and H<sub>2</sub>O<sub>2</sub>. Vitamin C was used as a control and the activity of 100 µg/mL vitamin C was set 100%. <b>B</b> O<sub>2</sub>•<sup>−</sup> free radical scavenging activity of EPS-A. The scavenging activity to O<sub>2</sub>•<sup>−</sup> of different concentration EPS-A was determined by removing O<sub>2</sub>•<sup>−</sup> generated from pyrogallol. Vitamin C was used as a control and the activity of 100 µg/mL Vitamin C was set 100%.</p>
Full article ">Figure 7
<p>Effect of EPS-A on the body weight (<b>A</b>) and splenic indices (<b>B</b>) of mice in a 14-day feeding test.</p>
Full article ">Figure 8
<p>Schematic diagram of EPS-A with the potential advantages (e.g., Sewage flocculation, Hygroscopic activity, Moisture retention, •OH free radical scavenging, O<sup>2−</sup> • free radicals scavenging). The safety assessment will be performed on mice.</p>
Full article ">Figure 9
<p>Standard curve of HPGPC-PSS series of standard samples.</p>
Full article ">
12 pages, 3585 KiB  
Article
Characterization of a Novel PolyM-Preferred Alginate Lyase from Marine Vibrio splendidus OU02
by Jingjing Zhuang, Keke Zhang, Xiaohua Liu, Weizhi Liu, Qianqian Lyu and Aiguo Ji
Mar. Drugs 2018, 16(9), 295; https://doi.org/10.3390/md16090295 - 22 Aug 2018
Cited by 35 | Viewed by 4667
Abstract
Alginate lyases are enzymes that degrade alginate into oligosaccharides which possess a variety of biological activities. Discovering and characterizing novel alginate lyases has great significance for industrial and medical applications. In this study, we reported a novel alginate lyase, AlyA-OU02, derived from the [...] Read more.
Alginate lyases are enzymes that degrade alginate into oligosaccharides which possess a variety of biological activities. Discovering and characterizing novel alginate lyases has great significance for industrial and medical applications. In this study, we reported a novel alginate lyase, AlyA-OU02, derived from the marine Vibrio splendidus OU02. The BLASTP searches showed that AlyA-OU02 belonged to polysaccharide lyase family 7 (PL7) and contained two consecutive PL7 domains, which was rare among the alginate lyases in PL7 family. Both the two domains, AlyAa and AlyAb, had lyase activities, while AlyAa exhibited polyM preference, and AlyAb was polyG-preferred. In addition, the enzyme activity of AlyAa was much higher than AlyAb at 25 °C. The full-length enzyme of AlyA-OU02 showed polyM preference, which was the same as AlyAa. AlyAa degraded alginate into di-, tri-, and tetra-alginate oligosaccharides, while AlyAb degraded alginate into tri-, tetra-, and penta-alginate oligosaccharides. The degraded products of AlyA-OU02 were similar to AlyAa. Our work provided a potential candidate in the application of alginate oligosaccharide production and the characterization of the two domains might provide insights into the use of alginate of this organism. Full article
Show Figures

Figure 1

Figure 1
<p>Sequence properties of the alginate lyase AlyA-OU02 from marine <span class="html-italic">Vibrio splendidus</span> OU02. (<b>A</b>) Module organization of AlyA-OU02. The two alginate_lyase 2 modules were putative catalytic domains (Ser<sup>40</sup> to Ala<sup>264</sup>, Trp<sup>278</sup> to His<sup>563</sup>). The full-length protein and the two alginate_lyase 2 modules were expressed to yield the recombinant protein AlyA-OU02 (Ser<sup>1</sup> to Tyr<sup>564</sup>), AlyA<sup>a</sup> (Ser<sup>1</sup> to Ser<sup>275</sup>), and AlyA<sup>b</sup> (Asn<sup>276</sup> to Tyr<sup>564</sup>). The indicated amino acid residues were hypothesized catalytic sites. (<b>B</b>) Comparison of the partial amino acid sequences of AlyA<sup>a</sup> and AlyA<sup>b</sup> with PL7 alginate lyases AlyVI from <span class="html-italic">Vibrio</span> sp. QY101 (AAP45155.1), Algb from <span class="html-italic">Vibrio</span> sp. W13 (AIY22661.1), NitAly from <span class="html-italic">Nitratiruptor</span> sp. SB155-2 (BAF69299.1), AlxM from <span class="html-italic">Photobacterium</span> sp. ATCC 43367 (CAA49630.1).</p>
Full article ">Figure 2
<p>SDS-PAGE analysis of purified AlyA-OU02, AlyA<sup>a</sup>, and AlyA<sup>b</sup>. Lane M, molecular weight markers; Lane 1, purified AlyA-OU02; Lane 2, purified AlyA<sup>a</sup>; Lane 3, purified AlyA<sup>b</sup>.</p>
Full article ">Figure 3
<p>Enzyme activity assay of AlyA-OU02, AlyA<sup>a</sup> and AlyA<sup>b</sup> toward alginate. The measurement was carried out in Tris-HCl buffer with 200 mM NaCl (pH 7.5) using 0.2% alginate as substrate.</p>
Full article ">Figure 4
<p>Effects of temperature and pH on enzyme activities of AlyA-OU02, AlyA<sup>a</sup>, and AlyA<sup>b</sup>. (<b>A</b>) Optimal temperatures of the enzymes were determined in Tris-HCl buffer with 200 mM NaCl (pH 7.5) at different temperatures. (<b>B</b>) Optimal pH of the enzymes was determined at 25 °C in NaAc-HAc buffer (pH 5–6) and Tris-HCl buffer (pH 7–10).</p>
Full article ">Figure 5
<p>Thermostabilities of AlyA-OU02, AlyA<sup>a</sup>, and AlyA<sup>b</sup>.</p>
Full article ">Figure 6
<p>Effects of NaCl and metal ions on enzymatic activities of AlyA-OU02, AlyA<sup>a</sup>, and AlyA<sup>b</sup>. (<b>A</b>) Influence of the concentration of NaCl. (<b>B</b>) Influence of the metal ions on AlyA-OU02. (<b>C</b>) Influence of the metal ions on AlyA<sup>a</sup>. (<b>D</b>) Influence of the metal ions on AlyA<sup>b</sup>. The enzymatic activity without metal ions served as the control, and the enzymatic activity was designated as 100%.</p>
Full article ">Figure 7
<p>Substrate specificities of AlyA-OU02, AlyA<sup>a</sup>, and AlyA<sup>b</sup>. (<b>A</b>) Relative enzyme activities of AlyA<sup>a</sup> toward alginate, polyM, and polyG. (<b>B</b>) Relative enzyme activities of AlyA<sup>b</sup> toward alginate, polyM, and polyG. (<b>C</b>) Relative enzyme activities of AlyA-OU02 toward alginate, polyM, and polyG.</p>
Full article ">Figure 8
<p>TLC analysis of the hydrolytic products of AlyA-OU02, AlyA<sup>a</sup>, and AlyA<sup>b</sup>. Solutions of 0.2% (<span class="html-italic">w</span>/<span class="html-italic">v</span>) alginate, polyG, and polyM were each incubated with 5 uM AlyA-OU02, AlyA<sup>a</sup>, and AlyA<sup>b</sup>, respectively, at 25 °C for 1 h. Lane 1: alginate; lane 2–4: hydrolytic products of AlyA-OU02, AlyA<sup>a</sup>, and AlyA<sup>b</sup> toward alginate; lane 5: polyG; lane 6–8: hydrolytic products of AlyA-OU02, AlyA<sup>a</sup>, and AlyA<sup>b</sup> toward polyG; lane 9: polyM; lane 10–12: hydrolytic products of AlyA-OU02, AlyA<sup>a</sup>, and AlyA<sup>b</sup> toward polyM; lane 13: DP2, DP3, DP4, and DP5 mean the alginate disaccharide, trisaccharide, tetrasaccharide, and pentasaccharide, respectively; lane 14: hydrolytic products of AlyA<sup>a</sup> toward polyM.</p>
Full article ">
15 pages, 1330 KiB  
Article
Extraction and Yield Optimisation of Fucose, Glucans and Associated Antioxidant Activities from Laminaria digitata by Applying Response Surface Methodology to High Intensity Ultrasound-Assisted Extraction
by Marco Garcia-Vaquero, Gaurav Rajauria, Brijesh Tiwari, Torres Sweeney and John O’Doherty
Mar. Drugs 2018, 16(8), 257; https://doi.org/10.3390/md16080257 - 30 Jul 2018
Cited by 72 | Viewed by 8306
Abstract
The objectives of this study were to employ response surface methodology (RSM) to investigate and optimize the effect of ultrasound-assisted extraction (UAE) variables, temperature, time and amplitude on the yields of polysaccharides (fucose and total glucans) and antioxidant activities (ferric reducing antioxidant power [...] Read more.
The objectives of this study were to employ response surface methodology (RSM) to investigate and optimize the effect of ultrasound-assisted extraction (UAE) variables, temperature, time and amplitude on the yields of polysaccharides (fucose and total glucans) and antioxidant activities (ferric reducing antioxidant power (FRAP) and 1,1-diphenyl-2-picryl-hydrazyl radical scavenging activity (DPPH)) from Laminaria digitata, and to explore the suitability of applying the optimum UAE conditions for L. digitata to other brown macroalgae (L. hyperborea and Ascophyllum nodosum). The RSM with three-factor, four-level Box-Behnken Design (BBD) was used to study and optimize the extraction variables. A second order polynomial model fitted well to the experimental data with R2 values of 0.79, 0.66, 0.64, 0.73 for fucose, total glucans, FRAP and DPPH, respectively. The UAE parameters studied had a significant influence on the levels of fucose, FRAP and DPPH. The optimised UAE conditions (temperature = 76 °C, time = 10 min and amplitude = 100%) achieved yields of fucose (1060.7 ± 70.6 mg/100 g dried seaweed (ds)), total glucans (968.6 ± 13.3 mg/100 g ds), FRAP (8.7 ± 0.5 µM trolox/mg freeze-dried extract (fde)) and DPPH (11.0 ± 0.2%) in L. digitata. Polysaccharide rich extracts were also attained from L. hyperborea and A. nodosum with variable results when utilizing the optimum UAE conditions for L. digitata. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Contour plots (2D) and response surface plots (3D) of (<b>I</b>) fucose (mg/100 g dried seaweed (ds)); (<b>II</b>) total glucans (mg/100 g ds); (<b>III</b>) FRAP (µM trolox/mg freeze-dried extract (fde)) and (<b>IV</b>) DPPH (%) extracted from <span class="html-italic">Laminaria digitata</span> as a function of (<b>a</b>) time to temperature (amplitude = 100%) (<b>b</b>) temperature to amplitude (time = 30 min) and (<b>c</b>) amplitude to time (temperature = 80 °C).</p>
Full article ">Figure 2
<p>Scheme summarizing the pre-treatments of the fresh macroalgae and the ultrasound-assisted extraction (UAE) used to generate seaweed extracts.</p>
Full article ">
14 pages, 3233 KiB  
Article
Glycosaminoglycans from a Sea Snake (Lapemis curtus): Extraction, Structural Characterization and Antioxidant Activity
by Mingyue Bai, Wenwei Han, Xia Zhao, Qingchi Wang, Yanyun Gao and Shiming Deng
Mar. Drugs 2018, 16(5), 170; https://doi.org/10.3390/md16050170 - 18 May 2018
Cited by 18 | Viewed by 4797
Abstract
Sea snakes have wide application prospects in medicine, health food and other fields. Several novel polysaccharides were successfully obtained from the skin and the meat of a sea snake (Lapemis curtus). The structures of polysaccharides LSP3 and LMP3, which were extracted [...] Read more.
Sea snakes have wide application prospects in medicine, health food and other fields. Several novel polysaccharides were successfully obtained from the skin and the meat of a sea snake (Lapemis curtus). The structures of polysaccharides LSP3 and LMP3, which were extracted and purified from Lapemis curtus, were determined to be new and highly heterogenic glycosaminoglycans (GAGs) by means of FT-IR, ESI-MS/MS and NMR. LSP3 is a hybrid dermatan sulfate (DS) and composed of 48% 4-sulfated disaccharides (Di4S), 42% 6-sulfated disaccharides (Di6S) and 5% disulfated disaccharides (Di2,6S), while LMP3 is a hybrid chondroitin sulfate (CS) and composed of 70% Di4S, 20% Di6S, and 8% Di2,6S. More importantly, LSP3 and LMP3 showed a strong scavenging ability of 1,1-diphenyl-2-picrylhydrazyl (DPPH) radicals, iron (Fe2+) chelating activity and total antioxidant capacity in vitro, especially LSP3, with high contents of uronic acid and sulfate, which possessed a higher scavenging ability of DPPH radicals than other fractions. These data suggested that the sea snake polysaccharides could be promising candidates for natural antioxidant ingredients. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Experimental flowchart of <span class="html-italic">Lapemis curtus</span> polysacchloarides.</p>
Full article ">Figure 2
<p>Isolation of the polysaccharides LSP and LMP.</p>
Full article ">Figure 3
<p>Monosaccharide composition of LSP3 and LMP3. Chromatograms of the acid hydrolysates of LSP3 and LMP3, which were hydrolyzed in 3 mol/L TFA for 3 h at 110 °C to ensure a high response value of IdoA.</p>
Full article ">Figure 4
<p>The IR spectra of LSP3 and LMP3.</p>
Full article ">Figure 5
<p>The ESI-MS and MS/MS spectra of LSP3. The green lines are arrow marks from the software system.</p>
Full article ">Figure 6
<p>The <sup>13</sup>C-NMR spectra of LSP3 (<b>a</b>) and LMP3 (<b>b</b>), and the HSQC spectrum of LMP3 (<b>c</b>).</p>
Full article ">Figure 7
<p>Schematic representation of the structure of LSP3, LMP3 and CS extracted from the cartilage of bovine trachea (BT), chicken sternum (Ch) and skate (Sk). Two commercial preparations were also used: C4S, from bovine trachea, and C6S, from shark cartilage [<a href="#B15-marinedrugs-16-00170" class="html-bibr">15</a>].</p>
Full article ">Figure 8
<p>Antioxidant properties of polysaccharide sub-fractions from <span class="html-italic">Lapemis curtus</span>. (<b>a</b>) DPPH radical scavenging activities, (<b>b</b>) iron chelating effect, (<b>c</b>) 2,2′-azino-bis(3-ethylbenzothiazoline-6-sulfonic acid) (ABTS) radical scavenging activities of different samples. LSP1, LSP2 and LSP3, extracted from the skin of <span class="html-italic">Lapemis curtus.</span> LMP1, LMP2 and LMP3 extracted from the meat of <span class="html-italic">Lapemis curtus.</span> BHT, EDTA and Vc were used as positive controls.</p>
Full article ">Figure 8 Cont.
<p>Antioxidant properties of polysaccharide sub-fractions from <span class="html-italic">Lapemis curtus</span>. (<b>a</b>) DPPH radical scavenging activities, (<b>b</b>) iron chelating effect, (<b>c</b>) 2,2′-azino-bis(3-ethylbenzothiazoline-6-sulfonic acid) (ABTS) radical scavenging activities of different samples. LSP1, LSP2 and LSP3, extracted from the skin of <span class="html-italic">Lapemis curtus.</span> LMP1, LMP2 and LMP3 extracted from the meat of <span class="html-italic">Lapemis curtus.</span> BHT, EDTA and Vc were used as positive controls.</p>
Full article ">
15 pages, 1969 KiB  
Article
Dietary Polysaccharide from Enteromorpha Clathrata Modulates Gut Microbiota and Promotes the Growth of Akkermansia muciniphila, Bifidobacterium spp. and Lactobacillus spp.
by Qingsen Shang, Ya Wang, Lin Pan, Qingfeng Niu, Chao Li, Hao Jiang, Chao Cai, Jiejie Hao, Guoyun Li and Guangli Yu
Mar. Drugs 2018, 16(5), 167; https://doi.org/10.3390/md16050167 - 17 May 2018
Cited by 56 | Viewed by 7293
Abstract
Recently, accumulating evidence has suggested that Enteromorpha clathrata polysaccharide (ECP) could contribute to the treatment of diseases. However, as a promising candidate for marine drug development, although ECP has been extensively studied, less consideration has been given to exploring its effect on gut [...] Read more.
Recently, accumulating evidence has suggested that Enteromorpha clathrata polysaccharide (ECP) could contribute to the treatment of diseases. However, as a promising candidate for marine drug development, although ECP has been extensively studied, less consideration has been given to exploring its effect on gut microbiota. In this light, given the critical role of gut microbiota in health and disease, we investigated here the effect of ECP on gut microbiota using 16S rRNA high-throughput sequencing. As revealed by bioinformatic analyses, ECP considerably changed the structure of the gut microbiota and significantly promoted the growth of probiotic bacteria in C57BL/6J mice. However, interestingly, ECP exerted different effects on male and female microbiota. In females, ECP increased the abundances of Bifidobacterium spp. and Akkermansia muciniphila, a next-generation probiotic bacterium, whereas in males, ECP increased the population of Lactobacillus spp. Moreover, by shaping a more balanced structure of the microbiota, ECP remarkably reduced the antigen load from the gut in females. Altogether, our study demonstrates for the first time a prebiotic effect of ECP on gut microbiota and forms the basis for the development of ECP as a novel gut microbiota modulator for health promotion and disease management. Full article
Show Figures

Figure 1

Figure 1
<p>Administration of ECP significantly changed the structure of the gut microbiota. PCA score plot of the first and second components for the gut microbiome in male (<b>A</b>) and female mice (<b>B</b>). Blue indicates control groups, red indicates low dose groups and green indicates high dose groups. The PCA score plot was constructed based on the OTUs of the microbiota in male mice and female mice.</p>
Full article ">Figure 2
<p>Dietary ECP increased the richness and diversity of the gut microbiota. Chao1 (<b>A</b>) and observed species (<b>B</b>) were used as the richness estimators. Shannon indices (<b>C</b>) was used as the diversity estimator. a: <span class="html-italic">p</span> &lt; 0.05 vs. MN group; b: <span class="html-italic">p</span> &lt; 0.05 vs. FN group.</p>
Full article ">Figure 3
<p>Response of the gut microbiota to ECP treatment at the phylum (<b>A</b>) and genus (<b>B</b>) levels. The relative abundances of the gut bacteria presented here were calculated by averaging the data obtained from the six replicates within each group.</p>
Full article ">Figure 4
<p>The taxonomic cladogram obtained from LEfSe analysis of gut microbiota in different groups. The microbial compositions of the male and female mice were compared at different evolutionary levels. A significant value of less than 0.05 was used as a threshold for the LEfSe analysis.</p>
Full article ">Figure 5
<p>The LDA score obtained from LEfSe analysis of gut microbiota in different groups. A LDA effect size of more than 2 was used as a threshold for the LEfSe analysis.</p>
Full article ">Figure 6
<p>Intake of ECP significantly increased the abundances of <span class="html-italic">Bifidobacterium</span> spp. (<b>A</b>), <span class="html-italic">Lactobacillus</span> spp. (<b>B</b>) and <span class="html-italic">A. muciniphila</span> (<b>C</b>) in C57BL/6J mice. The solid line represents the average abundances of <span class="html-italic">Bifidobacterium</span> spp. (<b>A</b>), <span class="html-italic">Lactobacillus</span> spp. (<b>B</b>) and <span class="html-italic">A. muciniphila</span> (<b>C</b>) of the six replicates within each group; the dash line represents the median abundances of <span class="html-italic">Bifidobacterium</span> spp. (<b>A</b>), <span class="html-italic">Lactobacillus</span> spp. (<b>B</b>) and <span class="html-italic">A. muciniphila</span> (<b>C</b>) of the six replicates within each group. The differences in the abundances of the probiotic bacteria between treated mice and control mice were evidenced to be significant at <span class="html-italic">p</span> &lt; 0.05 by LEfSe analysis.</p>
Full article ">Figure 7
<p>ECP treatment reduced body weight (<b>A</b>,<b>B</b>), energy intake (<b>C</b>,<b>D</b>) and serum LBP levels (<b>E</b>) in C57BL/6J mice. * <span class="html-italic">p</span> &lt; 0.05 vs. the control group.</p>
Full article ">

2017

Jump to: 2024, 2022, 2021, 2020, 2019, 2018, 2016, 2015, 2014, 2013, 2012, 2011, 2010

2153 KiB  
Article
The Anti-Inflammatory Effect and Structure of EPCP1-2 from Crypthecodinium cohnii via Modulation of TLR4-NF-κB Pathways in LPS-Induced RAW 264.7 Cells
by Xiaolei Ma, Baolong Xie, Jin Du, Aijun Zhang, Jianan Hao, Shuxun Wang, Jing Wang and Junrui Cao
Mar. Drugs 2017, 15(12), 376; https://doi.org/10.3390/md15120376 - 1 Dec 2017
Cited by 14 | Viewed by 5156
Abstract
Exopolysaccharide from Crypthecodinium cohnii (EPCP1-2) is a marine exopolysaccharide that evidences a variety of biological activities. We isolated a neutral polysaccharide from the fermentation liquid of Crypthecodinium cohnii (CP). In this study, a polysaccharide that is derived from Crypthecodinium cohnii were analyzed and [...] Read more.
Exopolysaccharide from Crypthecodinium cohnii (EPCP1-2) is a marine exopolysaccharide that evidences a variety of biological activities. We isolated a neutral polysaccharide from the fermentation liquid of Crypthecodinium cohnii (CP). In this study, a polysaccharide that is derived from Crypthecodinium cohnii were analyzed and its anti-inflammatory effect was evaluated on protein expression of toll-like receptor 4 and nuclear factor κB pathways in macrophages. The structural characteristics of EPCP1-2 were characterized by GC (gas chromatography) and GC-MS (gas Chromatography-Mass Spectrometer) analyses. The molecular weight was about 82.5 kDa. The main chain of EPCP1-2 consisted of (1→6)-linked mannopyranosyl, (1→6)-linked glucopyranosyl, branched-chain consisted of (1→3,6)-linked galactopyranosyl and terminal consisted of t-l-Rhapyranosyl. The in vitro anti-inflammatory activity was representated through assay of proliferation rate, pro-inflammatory factor (NO) and expressions of proteins on RAW 264.7, the macrophage cell line. The results revealed that EPCP1-2 exhibited significant anti-inflammatory activity by regulating the expression of toll-like receptor 4, mitogen-activated protein kinases, and Nuclear Factor-κB protein. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Stepwise elution curve of crude <span class="html-italic">Crypthecodinium cohnii</span> (EPCP) on size-exclusion chromatography column of Superose 6B (10/300) and (<b>b</b>) elution curve of polysaccharide fractions from Superose 6B on anion-exchange chromatography column of HiTrap Capto Q (16/25).</p>
Full article ">Figure 2
<p>Anti-inflammatory effect of EPCP1 (EPCP1-1, EPCP1-2, EPCP1-3, EPCP1-4) on RAW 264.7 induced by LPS. (Results are expressed as the mean ± S.D. of three separate experiments). Statistical significance was tested using a Student’s <span class="html-italic">t</span>-test. EPCP group ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 3
<p>Cytotoxicity assay of EPCP1-2 (50, 100, 200, 400, 800 μg/mL) on RAW 264.7. (Results are expressed as the mean ± S.D. of three separate experiments). Statistical significance was tested using a Student’s <span class="html-italic">t</span>-test.</p>
Full article ">Figure 4
<p>High Performance Liquid Chromatography (HPLC) analysis of EPCP1-2.</p>
Full article ">Figure 5
<p>Total ion gas chromatogram of methylation patterns of EPCP1-2 using GC-MS. Note: signals a–d represent 2,4-Me2-Gala<span class="html-italic">p</span>, 2,3,4-Me3-Man<span class="html-italic">p</span>, 2,3,4-Me3-Glc<span class="html-italic">p</span>, respectively. The <span class="html-italic">x</span>-axis represents retention time in minutes; the <span class="html-italic">y</span>-axis represents signal intensity by counts.</p>
Full article ">Figure 6
<p>Effects of EPCP1-2 on pro-inflammatory factor (NO) production in RAW 264.7. (Results are expressed as the mean ± S.D. of three separate experiments). Statistical significance was tested using a Student’s <span class="html-italic">t</span>-test. EPCP1-2 group ** <span class="html-italic">p</span> &lt; 0.01; LPS group (positive control) <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 7
<p>Inhibitory effect of EPCP1-2 isolated from <span class="html-italic">Crypthecodinium cohnii</span> on the NF-κB protein activation in LPS-induced RAW 264.7 cells. Cells were stimulated by LPS (5 μg/mL) for 1 h with or without the presence of EPCP1-2 (200, 400, 800 μg/mL). The protein levels of IκB-α and NF-κB were determined using ECL (electrochemiluminescence) immunoblotting method. (<b>a</b>) TLR4 protein and TAK1 protein expression were examined by Western blotting analysis. (<b>b</b>) Densitometric analysis showed the effects of EPCP1-2 on LPS-induced expressions of TLR4 and TAK1 proteins. (<b>c</b>) NF-κB p65 signal relative proteins expression was examined by Western blotting analysis. (<b>d</b>) Densitometric analysis showed the effects of EPCP1-2 on LPS-induced expressions of NF-κB p65 signal relative proteins. (<b>e</b>) Densitometric analysis showed the effects of EPCP1-2 on LPS-induced expressions of IKKα/β, IκB-α and NF-κB proteins. (<b>f</b>) Densitometric analysis showed the effects of EPCP1-2 on LPS-induced expressions of ERK1/2, JNK/SAPK, p38 MAPK proteins. (Results are expressed as the mean ± S.D. of three separate experiments). Statistical significance was tested using a Student’s <span class="html-italic">t</span>-test. EPCP1-2 group ** <span class="html-italic">p</span> &lt; 0.01; LPS group (positive control) <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 8
<p>Schematic diagram of the propose target for the anti-inflammatory effects of EPCP1-2 in the RAW 264.7 cells stimulated by LPS, potentially leading to the inhibition of the pro-inflammatory cytokines and related mediators.</p>
Full article ">
5977 KiB  
Article
Degradation of Polysaccharides from Grateloupia filicina and Their Antiviral Activity to Avian Leucosis Virus Subgroup J
by Yuhao Sun, Xiaolin Chen, Ziqiang Cheng, Song Liu, Huahua Yu, Xueqin Wang and Pengcheng Li
Mar. Drugs 2017, 15(11), 345; https://doi.org/10.3390/md15110345 - 3 Nov 2017
Cited by 22 | Viewed by 4493
Abstract
In this study, polysaccharides from Grateloupia filicinia (GFP) were extracted and several low molecular weight (Mw) G. filicina polysaccharides (LGFPs) were prepared by the hydrogen peroxide (H2O2) oxidation method. Additionally, the effect of different experimental conditions on the degradation [...] Read more.
In this study, polysaccharides from Grateloupia filicinia (GFP) were extracted and several low molecular weight (Mw) G. filicina polysaccharides (LGFPs) were prepared by the hydrogen peroxide (H2O2) oxidation method. Additionally, the effect of different experimental conditions on the degradation of GFP was determined. Results showed that the GFP degradation rate was positively related to H2O2 concentration and temperature, and negatively related to pH. Chemical analysis and Fourier transform infrared spectra (FT-IR) of GFP and LGFPs showed that the degradation caused a slight decrease of total sugar and sulfate content. However, there was no obvious change for monosaccharide contents. Then, the anti-ALV-J activity of GFP and LGFPs were determined in vitro. Results revealed that all of the samples could significantly inhibit ALV-J and lower Mw LGFPs exhibited a stronger suppression, and that the fraction LGFP-3 with Mw 8.7 kDa had the best effect. In addition, the reaction phase assays showed that the inhibition effect was mainly because of the blocking virus adsorption to host cells. Moreover, real-time PCR, western-blot, and IFA were further applied to evaluate the blocking effects of LGFP-3. Results showed that the gene relative expression and gp85 protein for LGFPS-3 groups were all reduced. Data from IFA showed that there was less virus infected cells for 1000 and 200 μg/mL LGFPS-3 groups when compared to virus control. Therefore, lower Mw polysaccharides from G. filicina might supply a good choice for ALV-J prevention and treatment. Full article
Show Figures

Figure 1

Figure 1
<p>Effects of temperature on the molecular weight (Mw) of <span class="html-italic">Grateloupia filicinia</span> polysaccharide (GFP). The reaction was carried out in pH 1 and 0.15% H<sub>2</sub>O<sub>2</sub> under different temperature conditions.</p>
Full article ">Figure 2
<p>Effect of H<sub>2</sub>O<sub>2</sub> concentration on the Mw of GFP. GFP was degraded at pH 1 and 90 °C under different H<sub>2</sub>O<sub>2</sub> concentration conditions.</p>
Full article ">Figure 3
<p>Effects of pH on the Mw of GFP. GFP was degraded in 0.3% H<sub>2</sub>O<sub>2</sub> at 90 °C under different pH conditions.</p>
Full article ">Figure 4
<p>HPLC profiles of the GFP and LGFPs.</p>
Full article ">Figure 5
<p>Fourier transform infrared (FT-IR) spectra of GFP and LGFPs in regions from 4000 to 500 cm<sup>−1</sup>.</p>
Full article ">Figure 6
<p>Expression of ALV-J p27 antigen. Antiviral activity of GFP and LGFPs were determined by ALV p27 antigen test kit. Results are recognized as positive when <span class="html-italic">S</span>/<span class="html-italic">P</span> value is greater than 0.2. Data are shown as the Mean + SD. Values with different letters in the same column (a–f) are significantly different (<span class="html-italic">p</span> &lt; 0.05) from each other.</p>
Full article ">Figure 7
<p>Expression of p27 after different modes of administration: LGFP-3 treated DF-1 cells before adsorption (BA); LGFP-3 treated virus at the adsorption phase (Ad); and, LGFP-3 treated DF-1 cells after adsorption (AA). The final concentrations of LGFP-3 were all 1 mg/mL in these three administration. DF-1 cells with and without inoculation were used as the virus and cell control, respectively. Data are shown as the Mean + SD. Values with different letters in the same column (a–c) are significantly different (<span class="html-italic">p</span> &lt; 0.05) from each other.</p>
Full article ">Figure 8
<p>Gene relative expression of ALV-J measured with real-time PCR. DF-1 cells with or without inoculation were used as the virus and cell control, respectively. Data are shown as the Mean + SD. Values with different letters in the same column (a–d) are significantly different (<span class="html-italic">p</span> &lt; 0.05) from each other.</p>
Full article ">Figure 9
<p>Expression of ALV-J gp85 protein evaluated by western-blot. DF-1 cells with or without inoculation were used as the virus and cell control, respectively.</p>
Full article ">Figure 10
<p>Expression of ALV-J gp85 protein evaluated by IFA. (<b>a</b>): virus control; (<b>b</b>): cell control; (<b>c</b>): ALV-J treated with 40 μg/mL LGFP-3; (<b>d</b>): ALV-J treated with 200 μg/mL LGFP-3; and, (<b>e</b>): ALV-J treated with 1000 μg/mL LGFP-3.</p>
Full article ">
1755 KiB  
Article
The Identification of a SIRT6 Activator from Brown Algae Fucus distichus
by Minna K. Rahnasto-Rilla, Padraig McLoughlin, Tomasz Kulikowicz, Maire Doyle, Vilhelm A. Bohr, Maija Lahtela-Kakkonen, Luigi Ferrucci, Maria Hayes and Ruin Moaddel
Mar. Drugs 2017, 15(6), 190; https://doi.org/10.3390/md15060190 - 21 Jun 2017
Cited by 37 | Viewed by 8183
Abstract
Brown seaweeds contain many bioactive compounds, including polyphenols, polysaccharides, fucosterol, and fucoxantin. These compounds have several biological activities, including anti-inflammatory, hepatoprotective, anti-tumor, anti-hypertensive, and anti-diabetic activity, although in most cases their mechanisms of action are not understood. In this study, extracts generated from [...] Read more.
Brown seaweeds contain many bioactive compounds, including polyphenols, polysaccharides, fucosterol, and fucoxantin. These compounds have several biological activities, including anti-inflammatory, hepatoprotective, anti-tumor, anti-hypertensive, and anti-diabetic activity, although in most cases their mechanisms of action are not understood. In this study, extracts generated from five brown algae (Fucus dichitus, Fucus vesiculosus (Linnaeus), Cytoseira tamariscofolia, Cytoseira nodacaulis, Alaria esculenta) were tested for their ability to activate SIRT6 resulting in H3K9 deacetylation. Three of the five macroalgal extracts caused a significant increase of H3K9 deacetylation, and the effect was most pronounced for F. dichitus. The compound responsible for this in vitro activity was identified by mass spectrometry as fucoidan. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>SIRT6 deacetylation activity in the presence of five species of brown algae extracts. The change of SIRT6 deacetylation in the presence of 0.5 mg/mL (grey) and 1.0 mg/mL (light grey) extracts is compared to controls with 500 µM NAD<sup>+</sup> and 40 µM H3K9Ac with 30 min of incubation time. The data are presented as means ± SD, <span class="html-italic">n</span> = 3.</p>
Full article ">Figure 2
<p>HPLC chromatogram of <span class="html-italic">F. distichus</span> and its separation into eight fractions using Zorbax Eclipse XDB-C18 column (4.6 mm × 50 mm, 1.8 µm). The collected fractions: F1 = 0.4–0.5 min; F7 = 5.4–5.5 min.</p>
Full article ">Figure 3
<p>SIRT6 deacetylation activity (fold activity relative to control) of fractions (F1–F8) from <span class="html-italic">F. distichus</span> in the presence of 0.5 mg/mL (grey) and 1.0 mg/mL (light grey) fractions (F1–F8) with 500 µM NAD<sup>+</sup> and 40 µM H3K9Ac with 30 min of incubation time. The data are presented as means ± SD, <span class="html-italic">n</span> = 3.</p>
Full article ">Figure 4
<p>HPLC–MS analysis of subfraction F1 in negative ionization mode with a scan range of <span class="html-italic">m</span>/<span class="html-italic">z</span> 150–600.</p>
Full article ">Figure 5
<p>Reported structural elements for fucoidan isolated from the brown seaweeds (<b>A</b>) <span class="html-italic">F. distichus</span>, (<b>B</b>) <span class="html-italic">F. vesiculosus</span>. Modified from [<a href="#B24-marinedrugs-15-00190" class="html-bibr">24</a>,<a href="#B25-marinedrugs-15-00190" class="html-bibr">25</a>].</p>
Full article ">Figure 6
<p>(<b>A</b>) Dose-response curve of the extract sub-fraction of F1 (●) in the presence of 500 µM NAD<sup>+</sup> and 40 µM H3K9Ac with 30 min of incubation time. The data are presented as means ± SD, <span class="html-italic">n</span> = 3; (<b>B</b>) Dose-response curve of fucoidan (■) on SIRT6 deacetylation activity in the presence of 500 µM NAD<sup>+</sup> and 40 µM H3K9Ac with 30 min of incubation time. The data are presented as means ± SD, <span class="html-italic">n</span> = 3.</p>
Full article ">Figure 7
<p>Western blot method for the in vitro SIRT6 deacetylation assay. Serially diluted concentrations of a SIRT6 stimulator (1–16 µg/mL) were incubated for 30 min at 37 °C in the presence of 1 μg/well of a purified recombinant GST-SIRT6 protein, 2 μg purified whole chicken core histones with 500 μM NAD<sup>+</sup> in 25 mM Tris-HCl, pH 8.0. (<b>A</b>) Acetylation level was detected with anti-H3K9Ac antibody and normalized to total H3 histone. Values indicate final fucoidan concentration in µg/mL. Molecular weight markers in kDa. (<b>B</b>) Quantification of H3K9 deacetylation. Values represent the averages of three experiments; error bars indicate standard deviation.</p>
Full article ">
2924 KiB  
Article
Preparation, Characterization and Properties of Alginate/Poly(γ-glutamic acid) Composite Microparticles
by Zongrui Tong, Yu Chen, Yang Liu, Li Tong, Jiamian Chu, Kecen Xiao, Zhiyu Zhou, Wenbo Dong and Xingwu Chu
Mar. Drugs 2017, 15(4), 91; https://doi.org/10.3390/md15040091 - 11 Apr 2017
Cited by 71 | Viewed by 6524
Abstract
Alginate (Alg) is a renewable polymer with excellent hemostatic properties and biocapability and is widely used for hemostatic wound dressing. However, the swelling properties of alginate-based wound dressings need to be promoted to meet the requirements of wider application. Poly(γ-glutamic acid) [...] Read more.
Alginate (Alg) is a renewable polymer with excellent hemostatic properties and biocapability and is widely used for hemostatic wound dressing. However, the swelling properties of alginate-based wound dressings need to be promoted to meet the requirements of wider application. Poly(γ-glutamic acid) (PGA) is a natural polymer with high hydrophility. In the current study, novel Alg/PGA composite microparticles with double network structure were prepared by the emulsification/internal gelation method. It was found from the structure characterization that a double network structure was formed in the composite microparticles due to the ion chelation interaction between Ca2+ and the carboxylate groups of Alg and PGA and the electrostatic interaction between the secondary amine group of PGA and the carboxylate groups of Alg and PGA. The swelling behavior of the composite microparticles was significantly improved due to the high hydrophility of PGA. Influences of the preparing conditions on the swelling behavior of the composites were investigated. The porous microparticles could be formed while compositing of PGA. Thermal stability was studied by thermogravimetric analysis method. Moreover, in vitro cytocompatibility test of microparticles exhibited good biocompatibility with L929 cells. All results indicated that such Alg/PGA composite microparticles are a promising candidate in the field of wound dressing for hemostasis or rapid removal of exudates. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) FT-IR spectra of alginate, PGA and composite microparticles with various contents; (<b>b</b>) double-network structure scheme of the composite microparticle.</p>
Full article ">Figure 2
<p>XPS C1s spectra of alginate (<b>a</b>), Alg/PGA82 (composite microparticles whose mass ratio is m<sub>Alg</sub>:m<sub>PGA</sub> = 8:2) (<b>b</b>), Alg/PGA73 (composite microparticles whose mass ratio is m<sub>Alg</sub>:m<sub>PGA</sub> = 7:3) (<b>c</b>) and PGA (<b>d</b>).</p>
Full article ">Figure 3
<p>XRD patterns of Alginate, PGA and composite microparticles.</p>
Full article ">Figure 4
<p>Morphology of the Alg/PGA Composite Microparticles. Note: SEM images of various microparticles at 500, 2000 and 20,000 magnification scales (<b>a</b>–<b>e</b> belongs to alginate microparticles and Alg/PGA composite microparticles with mass ratio is m<sub>Alg</sub>:m<sub>PGA</sub> = 9:1, 8:2, 7:3, 6:4 respectively).</p>
Full article ">Figure 5
<p>Effect of feeding ratio (<b>a</b>), concentration of span 80 (<b>b</b>), W/O ratio (<b>c</b>) and oil phase (<b>d</b>) on the swelling behavior of Alg/PGA composite microparticles.</p>
Full article ">Figure 6
<p>(<b>a</b>) Cytotoxicity assay of leach liquors. (<b>b</b>) Micrographs of cells cultured in leach liquors of samples Alg (<b>b.1</b>), Alg/PGA82 (<b>b.2</b>), Alg/PGA64 (<b>b.3</b>) and control group (<b>b.4</b>). Note: Samples Alg, Alg/PGA82 and Alg/PGA64 represent alginate microparticles, composite microparticles prepared with m<sub>Alg</sub>:m<sub>PGA</sub> = 8:2 and composite microparticles prepared with m<sub>Alg</sub>:m<sub>PGA</sub> = 6:4 respectively.</p>
Full article ">Figure 6 Cont.
<p>(<b>a</b>) Cytotoxicity assay of leach liquors. (<b>b</b>) Micrographs of cells cultured in leach liquors of samples Alg (<b>b.1</b>), Alg/PGA82 (<b>b.2</b>), Alg/PGA64 (<b>b.3</b>) and control group (<b>b.4</b>). Note: Samples Alg, Alg/PGA82 and Alg/PGA64 represent alginate microparticles, composite microparticles prepared with m<sub>Alg</sub>:m<sub>PGA</sub> = 8:2 and composite microparticles prepared with m<sub>Alg</sub>:m<sub>PGA</sub> = 6:4 respectively.</p>
Full article ">
2601 KiB  
Article
Degradation of Marine Algae-Derived Carbohydrates by Bacteroidetes Isolated from Human Gut Microbiota
by Miaomiao Li, Qingsen Shang, Guangsheng Li, Xin Wang and Guangli Yu
Mar. Drugs 2017, 15(4), 92; https://doi.org/10.3390/md15040092 - 24 Mar 2017
Cited by 70 | Viewed by 7416
Abstract
Carrageenan, agarose, and alginate are algae-derived undigested polysaccharides that have been used as food additives for hundreds of years. Fermentation of dietary carbohydrates of our food in the lower gut of humans is a critical process for the function and integrity of both [...] Read more.
Carrageenan, agarose, and alginate are algae-derived undigested polysaccharides that have been used as food additives for hundreds of years. Fermentation of dietary carbohydrates of our food in the lower gut of humans is a critical process for the function and integrity of both the bacterial community and host cells. However, little is known about the fermentation of these three kinds of seaweed carbohydrates by human gut microbiota. Here, the degradation characteristics of carrageenan, agarose, alginate, and their oligosaccharides, by Bacteroides xylanisolvens, Bacteroides ovatus, and Bacteroides uniforms, isolated from human gut microbiota, are studied. Full article
Show Figures

Figure 1

Figure 1
<p>Diagram of fragments generated from electrospray ionization collision-induced-dissociation mass spectrometry (ESI-CID-MS<sup>2</sup>) of neocarrabiose and neocarradiitol. (<b>a</b>) neocarrabiose; (<b>b</b>) neocarradiitol.</p>
Full article ">Figure 2
<p>Electrospray ionization collision-induced-dissociation mass spectrometry (ESI-CID-MS<sup>2</sup>) spectrums of products of K-DP17 degraded by 38F6. (<b>a</b>) ESI-CID-MS<sup>2</sup> spectrum of disaccharides generated from K-DP17 degradation by 38F6, (<b>b</b>) ESI-CID-MS<sup>2</sup> spectrums of reduced disaccharides generated from K-DP17 degradation by 38F6, (<b>c</b>) ESI-CID-MS<sup>2</sup> spectrums of tetrasaccharides generated from K-DP17 degradation by 38F6, (<b>d</b>) ESI-CID-MS<sup>2</sup> spectrums of reduced tetrasaccharides generated from K-DP17 degradation by 38F6.</p>
Full article ">Figure 2 Cont.
<p>Electrospray ionization collision-induced-dissociation mass spectrometry (ESI-CID-MS<sup>2</sup>) spectrums of products of K-DP17 degraded by 38F6. (<b>a</b>) ESI-CID-MS<sup>2</sup> spectrum of disaccharides generated from K-DP17 degradation by 38F6, (<b>b</b>) ESI-CID-MS<sup>2</sup> spectrums of reduced disaccharides generated from K-DP17 degradation by 38F6, (<b>c</b>) ESI-CID-MS<sup>2</sup> spectrums of tetrasaccharides generated from K-DP17 degradation by 38F6, (<b>d</b>) ESI-CID-MS<sup>2</sup> spectrums of reduced tetrasaccharides generated from K-DP17 degradation by 38F6.</p>
Full article ">Figure 3
<p>Thin-layer chromatography (TLC) analysis of degradation of neocarratetraose (NK-DP4) and neocarrahexaose (NK-DP6) by 38F6; 2: neocarrabiose (NK-DP2), 4: NK-DP4, 6: NK-DP6, 4J: NK-DP4 fermentation by 38F6, 6J: NK-DP6 fermentation by 38F6.</p>
Full article ">Figure 4
<p>Analysis of the products of agarose (AP) degraded by <span class="html-italic">B. uniformis</span> L8 (<b>a</b>) TLC analysis of degradation of AP by <span class="html-italic">B. uniformis</span> L8 at 0, 48, 96, 144 and 192 h; (<b>b</b>) HPLC chromatography of final products of AP degraded by <span class="html-italic">B. uniformis</span> L8 at 96 h; 1. monosaccharide standard (Man: mannose, Rha: rhamnose, GalA: galacturonic acid, Glc: glucose, Gal: Galactose, Xyl: xylose); 2. final products of AP degraded by <span class="html-italic">B. uniformis</span> L8 at 96 h.</p>
Full article ">Figure 5
<p>NMR <sup>13</sup>C spectrum of the oligosaccharides generated from agarose (AP) degradation by <span class="html-italic">B. uniformis</span> L8 at 48 h.</p>
Full article ">Figure 6
<p>Profile of degradation position and mode of action of enzymes on marine carbohydrates. (<b>a</b>) degradation position of carrageenan oligosaccharides by enzymes secreted by 38F6 (<span class="html-italic">B. xylanisolvens</span> and <span class="html-italic">E. coli</span>); (<b>b</b>) degradation position of agarose (AP) by enzymes from <span class="html-italic">B. uniformis</span> L8; (<b>c</b>) degradation position of alginate by enzymes from <span class="html-italic">B. ovatus</span> G19.</p>
Full article ">
2000 KiB  
Article
Immunomodulatory and Anti-IBDV Activities of the Polysaccharide AEX from Coccomyxa gloeobotrydiformis
by Qiang Guo, Qiang Shao, Wenping Xu, Lei Rui, Ryo Sumi, Fumio Eguchi and Zandong Li
Mar. Drugs 2017, 15(2), 36; https://doi.org/10.3390/md15020036 - 10 Feb 2017
Cited by 15 | Viewed by 5938
Abstract
A number of polysaccharides have been reported to show immunomodulatory and antiviral activities against various animal viruses. AEX is a polysaccharide extracted from the green algae, Coccomyxa gloeobotrydiformis. The aim of this study was to examine the function of AEX in regulating [...] Read more.
A number of polysaccharides have been reported to show immunomodulatory and antiviral activities against various animal viruses. AEX is a polysaccharide extracted from the green algae, Coccomyxa gloeobotrydiformis. The aim of this study was to examine the function of AEX in regulating the immune response in chickens and its capacity to inhibit the infectious bursal disease virus (IBDV), to gain an understanding of its immunomodulatory and antiviral ability. Here, preliminary immunological tests in vitro showed that the polysaccharide AEX can activate the chicken peripheral blood molecular cells’ (PBMCs) response by inducing the production of cytokines and NO, promote extracellular antigen presentation but negatively regulate intracellular antigen presentation in chicken splenic lymphocytes, and promote the proliferation of splenic lymphocytes and DT40 cells. An antiviral analysis showed that AEX repressed IBDV replication by the deactivation of viral particles or by interfering with adsorption in vitro and reduced the IBDV viral titer in the chicken bursa of Fabricius. Finally, in this study, when AEX was used as an adjuvant for the IBDV vaccine, specific anti-IBDV antibody (IgY, IgM, and IgA) titers were significantly decreased. These results indicate that the polysaccharide AEX may be a potential alternative approach for anti-IBDV therapy and an immunomodulator for the poultry industry. However, more experimentation is needed to find suitable conditions for it to be used as an adjuvant for the IBDV vaccine. Full article
Show Figures

Figure 1

Figure 1
<p>Multiple cytokines and iNOS expression were upregulated and NO production were increased by AEX in PBMCs. (<b>A</b>–<b>G</b>) PBMCs were isolated and cultured with AEX (0–100 μg/mL) or LPS (100 ng/mL) for 24 h. Then total RNA was extracted and analyzed by qRT-PCR for IFN-β, IL-1β, IL-6, TNF-α, IL-10, IL-12p40, and iNOS; (<b>H</b>) PBMCs were isolated and cultured with AEX (0–100 μg/mL) or LPS (100 ng/mL) for 24 h and 48 h. Then, the culture supernatants were collected and nitrite contents were determined by Griess reaction. Data represent means ± SEM from three wells per group. * <span class="html-italic">p</span> ≤ 0.05; ** <span class="html-italic">p</span> ≤ 0.01; *** <span class="html-italic">p</span> ≤ 0.001. Results are representative of two independent experiments.</p>
Full article ">Figure 2
<p>Inflammatory cytokines and surface molecules on splenic lymphocytes were regulated by AEX. Splenic lymphocytes were isolated and cultured with AEX (0–100 μg/mL) or LPS (100 ng/mL) for 24 h. Then, total RNA was extracted and analyzed by qRT-PCR for IL-1β (<b>A</b>), IL-6 (<b>B</b>), MHC I (<b>C</b>), MHC II (<b>D</b>), CD3ε (<b>E</b>), CD4 (<b>F</b>), and CD8α (<b>G</b>). Data represent means ± SEM from three wells per group. * <span class="html-italic">p</span> ≤ 0.05; ** <span class="html-italic">p</span> ≤ 0.01. Results are representative of two independent experiments.</p>
Full article ">Figure 3
<p>AEX promoted DT40 and splenic lymphocytes proliferation, but reduced peripheral blood lymphocyte proliferation in vitro. (<b>A</b>–<b>C</b>) DT40 cells, splenic lymphocytes and peripheral blood lymphocytes were cultured in 96-well plates. After being stimulated by AEX (0–250 μg/mL) or Con A (40 μg/mL) for 24 h, 48 h, and 72 h, respectively, the proliferation was examined by MTT method as described in the Materials and Methods section; (<b>D</b>,<b>E</b>) Splenic lymphocytes and peripheral blood lymphocytes were isolated and cultured with AEX (0–100 μg/mL) or LPS (100 ng/mL) for 24 h, respectively. Then, total RNA was extracted and analyzed by qRT-PCR for IL-2. Data represent means ± SEM from three wells per group. * <span class="html-italic">p</span> ≤ 0.05; ** <span class="html-italic">p</span> ≤ 0.01; *** <span class="html-italic">p</span> ≤ 0.001. Results are representative of two independent experiments.</p>
Full article ">Figure 4
<p>AEX inhibited IBDV replication by deactivating viral particle and interfering with adsorption. (<b>A</b>) The Vero cells were treated with different concentrations of AEX for 24 h and cell viability was measured by MTT assay; (<b>B</b>) The Vero cells infected with IBDV (MOI = 1.0) were treated with different concentrations of AEX, and viral titers in the supernatant were identified by TCID<sub>50</sub> assays on the Vero cells. PBS treatment wells served as the control; (<b>C</b>,<b>D</b>) The Vero cells infected with IBDV (MOI = 1.0) were treated with different concentrations of AEX, and immunofluorescence staining was performed using anti-IBDV antibodies. (<b>C</b>) Scale bar represents 100 μm; (<b>D</b>) IBDV positive cells number was calculated; (<b>E</b>) The Vero cells were incubated with IBDV (MOI = 1.0) for 1 h and treated with AEX at the indicated time points (0, 2, 4, 24 h p.i.). Viral titers in the supernatant were determined by TCID<sub>50</sub> assays on the Vero cells at 48 h p.i. (<b>F</b>) The Vero cells were infected with IBDV (MOI = 1.0) via different polysaccharide treatment method (IBDV pretreatment, adsorption, and after adsorption) and viral titers in the supernatant were determined by TCID<sub>50</sub> assays on the Vero cells at 48 h p.i. Data represent means ± SEM from three wells per group. * <span class="html-italic">p</span> ≤ 0.05; ** <span class="html-italic">p</span> ≤ 0.01. Results are representative of two independent experiments.</p>
Full article ">Figure 5
<p>AEX inhibited IBDV replication in chicken bursa of Fabricius. SPF chickens (6 chickens per group) were oral inoculated with different concentrations of AEX (0, 62.5, 125, 250 and 500 mg/kg body weight) per day. After three days, chickens were eye-dropped with IBDV (10<sup>5.5</sup>TCID<sub>50</sub>). Bursa of Fabricius were collected at 5 d.p.i. and divided into two parts. (<b>A</b>) Half of the bursa of Fabricius were grinded and virus titers were determined by TCID<sub>50</sub>; (<b>B</b>) The other half bursa of Fabricius were subjected to isolate total RNA and the expression lever of IBDV <span class="html-italic">VP2</span> was evaluated by qRT-PCR. Data represent means ± SEM from five wells per group. * <span class="html-italic">p</span> ≤ 0.05; ** <span class="html-italic">p</span> ≤ 0.01. Results are representative of two independent experiments.</p>
Full article ">Figure 6
<p>AEX reduced specific anti-IBDV antibody (IgY, IgM, and IgA) titers in the sera of chickens immunized with IBDV. SPF chickens (6 chickens per group) were immunized with IBDV vaccine by subcutaneous injection in the absence or presence of AEX (PBS used as control) thrice at a 14-day interval. The sera were collected (14 days after the first immunization, 7 days and 14 days after secondary immunization, and 7 days after tertiary immunization) and specific antibody (<b>A</b>) IgY; (<b>B</b>) IgM; (<b>C</b>) IgA were detected by ELISA. Data represent means ± SEM from five wells per group. <sup>a,b,c,d</sup> <span class="html-italic">p</span> ≤ 0.05 vs. <span class="html-italic">FA</span>. Results are representative of two independent experiments.</p>
Full article ">

2016

Jump to: 2024, 2022, 2021, 2020, 2019, 2018, 2017, 2015, 2014, 2013, 2012, 2011, 2010

1834 KiB  
Article
Purification and Characterization of a New Alginate Lyase from Marine Bacterium Vibrio sp. SY08
by Shangyong Li, Linna Wang, Jianhua Hao, Mengxin Xing, Jingjing Sun and Mi Sun
Mar. Drugs 2017, 15(1), 1; https://doi.org/10.3390/md15010001 - 23 Dec 2016
Cited by 59 | Viewed by 6815
Abstract
Unsaturated alginate disaccharides (UADs), enzymatically derived from the degradation of alginate polymers, are considered powerful antioxidants. In this study, a new high UAD-producing alginate lyase, AlySY08, has been purified from the marine bacterium Vibrio sp. SY08. AlySY08, with a molecular weight of about [...] Read more.
Unsaturated alginate disaccharides (UADs), enzymatically derived from the degradation of alginate polymers, are considered powerful antioxidants. In this study, a new high UAD-producing alginate lyase, AlySY08, has been purified from the marine bacterium Vibrio sp. SY08. AlySY08, with a molecular weight of about 33 kDa and a specific activity of 1070.2 U/mg, showed the highest activity at 40 °C in phosphate buffer at pH 7.6. The enzyme was stable over a broad pH range (6.0–9.0) and retained about 75% activity after incubation at 40 °C for 2 h. Moreover, the enzyme was active in the absence of salt ions and its activity was enhanced by the addition of NaCl and KCl. AlySY08 resulted in an endo-type alginate lyase that degrades both polyM and polyG blocks, yielding UADs as the main product (81.4% of total products). All these features made AlySY08 a promising candidate for industrial applications in the production of antioxidants from alginate polysaccharides. Full article
Show Figures

Figure 1

Figure 1
<p>Phylogenetic tree of strain SY08 and related bacteria. The tree ID is based on a maximum parsimony analysis of the 16S rDNA sequences. The obtained 16S rDNA sequence was searched for and aligned by using the BLASTn and ClustalX programs, respectively. The phylogenetic tree was obtained by using MEGA 4.0 software.</p>
Full article ">Figure 2
<p>SDS-PAGE analysis of purified AlySY08. The purified AlySY08 was resolved by 10% acrylamide (<span class="html-italic">w</span>/<span class="html-italic">v</span>) SDS-PAGE followed by staining with Coomassie Blue G-250. Lane M, molecular weight markers; Lane 1, purified AlySY08.</p>
Full article ">Figure 3
<p>Effects of pH and temperature on the activity and stability of AlySY08. (<b>a</b>) The optimum pH for AlySY08 was determined by measuring its activity at 40 °C in 50 mM Na<sub>2</sub>HPO<sub>4</sub>-citric acid buffer (filled circle), 50 mM Na<sub>2</sub>HPO<sub>4</sub>-NaH<sub>2</sub>PO<sub>4</sub> buffer (open circle), 50 mM Tris-HCl buffer (filled triangle) and 50 mM Gly-NaOH buffer (filled square); (<b>b</b>) The optimal temperature for AlySY08 was determined by measuring its activity at various temperatures (10–60 °C); (<b>c</b>) pH stability of AlySY08. The residual activity was measured at 40 °C in 50 mM phosphate buffer (pH 7.6) after incubation in the buffers reported above at 4 °C for 6 h; (<b>d</b>) Thermostability of AlySY08. The enzyme was incubated at 30 °C (filled circle), 40 °C (open rhombus), 45 °C (open square) and 50 °C (filled triangle) for various times. The residual activity was then determined at 40 °C. The activity of control (100% relative activity) is 12.6 U/mL.</p>
Full article ">Figure 4
<p>Effect of NaCl on enzymatic activity of AlySY08. The activity of AlySY08 in the absence of NaCl was retained at 100%. All the experiments were conducted in triplicate.</p>
Full article ">Figure 5
<p>Size-exclusion chromatography of the alginate degradation products by AlySY08. The elution volumes of the dimer (DP2), trimer (DP3), tetramer (DP4), and pentamer (DP5) are 16.1 mL, 14.9 mL, 14.1 mL and 13.7 mL, respectively. The ratios of dimers present in the degradation products were analyzed by the peak integration function on the UNICORN 5.31 software (GE Healthcare, Madison, WI, USA).</p>
Full article ">Figure 6
<p>TLC and ESI-MS analysis of the main products of AlySY08. (<b>a</b>) TLC analysis. The reaction products were separated on a HPTLC plate with <span class="html-italic">n</span>-butanol/formic acid/water (2:1:1, by vol) and visualized with a diphenylamine/aniline/phosphate reagent. Lane M: standard UAOs mixture, disaccharide (DP2) and trisaccharide (DP3); Lane 0 sodium alginate; Lane 1 reaction products; (<b>b</b>) ESI-MS analysis of the main products by AlySY08.</p>
Full article ">
2673 KiB  
Article
Identification of a Pro-Angiogenic Potential and Cellular Uptake Mechanism of a LMW Highly Sulfated Fraction of Fucoidan from Ascophyllum nodosum
by Nicolas Marinval, Pierre Saboural, Oualid Haddad, Murielle Maire, Kevin Bassand, Frederic Geinguenaud, Nadia Djaker, Khadija Ben Akrout, Marc Lamy de la Chapelle, Romain Robert, Olivier Oudar, Erwan Guyot, Christelle Laguillier-Morizot, Angela Sutton, Cedric Chauvierre, Frederic Chaubet, Nathalie Charnaux and Hanna Hlawaty
Mar. Drugs 2016, 14(10), 185; https://doi.org/10.3390/md14100185 - 17 Oct 2016
Cited by 36 | Viewed by 7433
Abstract
Herein we investigate the structure/function relationships of fucoidans from Ascophyllum nodosum to analyze their pro-angiogenic effect and cellular uptake in native and glycosaminoglycan-free (GAG-free) human endothelial cells (HUVECs). Fucoidans are marine sulfated polysaccharides, which act as glycosaminoglycans mimetics. We hypothesized that the size [...] Read more.
Herein we investigate the structure/function relationships of fucoidans from Ascophyllum nodosum to analyze their pro-angiogenic effect and cellular uptake in native and glycosaminoglycan-free (GAG-free) human endothelial cells (HUVECs). Fucoidans are marine sulfated polysaccharides, which act as glycosaminoglycans mimetics. We hypothesized that the size and sulfation rate of fucoidans influence their ability to induce pro-angiogenic processes independently of GAGs. We collected two fractions of fucoidans, Low and Medium Molecular Weight Fucoidan (LMWF and MMWF, respectively) by size exclusion chromatography and characterized their composition (sulfate, fucose and uronic acid) by colorimetric measurement and Raman and FT-IR spectroscopy. The high affinities of fractionated fucoidans to heparin binding proteins were confirmed by Surface Plasmon Resonance. We evidenced that LMWF has a higher pro-angiogenic (2D-angiogenesis on Matrigel) and pro-migratory (Boyden chamber) potential on HUVECs, compared to MMWF. Interestingly, in a GAG-free HUVECs model, LMWF kept a pro-angiogenic potential. Finally, to evaluate the association of LMWF-induced biological effects and its cellular uptake, we analyzed by confocal microscopy the GAGs involvement in the internalization of a fluorescent LMWF. The fluorescent LMWF was mainly internalized through HUVEC clathrin-dependent endocytosis in which GAGs were partially involved. In conclusion, a better characterization of the relationships between the fucoidan structure and its pro-angiogenic potential in GAG-free endothelial cells was required to identify an adapted fucoidan to enhance vascular repair in ischemia. Full article
Show Figures

Figure 1

Figure 1
<p>Raman and Fourrier Tansform Infrared (FT-IR) Spectroscopy analysis. Fucoidan spectra are represented in black for crude Ascophyscient (ASPHY), dark grey for the low molecular weight fucoidan (LMWF) and light grey for the medium molecular weight fucoidan (MMWF) for (<b>A</b>) Raman and (<b>B</b>) FT-IR (in H<sub>2</sub>O and in D<sub>2</sub>O). The numbers indicates the characteristics bands for polysaccharides.</p>
Full article ">Figure 2
<p>Affinity measurement of fucoidans to SDF-1/CXCL12, RANTES/CCL5 and VEGF. The binding responses of ASPHY, MMWF, LMWF and low molecular weight heparin (LMWH) to SDF-1/CXCL12, RANTES/CCL5 and VEGF were measured by Surface Plasmon Resonance. We immobilized biotinylated SDF-1/CXCL12, RANTES/CCL5 or VEGF on streptavidin chip. Each polysaccharide was injected over flow of a BIAcore sensor chip pre-coated with streptavidin biotinylated SDF-1/CXCL12, RANTES/CCL5 or VEGF. Each set of sensorgrams was obtained by injecting increasing concentration of polysaccharides (1.2, 3.7, 11.1, 33.3, and 100 nM). The response unit (RU) was recorded as a function of time (sec) and the affinities are expressed in molar (M) with the equilibrium dissociation constant KD (Kd/Ka). LMWH was used as a positive control of sulfated polysaccharide whereas non-sulfated dextran was used as a negative control (not shown). Affinity of polysaccharides to (<b>A</b>) SDF-1/CXCL12; (<b>B</b>) RANTES/CCL5 and (<b>C</b>) VEGF, and their corresponding representative sensorgrams.</p>
Full article ">Figure 3
<p>Effect of fucoidans on cell viability. The viability of HUVECs was analyzed by using MTT assay after fucoidan treatment for 24 h. The absorbance was read with a spectrophotometer (at 570 nm). HUVECs were incubated with (<b>A</b>) LMWF and (<b>B</b>) MMWF at increasing concentration (1, 10, 100, and 1000 μg/mL); (<b>C</b>) HUVECs were incubated 24 h with polysaccharides (dextran, LMWH, LMWF, MMWF and ASPHY) at 10 μg/mL. Values are expressed as means ± SEM (<span class="html-italic">n</span> ≥ 3). AU-Arbitrary units. * <span class="html-italic">p</span> &lt; 0.05 versus Untreated.</p>
Full article ">Figure 4
<p>Pro-angiogenic potential of fucoidans on GAG-free HUVECs. (<b>A</b>) HUVECs pre-treated or not with βDX (4-Nitrophenyl-β-<span class="html-small-caps">d</span>-Xylopyranoside) were seeded on Matrigel and incubated with dextran, LMWH, LMWF, MMWF or ASPHY for 6 h. The cells were then stained with Hemalun Mayer’s and photographed for analysis. Values are expressed in number of nodes per well. ** <span class="html-italic">p</span> &lt; 0.01 LMWF or LMWH versus Untreated (all without βDX). # <span class="html-italic">p</span> &lt; 0.05 LMWF versus Untreated (all with βDX); (<b>B</b>) Endogenous GAGs expression analyzed by flow cytometry on HUVECs pre-treated or not 48 h with β<span class="html-italic">DX</span>; (<b>C</b>) PD98059, a pharmacological inhibitor of ERK1/2 and (<b>D</b>) LY294002, a pharmacological inhibitor of PI3K/AKT were added in HUVEC culture, then HUVECs were seeded on Matrigel for 6 h and vascular network formation was observed as described before. Values are expressed in nodes per well (<span class="html-italic">n</span> ≥ 3). * <span class="html-italic">p</span> &lt; 0.05 LMWF, PD98059, LY294002, LMWF + PD98059, LY294002 + LMWF versus Untreated; # <span class="html-italic">p</span> &lt; 0.05 LMWF + PD98059 versus PD98059.</p>
Full article ">Figure 5
<p>Pro-migratory potential of fucoidans on GAG-free HUVECs. HUVECs were seeded and incubated 24 h in the upper chamber with the polysaccharides dextran, LMWH, LMWF, MMWF or ASPHY at 10 μg/mL. The basal migration was performed in complete medium with 12% FBS. In the aim to remove the GAGs, the cells were pre-treated with βDX for 48 h, then the migration assay was performed with the same treatments as described above. The cells were fixed, stained with Mayer’s hemalun solution and counted after migration. Values are expressed as cell number per well. * <span class="html-italic">p</span> &lt; 0.05 ASPHY versus Untreated (all without βDX); ** <span class="html-italic">p</span> &lt; 0.01 LMWH and LMWF versus Untreated (all without βDX); $ <span class="html-italic">p</span> &lt; 0.05 LMWH and LMWF versus Untreated (all with βDX); # <span class="html-italic">p</span> &lt; 0.05 LMWH and LMWF and ASPHY (all without βDX) versus LMWH and LMWF and ASPHY (all with βDX).</p>
Full article ">Figure 6
<p>LMWF-Alexa localization in HUVECs by confocal microscopy. LMWF was previously coupled with the fluorophore Alexa Fluor 555. (<b>A</b>) LMWF-Alexa was added in HUVEC culture medium at 10 μg/mL during 2 h at 37 °C and 4 °C. Dextran-FITC and Alexa fluor alone (Alexa) were used as negative control. Pictures were taken by confocal microscopy and the intensity of the accumulated fluorescence per cell was quantified by using specific quantification software (DAPI—blue, LMWF-Alexa—red, Dextran-FITC—green, bar = 10 μm). Values are expressed as percentage of the intensity. * <span class="html-italic">p</span> &lt; 0.01 37 °C versus 4 °C; (<b>B</b>) Endogenous GAGs expression on HUVECs treated by heparinase I, II, and III and chondroitinase ABC (H/C); (<b>C</b>) LMWF-Alexa was incubated with HUVECs (30 min, 2 h, and 6 h) with or without heparinase I, II, and III and chondroitinase ABC (H/C) and the intensity of fluorescence per cell was measured by flow cytometry. Values are expressed as percentage of the intensity normalized on the maximum intensity reached at 2 h. The right panel shows the confocal observation. (DAPI—blue, LMWF-Alexa—red, bar = 10 μm). * <span class="html-italic">p</span> &lt; 0.05, 2 h, and 6 h versus 30 min (all Untreated with H/C); $ <span class="html-italic">p</span> &lt; 0.05, 6 h versus 30 min (all treated with H/C); # <span class="html-italic">p</span> &lt; 0.05, 2 h of treated with H/C versus 2 h of Untreated.</p>
Full article ">Figure 6 Cont.
<p>LMWF-Alexa localization in HUVECs by confocal microscopy. LMWF was previously coupled with the fluorophore Alexa Fluor 555. (<b>A</b>) LMWF-Alexa was added in HUVEC culture medium at 10 μg/mL during 2 h at 37 °C and 4 °C. Dextran-FITC and Alexa fluor alone (Alexa) were used as negative control. Pictures were taken by confocal microscopy and the intensity of the accumulated fluorescence per cell was quantified by using specific quantification software (DAPI—blue, LMWF-Alexa—red, Dextran-FITC—green, bar = 10 μm). Values are expressed as percentage of the intensity. * <span class="html-italic">p</span> &lt; 0.01 37 °C versus 4 °C; (<b>B</b>) Endogenous GAGs expression on HUVECs treated by heparinase I, II, and III and chondroitinase ABC (H/C); (<b>C</b>) LMWF-Alexa was incubated with HUVECs (30 min, 2 h, and 6 h) with or without heparinase I, II, and III and chondroitinase ABC (H/C) and the intensity of fluorescence per cell was measured by flow cytometry. Values are expressed as percentage of the intensity normalized on the maximum intensity reached at 2 h. The right panel shows the confocal observation. (DAPI—blue, LMWF-Alexa—red, bar = 10 μm). * <span class="html-italic">p</span> &lt; 0.05, 2 h, and 6 h versus 30 min (all Untreated with H/C); $ <span class="html-italic">p</span> &lt; 0.05, 6 h versus 30 min (all treated with H/C); # <span class="html-italic">p</span> &lt; 0.05, 2 h of treated with H/C versus 2 h of Untreated.</p>
Full article ">Figure 7
<p>Internalization pathways of LMWF-Alexa in HUVECs analyzed by confocal microscopy. (<b>A</b>) HUVECs were incubated 2 h with LMWF-Alexa, fixed, permeabilized and the clathrin or caveolin-1 was revealed by immunofluorescence. A non-specific isotype of immunoglobulin was used as negative control (Isotype). The pictures were taken by confocal microscopy and the staining overlaped in merge (DAPI—blue, LMWF-Alexa—red, Clathrin—green, bar = 10 μm) high view inserts. The intensity of the fluorescence was quantified and the co-localization of markers was measured with the rate red/green represented in the histogram and dot plots. The intensity of fluorescence in HUVECs was quantified using specific software. * <span class="html-italic">p</span> &lt; 0.05 Caveolin-1 versus Isotype; <b>*</b>* <span class="html-italic">p</span> &lt; 0.01 Clathrin versus Isotype; (<b>B</b>) HUVECs were pre-treated or not (control) with specific inhibitors of endocytosis: Cytochalasin D (CytD-inhibits phagocytosis and micropynocytosis), Dynasore (inhibits clathrin mediated endocytosis) and Filipin (inhibits lipid raft formation) before adding LMWF-Alexa in the culture medium for 2 h. The intensity of fluorescence by cells was quantified as described above.* <span class="html-italic">p</span> &lt; 0.05 Dynasore versus Control.</p>
Full article ">
4064 KiB  
Article
Molecular Weight-Dependent Immunostimulative Activity of Low Molecular Weight Chitosan via Regulating NF-κB and AP-1 Signaling Pathways in RAW264.7 Macrophages
by Bin Zheng, Zheng-Shun Wen, Yun-Juan Huang, Mei-Sheng Xia, Xing-Wei Xiang and You-Le Qu
Mar. Drugs 2016, 14(9), 169; https://doi.org/10.3390/md14090169 - 20 Sep 2016
Cited by 42 | Viewed by 7336
Abstract
Chitosan and its derivatives such as low molecular weight chitosans (LMWCs) have been found to possess many important biological properties, such as antioxidant and antitumor effects. In our previous study, LMWCs were found to elicit a strong immunomodulatory response in macrophages dependent on [...] Read more.
Chitosan and its derivatives such as low molecular weight chitosans (LMWCs) have been found to possess many important biological properties, such as antioxidant and antitumor effects. In our previous study, LMWCs were found to elicit a strong immunomodulatory response in macrophages dependent on molecular weight. Herein we further investigated the molecular weight-dependent immunostimulative activity of LMWCs and elucidated its mechanism of action on RAW264.7 macrophages. LMWCs (3 kDa and 50 kDa of molecular weight) could significantly enhance the mRNA expression levels of COX-2, IL-10 and MCP-1 in a molecular weight and concentration-dependent manner. The results suggested that LMWCs elicited a significant immunomodulatory response, which was dependent on the dose and the molecular weight. Regarding the possible molecular mechanism of action, LMWCs promoted the expression of the genes of key molecules in NF-κB and AP-1 pathways, including IKKβ, TRAF6 and JNK1, and induced the phosphorylation of protein IKBα in RAW264.7 macrophage. Moreover, LMWCs increased nuclear translocation of p65 and activation of activator protein-1 (AP-1, C-Jun and C-Fos) in a molecular weight-dependent manner. Taken together, our findings suggested that LMWCs exert immunostimulative activity via activation of NF-κB and AP-1 pathways in RAW264.7 macrophages in a molecular weight-dependent manner and that 3 kDa LMWC shows great potential as a novel agent for the treatment of immune suppression diseases and in future vaccines. Full article
Show Figures

Figure 1

Figure 1
<p>Effect of LMWCs on the mRNA expression levels of COX-2 (<b>A</b>), IL-10 (<b>B</b>) and MCP-1 (<b>C</b>) in RAW264.7 macrophage. Each cell population (1 × 10<sup>6</sup> cells/mL) was treated with LMWCs (3 kDa and 50 kDa) at the indicated concentrations of 2.5, 10 and 40 μg/mL or LPS (1 μg/mL) for 24 h, respectively. The untreated cells are used as the control. These represent mean values of three independent experiments. Values are presented as means ± SD (<span class="html-italic">n</span> = 3, three independent experiments). Bars with different letters (a, b, c, d, e, f) are statistically different (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>Effects of LMWCs on the mRNA expression levels of IKKβ in RAW264.7 macrophage. Each cell population (4 × 10<sup>5</sup> cells/mL) was treated with LMWCs (3 kDa and 50 kDa) at the indicated concentrations of 40 μg/mL or LPS (1 μg/mL) for 12 h, respectively. The untreated cells are used as the control. These represent mean values of three independent experiments. Values are presented as means ± SD (<span class="html-italic">n</span> = 3). Bars with different letters (a, b, c, d) are statistically different (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>Effects of LMWCs on the mRNA expression levels of key molecules (TRAF6 (<b>A</b>), JNK1 (<b>B</b>)) from RAW264.7 macrophage. Each cell population (4 × 10<sup>5</sup> cells/mL) was treated with LMWCs (3 kDa and 50 kDa) at the indicated concentrations of 40 μg/mL or LPS (1 μg/mL) for 12 h, respectively. The untreated cells are used as the control. These represent mean values of three independent experiments. Values are presented as means ± SD (<span class="html-italic">n</span> = 3). Bars with different letters (a, b, c) are statistically different (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Effect of LMWCs on the phosphorylation of IKBα in RAW264.7 macrophage. (<b>A</b>) Each cell population (4 × 10<sup>5</sup> cells/mL) was treated with LMWCs (3 kDa and 50 kDa) at the indicated concentrations of 40 μg/mL for 12 h, respectively (1: Control; 2: LPS; 3: 3 kDa chitosan; 4: 50 kDa chitosan); (<b>B</b>) Each cell population (4 × 10<sup>5</sup> cells/mL) was treated with LMWCs (3 kDa and 50 kDa) at the indicated concentrations of 40 μg/mL and LPS (1 μg/mL) for 12 h after pre-incubation with 20 μmol/L of wedelolactone (Wed) for 12 h, respectively (1: Control; 2: Wedelolactone (Wed); 3: Wed + LPS; 4: Wed + 3 kDa chitosan; 5: Wed + 50 kDa chitosan). The figures shown are representative of three independent experiments. These represent mean values of three independent experiments. Values are presented as means ± SD (<span class="html-italic">n</span> = 3). Bars with different letters (a, b, c, d) are statistically different (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Effect of LMWCs on the nuclear translocation of p65 in the NF-κB pathway in RAW264.7 macrophages. Each cell population (1 × 10<sup>6</sup> cells/mL) was treated with LMWCs (3 kDa and 50 kDa) at the indicated concentrations of 40 μg/mL for 24 h, respectively (1: Control; 2: 50 kDa chitosan; 3: 3 kDa chitosan; 4: LPS). The figures shown are representative of three independent experiments. These represent mean values of three independent experiments. Values are presented as means ± SD (<span class="html-italic">n</span> = 3). Bars with different letters (a, b, c, d) are statistically different (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 6
<p>Effect of LMWCs on activator protein-1 (AP-1) in RAW264.7 macrophages. Each cell population (1 × 10<sup>6</sup> cells/mL) was treated with LMWCs (3 kDa and 50 kDa) at the indicated concentrations of 40 μg/mL for 24 h, respectively (1: Control; 2: 50 kDa chitosan; 3: 3 kDa chitosan; 4: LPS). The figures shown are representative of three independent experiments. These represent mean values of three independent experiments. Values are presented as means ± SD (<span class="html-italic">n</span> = 3). Bars with different letters (a, b, c, d) are statistically different (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 7
<p>Schematic diagram of the targets of LMWCs.</p>
Full article ">
806 KiB  
Review
Fucoidans in Nanomedicine
by Lucas Chollet, Pierre Saboural, Cédric Chauvierre, Jean-Noël Villemin, Didier Letourneur and Frédéric Chaubet
Mar. Drugs 2016, 14(8), 145; https://doi.org/10.3390/md14080145 - 29 Jul 2016
Cited by 87 | Viewed by 11967
Abstract
Fucoidans are widespread cost-effective sulfated marine polysaccharides which have raised interest in the scientific community over last decades for their wide spectrum of bioactivities. Unsurprisingly, nanomedicine has grasped these compounds to develop innovative therapeutic and diagnostic nanosystems. The applications of fucoidans in nanomedicine [...] Read more.
Fucoidans are widespread cost-effective sulfated marine polysaccharides which have raised interest in the scientific community over last decades for their wide spectrum of bioactivities. Unsurprisingly, nanomedicine has grasped these compounds to develop innovative therapeutic and diagnostic nanosystems. The applications of fucoidans in nanomedicine as imaging agents, drug carriers or for their intrinsic properties are reviewed here after a short presentation of the main structural data and biological properties of fucoidans. The origin and the physicochemical specifications of fucoidans are summarized in order to discuss the strategy of fucoidan-containing nanosystems in Human health. Currently, there is a need for reproducible, well characterized fucoidan fractions to ensure significant progress. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Evolution of published articles reporting fucoidans (from Web of Science). Left axis: number of articles for “Fucoidan*”, right axis: number of articles for “Fucoidan* + Nano*”.</p>
Full article ">Figure 2
<p>Repeating chemical structures of some fucoidans from brown algae (<b>A</b>) <span class="html-italic">Chorda filum</span> [<a href="#B28-marinedrugs-14-00145" class="html-bibr">28</a>]; (<b>B</b>) <span class="html-italic">Ascophyllum nodosum</span>, <span class="html-italic">Fucus vesiculosus,</span> and <span class="html-italic">Fucus evanescens</span> [<a href="#B22-marinedrugs-14-00145" class="html-bibr">22</a>,<a href="#B23-marinedrugs-14-00145" class="html-bibr">23</a>,<a href="#B31-marinedrugs-14-00145" class="html-bibr">31</a>] and from marine invertebrates: sea cucumber (<span class="html-italic">Holothuriodea</span>) (<b>C</b>) <span class="html-italic">Ludwigothuria grisea</span> [<a href="#B29-marinedrugs-14-00145" class="html-bibr">29</a>]; (<b>D</b>) <span class="html-italic">Strongylocentrotus droebachiensis</span> [<a href="#B17-marinedrugs-14-00145" class="html-bibr">17</a>], and (<b>E</b>) <span class="html-italic">Strongylocentrotus franciscanus</span> [<a href="#B30-marinedrugs-14-00145" class="html-bibr">30</a>].</p>
Full article ">
3286 KiB  
Article
Electrospinning of Nanodiamond-Modified Polysaccharide Nanofibers with Physico-Mechanical Properties Close to Natural Skins
by Mina Mahdavi, Nafiseh Mahmoudi, Farzad Rezaie Anaran and Abdolreza Simchi
Mar. Drugs 2016, 14(7), 128; https://doi.org/10.3390/md14070128 - 7 Jul 2016
Cited by 55 | Viewed by 7972
Abstract
Electrospinning of biopolymers has gained significant interest for the fabrication of fibrous mats for potential applications in tissue engineering, particularly for wound dressing and skin regeneration. In this study, for the first time, we report successful electrospinning of chitosan-based biopolymers containing bacterial cellulous [...] Read more.
Electrospinning of biopolymers has gained significant interest for the fabrication of fibrous mats for potential applications in tissue engineering, particularly for wound dressing and skin regeneration. In this study, for the first time, we report successful electrospinning of chitosan-based biopolymers containing bacterial cellulous (33 wt %) and medical grade nanodiamonds (MND) (3 nm; up to 3 wt %). Morphological studies by scanning electron microscopy showed that long and uniform fibers with controllable diameters from 80 to 170 nm were prepared. Introducing diamond nanoparticles facilitated the electrospinning process with a decrease in the size of fibers. Fourier transform infrared spectroscopy determined hydrogen bonding between the polymeric matrix and functional groups of MND. It was also found that beyond 1 wt % MND, percolation networks of nanoparticles were formed which affected the properties of the nanofibrous mats. Uniaxial tensile testing of the woven mats determined significant enhancement of the strength (from 13 MPa to 25 MP) by dispersion of 1 wt % MND. The hydrophilicity of the mats was also remarkably improved, which was favorable for cell attachment. The water vapor permeability was tailorable in the range of 342 to 423 µg·Pa−1·s−1·m−1. The nanodiamond-modified mats are potentially suitable for wound healing applications. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Effect of medical grade nanodiamonds (MND) on the morphology and size distribution of electrospun fibers: (<b>a</b>) chitosan/bacterial cellulose (CS/BC) without MND; (<b>c</b>) contain 1%; (<b>e</b>) 2% and (<b>g</b>) 3% MND particles, respectively. (<b>b</b>), (<b>d</b>), (<b>f</b>) and (<b>h</b>) show the fiber diameter distribution diagrams of each specimen.</p>
Full article ">Figure 2
<p>Formation of large nanoparticle clusters upon electrospinning. The concentration of MND (%) is (<b>a</b>) 2 and (<b>b</b>) 3.</p>
Full article ">Figure 3
<p>Fourier transform infrared (FT-IR) spectrum of (<b>a</b>) CS/BC polymer; (<b>b</b>) the nanocomposite fiber containing 3% MND; and (<b>c</b>) pristine MND.</p>
Full article ">Figure 4
<p>Effect of diamond particles on the hydrophilicity of electrospun CS/BC mats.</p>
Full article ">Figure 5
<p>Stress-strain curves of electrospun mats containing different amounts of diamond nanoparticles.</p>
Full article ">Figure 6
<p>Weight change per unit of area of the mats versus time for CS/BC mats containing different amounts of MND.</p>
Full article ">Figure 7
<p>Cell viability of CS/BC mats dependent on the MND content at two incubated times.</p>
Full article ">
7295 KiB  
Article
Anticancer Effect of Fucoidan on DU-145 Prostate Cancer Cells through Inhibition of PI3K/Akt and MAPK Pathway Expression
by Gang-Sik Choo, Hae-Nim Lee, Seong-Ah Shin, Hyeong-Jin Kim and Ji-Youn Jung
Mar. Drugs 2016, 14(7), 126; https://doi.org/10.3390/md14070126 - 7 Jul 2016
Cited by 55 | Viewed by 8554
Abstract
In this study, we showed that PI3K/Akt signaling mediates fucoidan’s anticancer effects on prostate cancer cells, including suppression of proliferation. Fucoidan significantly decreased viability of DU-145 cancer cells in a concentration-dependent manner as shown by MTT [3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide] assay. The drug also significantly [...] Read more.
In this study, we showed that PI3K/Akt signaling mediates fucoidan’s anticancer effects on prostate cancer cells, including suppression of proliferation. Fucoidan significantly decreased viability of DU-145 cancer cells in a concentration-dependent manner as shown by MTT [3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide] assay. The drug also significantly increased chromatin condensation, which indicates apoptosis, in a concentration-dependent manner as shown by DAPI (4′,6-diamidino-2-phenylindole) staining. Fucoidan increased expression of Bax, cleaved poly-ADP ribose polymerase and cleaved caspase-9, and decreased of the Bcl-2, p-Akt, p-PI3K, p-P38, and p-ERK in a concentration-dependent manner. In vivo, fucoidan (at 5 and 10 mg/kg) significantly decreased tumor volume, and increased apoptosis as assessed by the TUNEL (terminal deoxynucleotidyl transferase dUTP nick end labeling) assay, confirming the tumor inhibitory effect. The drug also increased expression of p-Akt and p-ERK as shown by immunohistochemistry staining. Therefore, fucoidan may be a promising cancer preventive medicine due to its growth inhibitory effects and induction of apoptosis in human prostate cancer cells. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Effect of fucoidan on the cell viability of DU-145 cells. DU-145 cells (2 × 10<sup>4</sup> cells/mL) were treated with 0, 250, 500, 750, 1000 μg/mL fucoidan in RPMI-1640 medium containing 5% FBS for 24 h. The growth inhibition was measured by the MTT assay. Data are mean standard deviation (SD) for three samples. The significance was determined by Student’s <span class="html-italic">t</span>-test (* <span class="html-italic">p</span> &lt; 0.05 compared with untreated control).</p>
Full article ">Figure 2
<p>Effect of fucoidan on the chromatin condensation in DU-145 cells. (<b>A</b>) DU-145 cells were treated with 0, 500, 1000 μg/mL fucoidan or vehicle in RPMI-1640 medium containing 5% FBS for 24 h, and cell were stained with DAPI. The arrows indicate chromatin condensation in the cancer cell. (<b>B</b>) DU-145 cells were treated with fucoidan (0, 500, 1000 μg/mL) for 24 h. Apoptosis cells were counted under a light microscope and expressed as the average of five fields. Each bar represents the mean ± SD calculated from independent experiments. Significance was determined by Dunnett’s <span class="html-italic">t</span>-test with * <span class="html-italic">p</span> &lt; 0.05 compared as statistically significant compared with non-treated controls.</p>
Full article ">Figure 3
<p>Effect of fucoidan on the apoptotic pathway in DU-145 cells. DU-145 cells were treated with fucoidan 0, 500, and 1000 μg/mL for 24 h and cell were harvested to measure protein levels of Bax, Bcl-2, cspase-9, and PARP by western blotting. The blots were also probed with β-actin antibodies to confirm equal sample loading.</p>
Full article ">Figure 4
<p>Effect of fucoidan on the activation of PI3K/Akt pathway in DU-145 cells. Cells were treated with fucoidan 0, 500 and 1000 μg/mL for 24 h. Cell lysates were prepared as described in the materials and methods and analyzed by 12% SDS-PAGE followed by western blotting. (<b>A</b>) The membranes were incubated with PI3K/AKT pathway antibodies. (<b>B</b>) The membranes were incubated with MAPKs pathway antibodies. Each bar represents the mean ± SD calculated from independent experiments. Significance was determined by Dunnett’s <span class="html-italic">t</span>-test with * <span class="html-italic">p</span> &lt;0.05 compared as statistically significant compared with non-treated controls.</p>
Full article ">Figure 5
<p>Inhibition of DU-145 prostate tumor growth and enhancement of apoptosis in DU-145 prostate tumors by the fucoidan. (<b>A</b>) To identify the effect of fucoidan in DU-145 prostate tumor growth, nude mice were treated with fucoidan (0, 5, 10 mg/kg) for 21 days (<span class="html-italic">n</span> = 5). (<b>B</b>) The graph expresses final tumor weight. (<b>C</b>) The graph is nude mice weight. Each value was expressed as mean ± SE of five mice. Significance was determined by Dunnett’s <span class="html-italic">t</span>-test with * <span class="html-italic">p</span> &lt;0.05 compared as statistically significant compared with non-treated controls.</p>
Full article ">Figure 6
<p>Induction of apoptosis by fucoidan in DU-145 cells. Nude mice were treated with fucoidan for 21 days and apoptosis was assessed by terminal deoxynucleotidyltransferase-mediated Dutp nick-ned labeling (TUNEL) assay. Tumor tissues were observed under a microscope and photographed at a ×200 magnification. The percentage of labeled with TUNEL-positive apoptotic cells was calculated from 1,000 scored cells. Paraffin-embedded tumors were cut into 5 μm sections. Each bar represents the mean ± SD calculated from independent experiments. Significance was determined by Dunnett’s <span class="html-italic">t</span>-test with * <span class="html-italic">p</span> &lt;0.05 compared as statistically significant compared with non-treated controls. Scale bar, 10 µm.</p>
Full article ">Figure 7
<p>Effect of fucoidan on p-Akt and p-ERK expression in DU-145 prostate tumors. Nude mice were administered fucoidan (0, 5 and 10 mg/kg) for three weeks and assayed by immunohistochemistry using p-Akt and p-ERK antibodes. Tumor tissues were observed under a microscope and photographed at a ×400 magnification. Paraffin-embedded tumors were sectioned to a thickness of 5 µm. Scale bar, 5 µm.</p>
Full article ">Figure 8
<p>Histological observation of nude mice treated intraperitoneally with fucoidan. Fucoidan was administered at a dose of 5 or 10 mg/kg five times per week, for a total 21 injections. On day 21, mice were sacrificed, and tumors excised and evaluated by hematoxylin &amp; eosin (H &amp; E) staining (×200). The dose of fucoidan had no detectable toxicological effect on nude mice. Scale bar, 10 µm.</p>
Full article ">
416 KiB  
Article
Toxicological Evaluation of Low Molecular Weight Fucoidan in Vitro and in Vivo
by Pai-An Hwang, Ming-De Yan, Hong-Ting Victor Lin, Kuan-Lun Li and Yen-Chang Lin
Mar. Drugs 2016, 14(7), 121; https://doi.org/10.3390/md14070121 - 24 Jun 2016
Cited by 46 | Viewed by 7764
Abstract
For a long time, fucoidan has been well known for its pharmacological activities, and recently low molecular weight fucoidan (LMF) has been used in food supplements and pharmaceutical products. In the present study, LMF was extracted from Laminaria japonica by enzyme hydrolysis. The [...] Read more.
For a long time, fucoidan has been well known for its pharmacological activities, and recently low molecular weight fucoidan (LMF) has been used in food supplements and pharmaceutical products. In the present study, LMF was extracted from Laminaria japonica by enzyme hydrolysis. The toxicity of LMF in mouse and rat models was determined by many methods, such as total arsenic content, bacterial reverse mutation assay, chromosome aberration assay, and in vivo micronucleus assay. The present findings showed that LMF at 5000 μg/mL exhibited no mutagenicity. It also produced no formatting disruption of red blood cells in vivo. At 2000 mg/kg BW/day there were no toxicological indications. LMF is expected to be used as a safe food supplement. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Growth curves for male (open symbols) and female (solid symbols) rats treated with LMF-LJ for 28 days. Values were expressed as mean ± SD, <span class="html-italic">n</span> = 10.</p>
Full article ">
3766 KiB  
Article
Effect of Experimental Parameters on Alginate/Chitosan Microparticles for BCG Encapsulation
by Liliana A. Caetano, António J. Almeida and Lídia M.D. Gonçalves
Mar. Drugs 2016, 14(5), 90; https://doi.org/10.3390/md14050090 - 11 May 2016
Cited by 82 | Viewed by 10329
Abstract
The aim of the present study was to develop novel Mycobacterium bovis bacille Calmette-Guérin (BCG)-loaded polymeric microparticles with optimized particle surface characteristics and biocompatibility, so that whole live attenuated bacteria could be further used for pre-exposure vaccination against Mycobacterium tuberculosis by the intranasal [...] Read more.
The aim of the present study was to develop novel Mycobacterium bovis bacille Calmette-Guérin (BCG)-loaded polymeric microparticles with optimized particle surface characteristics and biocompatibility, so that whole live attenuated bacteria could be further used for pre-exposure vaccination against Mycobacterium tuberculosis by the intranasal route. BCG was encapsulated in chitosan and alginate microparticles through three different polyionic complexation methods by high speed stirring. For comparison purposes, similar formulations were prepared with high shear homogenization and sonication. Additional optimization studies were conducted with polymers of different quality specifications in a wide range of pH values, and with three different cryoprotectors. Particle morphology, size distribution, encapsulation efficiency, surface charge, physicochemical properties and biocompatibility were assessed. Particles exhibited a micrometer size and a spherical morphology. Chitosan addition to BCG shifted the bacilli surface charge from negative zeta potential values to strongly positive ones. Chitosan of low molecular weight produced particle suspensions of lower size distribution and higher stability, allowing efficient BCG encapsulation and biocompatibility. Particle formulation consistency was improved when the availability of functional groups from alginate and chitosan was close to stoichiometric proportion. Thus, the herein described microparticulate system constitutes a promising strategy to deliver BCG vaccine by the intranasal route. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Microparticles domain formation using high, medium and low molecular weight chitosan. The pH of alginate and chitosan solutions was initially set to 4.9 and 4.6, respectively. Three different systems were identified: clear solution (♦), opalescent/colloidal suspension (■), and aggregates (▲).</p>
Full article ">Figure 2
<p>Effect of alginate to chitosan mass ratio on particle surface charge. The pH of alginate and chitosan solutions was initially set to 4.9 and 4.6, respectively. Zeta potential of microparticles prepared with chitosan of low molecular weight (□), medium molecular weight (◌), and high molecular weight (∆). Results are presented as mean ± SD (<span class="html-italic">n</span> ≥ 3).</p>
Full article ">Figure 3
<p>(<b>A</b>) Particle size distribution of plain chitosan-alginate microparticles of 1:1 ALG/CS mass ratio, prepared with chitosan of low molecular weight and alginates of decreasing G-content, according to Method (II) (solid) and Method (III) (dashed); (<b>B</b>) Zeta potential of plain chitosan-alginate microparticles of 1:1 ALG/CS mass ratio, prepared with chitosan of low molecular weight and alginates of decreasing G-content, according to Method (II) (□) and Method (III) (◌).The pH of alginate, chitosan and TPP solutions was initially set to 6.7, 4.1 and 9.0, respectively. Results are presented as mean ± SD (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 4
<p>FT-IR spectra of plain chitosan-alginate “F14_Low” microparticles (1:1 ALG/CS mass ratio) with increasing pH of the microparticles suspension. Bands wave numbers (cm<sup>−1</sup>) are as follows: 1641 (amide bond), 1613 (symmetric COO<sup>−</sup> stretching vibration), 1569 (strong protonated amino peak—from partial N-deacetylation of chitin), and 1415 (asymmetric COO<sup>−</sup> stretching vibration).</p>
Full article ">Figure 5
<p>(<b>A</b>) Polarized light micrograph (100×) of “F13_Medium” microparticles (0.8:1 ALG/CS) prepared according to Method (III) with chitosan of medium molecular weight; (<b>B</b>) Contrast phase micrograph (40×) of “F14_Low<span class="html-italic">”</span> microparticles (1:1 ALG/CS) prepared according to Method (II) with chitosan of low molecular weight.</p>
Full article ">Figure 6
<p>Particle size distribution of microparticles produced with alginate to chitosan ratio of 4.23:1 (F0), 0.8:1 (F13), and 1:1 (F14). F0 microparticles prepared according to Method (I) by alginate ionotropic pre-gelation with CaCl<sub>2</sub> followed by chitosan coating; F13-F14 microparticles prepared according to Method (III) by chitosan pre-gelation with TPP followed by alginate coating. LMW, low molecular weight chitosan; MMW, medium molecular weight chitosan; HMW, high molecular weight chitosan.</p>
Full article ">Figure 7
<p>Cell viability of BCG (<b>A</b>) After 3 weeks of incubation, agar (Middlebrook 7H10 medium supplemented with OADC) plates inoculated with bacteria of both strains (Pasteur and GFP) that presented a number of colonies of statistical relevance were used to calculate the CFU/mL, by multiplying the colony forming units by the plating factor and the dilution factor. The CFU/mL provides an approximation of the cell density of the original culture. BCG suspension in 0.9% NaCl was used as control; (<b>B</b>) BCG Pasteur viability following BCG microencapsulation in “F14_Low” chitosan-alginate microparticles (no fill), BCG suspension in 0.025% low molecular weight chitosan weight (horizontal lines), or BCG suspension in 0.9% NaCl (angled lines). Results are expressed as mean ± S.D.; <span class="html-italic">n</span> = 3.</p>
Full article ">Figure 8
<p>Relative cell viability of THP-1 cell line measured by the MTT reduction. Columns: black—control cells with culture medium; dark grey—BCG-GFP/0.9% NaCl; light grey—BCG-GFP/0.025% LMW chitosan; dotted white—BCG Pasteur/F13_Medium microparticles; dotted grey—BCG Pasteur/F13_High microparticles (1 × 108 CFUs/mL). Results are expressed as mean ± SD (<span class="html-italic">n</span> = 3). Statistical differences between the control group and formulations are reported as: *** <span class="html-italic">p</span> &lt; 0.001, ** <span class="html-italic">p</span> &lt; 0.01, * <span class="html-italic">p</span> &lt; 0.05. Cell viability (% of control) = [A] test/[A] control × 100.</p>
Full article ">Figure 9
<p>Microparticles formation by alginate ionotropic pre-gelation with CaCl<sub>2</sub> followed by chitosan addition (adapted from [<a href="#B33-marinedrugs-14-00090" class="html-bibr">33</a>]).</p>
Full article ">Figure 10
<p>Microparticles formation by chitosan gel matrix formation with sodium alginate followed by TPP addition (adapted from [<a href="#B32-marinedrugs-14-00090" class="html-bibr">32</a>,<a href="#B34-marinedrugs-14-00090" class="html-bibr">34</a>]).</p>
Full article ">Figure 11
<p>Microparticles formation by chitosan precipitation with TPP followed by alginate addition.</p>
Full article ">
3038 KiB  
Review
Marine Origin Polysaccharides in Drug Delivery Systems
by Matias J. Cardoso, Rui R. Costa and João F. Mano
Mar. Drugs 2016, 14(2), 34; https://doi.org/10.3390/md14020034 - 5 Feb 2016
Cited by 219 | Viewed by 17265
Abstract
Oceans are a vast source of natural substances. In them, we find various compounds with wide biotechnological and biomedical applicabilities. The exploitation of the sea as a renewable source of biocompounds can have a positive impact on the development of new systems and [...] Read more.
Oceans are a vast source of natural substances. In them, we find various compounds with wide biotechnological and biomedical applicabilities. The exploitation of the sea as a renewable source of biocompounds can have a positive impact on the development of new systems and devices for biomedical applications. Marine polysaccharides are among the most abundant materials in the seas, which contributes to a decrease of the extraction costs, besides their solubility behavior in aqueous solvents and extraction media, and their interaction with other biocompounds. Polysaccharides such as alginate, carrageenan and fucoidan can be extracted from algae, whereas chitosan and hyaluronan can be obtained from animal sources. Most marine polysaccharides have important biological properties such as biocompatibility, biodegradability, and anti-inflammatory activity, as well as adhesive and antimicrobial actions. Moreover, they can be modified in order to allow processing them into various shapes and sizes and may exhibit response dependence to external stimuli, such as pH and temperature. Due to these properties, these biomaterials have been studied as raw material for the construction of carrier devices for drugs, including particles, capsules and hydrogels. The devices are designed to achieve a controlled release of therapeutic agents in an attempt to fight against serious diseases, and to be used in advanced therapies, such as gene delivery or regenerative medicine. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Interrelations of marine origin polysaccharides in drug delivery systems for advances therapies and applications.</p>
Full article ">Figure 2
<p>Marine origin polysaccharides categorized by electrostatic nature and carboxylated/sulfated structure.</p>
Full article ">Figure 3
<p>Optical microscope images of alginate microspheres before (<b>A</b>) and after (<b>B</b>) ultrasound exposure. Reprinted with permission from [<a href="#B51-marinedrugs-14-00034" class="html-bibr">51</a>], Copyright © 2014 American Chemical Society.</p>
Full article ">Figure 4
<p>Transmission electron microscopy (TEM) micrograph of chitosan/carrageenan nanoparticles (<b>A</b>). Ovalbumin release profile from chitosan-carrageenan nanoparticles (<b>B</b>). Adapted with permission from [<a href="#B108-marinedrugs-14-00034" class="html-bibr">108</a>], Copyright © 2009 Wiley Periodicals, Inc.</p>
Full article ">Figure 5
<p>TEM image of chitosan/fucoidan nanoparticles (<b>A</b>). Gentamicin release kinetics from chitosan/fucoidan particles (<b>B</b>). Adapted with permission from [<a href="#B119-marinedrugs-14-00034" class="html-bibr">119</a>], Copyright © 2014 distributed under a Creative Commons Attribution License.</p>
Full article ">
596 KiB  
Review
Fucoidan as a Potential Therapeutic for Major Blinding Diseases—A Hypothesis
by Alexa Klettner
Mar. Drugs 2016, 14(2), 31; https://doi.org/10.3390/md14020031 - 3 Feb 2016
Cited by 36 | Viewed by 8380
Abstract
Fucoidan is a heterogeneous group of sulfated polysaccharide with a high content of l-fucose, which can be extracted from brown algae and marine invertebrates. It has many beneficial biological activities that make fucoidan an interesting candidate for therapeutic application in a variety [...] Read more.
Fucoidan is a heterogeneous group of sulfated polysaccharide with a high content of l-fucose, which can be extracted from brown algae and marine invertebrates. It has many beneficial biological activities that make fucoidan an interesting candidate for therapeutic application in a variety of diseases. Age-related macular degeneration and diabetic retinopathy are major causes for vision loss and blindness in the industrialized countries and increasingly in the developing world. Some of the characteristics found in certain fucoidans, such as its anti-oxidant activity, complement inhibition or interaction with the Vascular Endothelial Growth factor, which would be of high interest for a potential application of fucoidan in age-related macular degeneration or diabetic retinopathy. However, the possible usage of fucoidan in ophthalmological diseases has received little attention so far. In this review, biological activities of fucoidan that could be of interest regarding these diseases will be discussed. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic of potential beneficial effects of fucoidan, depicted in red, on age related macular degeneration (AMD) or diabetic retinopathy (DR). Additional abbreviations: complement component (C), complement factor B (CFB), endothelial nitric oxide synthase (eNOS), glucose transporter type 4 (Glut4), nitric oxide (NO), nuclear factor erythroid-2 related factor 2 (Nrf2), reactive oxygen species (ROS), Vascular Endothelial Growth Factor (VEGF).</p>
Full article ">

2015

Jump to: 2024, 2022, 2021, 2020, 2019, 2018, 2017, 2016, 2014, 2013, 2012, 2011, 2010

1534 KiB  
Article
Structural and Immunological Activity Characterization of a Polysaccharide Isolated from Meretrix meretrix Linnaeus
by Li Li, Heng Li, Jianying Qian, Yongfeng He, Jialin Zheng, Zhenming Lu, Zhenghong Xu and Jinsong Shi
Mar. Drugs 2016, 14(1), 6; https://doi.org/10.3390/md14010006 - 29 Dec 2015
Cited by 15 | Viewed by 5948
Abstract
Polysaccharides from marine clams perform various biological activities, whereas information on structure is scarce. Here, a water-soluble polysaccharide MMPX-B2 was isolated from Meretrix meretrix Linnaeus. The proposed structure was deduced through characterization and its immunological activity was investigated. MMPX-B2 consisted of d-glucose [...] Read more.
Polysaccharides from marine clams perform various biological activities, whereas information on structure is scarce. Here, a water-soluble polysaccharide MMPX-B2 was isolated from Meretrix meretrix Linnaeus. The proposed structure was deduced through characterization and its immunological activity was investigated. MMPX-B2 consisted of d-glucose and d-galctose residues at a molar ratio of 3.51:1.00. The average molecular weight of MMPX-B2 was 510 kDa. This polysaccharide possessed a main chain of (1→4)-linked-α-d-glucopyranosyl residues, partially substituted at the C-6 position by a few terminal β-d-galactose residues or branched chains consisting of (1→3)-linked β-d-galactose residues. Preliminary immunological tests in vitro showed that MMPX-B2 could stimulate the murine macrophages to release various cytokines, and the structure-activity relationship was then established. The present study demonstrated the potential immunological activity of MMPX-B2, and provided references for studying the active ingredients in M. meretrix. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Purification of MMPX-B2. (<b>a</b>) Elution profile of crude polysaccharides by DEAE-52 cellulose; (<b>b</b>) Purification profile of MMPX-B by Superdex 200; (<b>c</b>) HP-GPC profile of MMPX-B2.</p>
Full article ">Figure 2
<p>GC profile of MMPX-B2 with acid hydrolysis and acetylation. (<b>A</b>) <span class="html-small-caps">d</span>-glucose; (<b>B</b>) <span class="html-small-caps">d</span>-galactose; (<b>C</b>) internal standard inositol.</p>
Full article ">Figure 3
<p>FT-IR spectrum of MMPX-B2.</p>
Full article ">Figure 4
<p>NMR spectra of MMPX-B2. (<b>a</b>) <sup>13</sup>C-NMR; (<b>b</b>) <sup>1</sup>H-NMR.</p>
Full article ">Figure 5
<p>Proposed structural model of MMPX-B2.</p>
Full article ">Figure 6
<p>Effects of MMPX-B2 on macrophages-mediated immunity <span class="html-italic">in vitro</span>. (<b>a</b>) Nitrate accumulation; (<b>b</b>) TNF-α; (<b>c</b>) IL-1β; (<b>d</b>) IL-6. ** is representative of <span class="html-italic">p</span> &lt; 0.01 and *** is representative of <span class="html-italic">p</span> &lt; 0.001, when compared to the control group.</p>
Full article ">
2961 KiB  
Article
Characterization and Comparison of the Structural Features, Immune-Modulatory and Anti-Avian Influenza Virus Activities Conferred by Three Algal Sulfated Polysaccharides
by Lin Song, Xiaolin Chen, Xiaodong Liu, Fubo Zhang, Linfeng Hu, Yang Yue, Kecheng Li and Pengcheng Li
Mar. Drugs 2016, 14(1), 4; https://doi.org/10.3390/md14010004 - 29 Dec 2015
Cited by 65 | Viewed by 7693
Abstract
Three marine macroalgae, i.e., Grateloupia filicina, Ulva pertusa and Sargassum qingdaoense, were selected as the deputies of Rhodophyta, Chlorophyta and Ochrophyta for comparative analysis of the molecular structures and biological activities of sulfated polysaccharides (SP). The ratio of water-soluble polysaccharides, [...] Read more.
Three marine macroalgae, i.e., Grateloupia filicina, Ulva pertusa and Sargassum qingdaoense, were selected as the deputies of Rhodophyta, Chlorophyta and Ochrophyta for comparative analysis of the molecular structures and biological activities of sulfated polysaccharides (SP). The ratio of water-soluble polysaccharides, the monosaccharide composition and the sulfated contents of three extracted SPs were determined, and their structures were characterized by Fourier transformation infrared spectroscopy. In addition, biological activity analysis showed that all three SPs had immune-modulatory activity both in vitro and in vivo, and SPs from S. qingdaoense had the best effect. Further bioassays showed that three SPs could not only enhance the immunity level stimulated by inactivated avian influenza virus (AIV) in vivo but also significantly inhibited the activity of activated AIV (H9N2 subtype) in vitro. G. filicina SP exhibited the strongest anti-AIV activity. These results revealed the variations in structural features and bioactivities among three SPs and indicated the potential adjuvants for immune-enhancement and anti-AIV. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>FT-IR spectra of three extracted sulfated polysaccharides. GFP, <span class="html-italic">Grateloupia filicina</span>; UPP, <span class="html-italic">Ulva Pertusa</span>; SQP, <span class="html-italic">Sargassum qingdaoens</span>.</p>
Full article ">Figure 2
<p>Mouse spleen cell proliferation effects of GFP, UPP and SQP. Mock treated with PBS instead of polysaccharides as a negative control. Values with different letters in the same column (a–d) are significantly different (<span class="html-italic">p</span> &lt; 0.05) from each other. Data are shown as the Mean + SD and are fully representative of the individual experiment.</p>
Full article ">Figure 3
<p>Avian influenza virus (AIV)-specific antibody titer detection. Kunming mice were immunized with an AIV vaccine and polysaccharides, following the prime-boost vaccination programme (days 0 and 14), respectively. (<b>A</b>) GFP; (<b>B</b>) UPP; (<b>C</b>) SQP. Values with different letters in the same column (a–d) are significantly different (<span class="html-italic">p</span> &lt; 0.05) from each other. Data are shown as the Mean + SD and are fully representative of an individual experiment.</p>
Full article ">Figure 4
<p>Cytokine production stimulating effect of GFP, UPP and SQP. Kunming mice were immunized with an AIV vaccine and polysaccharides, and sera were collected on day 28 after two immunizations to detect the cytokines IFN-γ (<b>A</b>) and IL-4 (<b>B</b>). Values with different letters in the same column (a–d) are significantly different (<span class="html-italic">p</span> &lt; 0.05) from each other. Data are shown as the Mean + SD and are fully representative for the individual experiment.</p>
Full article ">Figure 5
<p>T-cell subpopulation tests. The blood cells of the treated mice were collected and analyzed with flow cytometry. (<b>A</b>) CD3+CD4+. (<b>B</b>) CD3+CD8+. Values with different letters in the same column (a–d) are significantly different (<span class="html-italic">p</span> &lt; 0.05) from each other. Data are shown as the Mean + SD and are fully representative of the individual experiment.</p>
Full article ">Figure 6
<p>Haemagglutination test (HA) of the cell culture and relative expression of H9N2. Antiviral activity <span class="html-italic">in vitro</span> was measured with a HA test (<b>A</b>) and RT-PCR (<b>B</b>). Data from samples without polysaccharides were used as a basic control. Values with different letters in the same column (a–c) are significantly different (<span class="html-italic">p</span> &lt; 0.05) from each other. Data are shown as the Mean + SD and are fully representative of the individual experiment.</p>
Full article ">
1701 KiB  
Article
Heparanase and Syndecan-4 Are Involved in Low Molecular Weight Fucoidan-Induced Angiogenesis
by Oualid Haddad, Erwan Guyot, Nicolas Marinval, Fabien Chevalier, Loïc Maillard, Latifa Gadi, Christelle Laguillier-Morizot, Olivier Oudar, Angela Sutton, Nathalie Charnaux and Hanna Hlawaty
Mar. Drugs 2015, 13(11), 6588-6608; https://doi.org/10.3390/md13116588 - 28 Oct 2015
Cited by 8 | Viewed by 6579
Abstract
Induction of angiogenesis is a potential treatment for chronic ischemia. Low molecular weight fucoidan (LMWF), the sulfated polysaccharide from brown seaweeds, has been shown to promote revascularization in a rat limb ischemia, increasing angiogenesis in vivo. We investigated the potential role of [...] Read more.
Induction of angiogenesis is a potential treatment for chronic ischemia. Low molecular weight fucoidan (LMWF), the sulfated polysaccharide from brown seaweeds, has been shown to promote revascularization in a rat limb ischemia, increasing angiogenesis in vivo. We investigated the potential role of two heparan sulfate (HS) metabolism enzymes, exostosin-2 (EXT2) and heparanase (HPSE), and of two HS-membrane proteoglycans, syndecan-1 and -4 (SDC-1 and SDC-4), in LMWF induced angiogenesis. Our results showed that LMWF increases human vascular endothelial cell (HUVEC) migration and angiogenesis in vitro. We report that the expression and activity of the HS-degrading HPSE was increased after LMWF treatment. The phenotypic tests of LMWF-treated and EXT2- or HPSE-siRNA-transfected cells indicated that EXT2 or HPSE expression significantly affect the proangiogenic potential of LMWF. In addition, LMWF increased SDC-1, but decreased SDC-4 expressions. The effect of LMWF depends on SDC-4 expression. Silencing EXT2 or HPSE leads to an increased expression of SDC-4, providing the evidence that EXT2 and HPSE regulate the SDC-4 expression. Altogether, these data indicate that EXT2, HPSE, and SDC-4 are involved in the proangiogenic effects of LMWF, suggesting that the HS metabolism changes linked to LMWF-induced angiogenesis offer the opportunity for new therapeutic strategies of ischemic diseases. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Effects of Low molecular weight fucoidan (LMWF) on endothelial cell abilities: migration and 2D-angiogenesis. Human vascular endothelial cells (HUVEC) were incubated with 10 µg/mL LMWF for 24 h and the migration (<b>A</b>), the lamellipodia formation (<b>B</b>) and the capillary tube formation (length and area) (<b>C</b>) were determined. (<b>A</b>) Migration chamber assay. HUVECs incubated with or without 10 μg/mL LMWF, were allowed to migrate through the porous fibronectin-coated membrane. They were stained with Mayer’s hemalum and counted. The results are expressed as cell number per field; (<b>B</b>) Lamellipodia formation. LMWF induced the formation of lamellipodia and ruffles (white arrows indicate lamellipodia/ruffle formation, DAPI-nucleus (blue), Phalloidin-F-actin (red)). Bar = 10 µm; (<b>C</b>) Capillary tube formation (2D-angiogenesis assay) on Matrigel. Left and right panels show the length (<b>left</b>) and area (<b>right</b>) of endothelial capillaries formed by HUVECs treated with or without 10 µg/mL LMWF. Lower right panel shows a representative image of capillary network, as photographed with phase contrast microscopy (magnification ×100). * <span class="html-italic">p</span> &lt; 0.05 <span class="html-italic">versus</span> control untreated (UT) cells. A.U.: arbitrary unit.</p>
Full article ">Figure 2
<p>LMWF and glycosaminoglycan chain level in HUVECs. (<b>A</b>) Glycosaminoglycan quantification. HUVECs were incubated with 10 µg/mL LMWF for 24 h and the amount of total GAGs, CS and HS chains were determined in the supernatant according to a dimethyl-methylene blue (DMMB) assay; (<b>B</b>) Exostosin-1 and -2 (EXT1 or EXT2) mRNA levels were determined by real-time RT-PCR in cells treated with or without 10 µg/mL LMWF. (<b>C</b>) EXT1 or EXT2 protein levels were determined by western blot in cells treated with or without 10 µg/mL LMWF. Right panel shows a representative image of the western blot assay. * <span class="html-italic">p</span> &lt; 0.05 <span class="html-italic">versus</span> control untreated (UT) cells. A.U.: arbitrary unit.</p>
Full article ">Figure 3
<p>LMWF and heparanase in HUVECs. (<b>A</b>) <span class="html-italic">HPSE</span> mRNA levels were determined by real-time RT-PCR in cells treated with or without 10 µg/mL LMWF; (<b>B</b>) HPSE protein levels were determined by western blot in the supernatant or in the lysate of cells treated with or without 10 µg/mL LMWF. Lower panel shows a representative image of the western blot assay. (<b>C</b>) Heparanase activity was checked in the lysate of LMWF-treated cells. HPSE activity in untreated cells was arbitrary set to 100%. * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.0005, LMWF-treated cells <span class="html-italic">versus</span> LMWF-untreated cells (UT). A.U.: arbitrary unit.</p>
Full article ">Figure 4
<p>Effects of LMWF on the SDC expression in HUVECs. <span class="html-italic">SDC-1</span> and <span class="html-italic">SDC-4</span> mRNA or protein levels in endothelial cells treated or not with 10 µg/mL LMWF were analyzed respectively by real time RT-PCR (<b>A</b>) or western blot (<b>B</b>,<b>C</b>). SDC-1 and SDC-4 ectodomains in the supernatant of cells treated with or without 10 µg/mL LMWF were analyzed by dot blot (<b>D</b>). <span class="html-italic">* p</span> &lt; 0.05, <span class="html-italic">** p</span> &lt; 0.005, significantly different to LMWF-untreated cells (UT). A.U.: arbitrary unit.</p>
Full article ">Figure 5
<p>Effects of LMWF on the SDC distribution in rat balloon injured artery. SDC-1 and SDC-4 expressions were assessed using immunohistochemistry in rat model of intimal hyperplasia. (<b>A</b>) SDC-1 or (<b>D</b>) SDC-4 expressions in non injured arteries; (<b>B</b>) SDC-1 or (<b>E</b>) SDC-4 expressions in injured arteries treated with NaCl; (<b>C</b>) SDC-1 or (<b>F</b>) SDC-4 expressions in injured arteries treated with LMWF. White arrows indicate SDC expressions in the neointima layer in high power view inserts (red). Green: autofluorescence of the elastic fibers of the lamina. Magnification ×100, L: lumen, N: neointima, M: media, A: adventitia.</p>
Full article ">Figure 6
<p>Assessment of EXT involvement in biological effects of LMWF in HUVECs. HUVECs were transfected with <span class="html-italic">EXT1</span>-siRNA or <span class="html-italic">EXT2</span>-siRNA or with <span class="html-italic">SNC</span>-siRNA control. (<b>A</b>) <span class="html-italic">EXT1</span> or <span class="html-italic">EXT2</span> mRNA levels were determined in <span class="html-italic">EXT1</span>-siRNA- or <span class="html-italic">EXT2</span>-siRNA- or <span class="html-italic">SNC</span>-siRNA-transfected control cell by real-time RT PCR. <span class="html-italic">EXT1</span>- or <span class="html-italic">EXT2</span> mRNA level normalized to <span class="html-italic">GAPDH</span> mRNA level in <span class="html-italic">SNC</span>-siRNA-transfected control cells was arbitrarily set to 1; (<b>B</b>) The binding of 10E4 anti-HS antibodies to <span class="html-italic">EXT2</span>- or <span class="html-italic">EXT1</span>-siRNA transfected cells was compared to that of <span class="html-italic">SNC</span>-siRNA-transfected cells; (<b>C</b>) Migration was assayed in cells treated with or without 10 µg/mL LMWF; (<b>D</b>) 2D-angiogenesis was assayed in cells treated with or without 10 µg/mL LMWF. The difference in the capillary network length between LMWF-treated and untreated cells in each RNA interference condition (<span class="html-italic">EXT1 or EXT2</span> silencing) was compared to that in <span class="html-italic">SNC</span>-siRNA-transfected cells. Control LMWF induction was arbitrary set at 100% for <span class="html-italic">SNC</span>-siRNA-transfected cells. <span class="html-italic">* p</span> &lt; 0.05, **<span class="html-italic">* p</span> &lt; 0.0005 <span class="html-italic">versus SNC</span>-siRNA-transfected control cells. A.U.: arbitrary unit.</p>
Full article ">Figure 7
<p>Assessment of HPSE involvement in biological effects of LMWF in HUVECs. HUVECs were transfected with <span class="html-italic">HPSE</span>-siRNA or <span class="html-italic">SNC</span>-siRNA control. (<b>A</b>) <span class="html-italic">HPSE</span> mRNA levels were determined in <span class="html-italic">HPSE</span>-siRNA- or <span class="html-italic">SNC</span>-siRNA-transfected cells by real-time RT-PCR. <span class="html-italic">HPSE</span> mRNA level normalized to <span class="html-italic">GAPDH</span> mRNA level in <span class="html-italic">SNC</span>-siRNA-transfected control cells was arbitrary set to 1; (<b>B</b>) 2D-angiogenesis was assayed in cells treated with or without 10 µg/mL LMWF. The difference in the capillary network length between LMWF-treated and untreated cells in <span class="html-italic">HPSE</span> RNA interference condition was compared to that in <span class="html-italic">SNC</span>-siRNA-transfected cells; (<b>C</b>) Migration was assayed in cells treated with or without 10 µg/mL LMWF. Control LMWF induction was arbitrary set at 100% for <span class="html-italic">SNC</span>-siRNA-transfected cells. <span class="html-italic">* p</span> &lt; 0.05, <span class="html-italic">** p</span> &lt; 0.005 <span class="html-italic">versus SNC</span>-siRNA-transfected control cells. A.U.: arbitrary unit.</p>
Full article ">Figure 8
<p>Assessment of SDC involvement in biological effects of LMWF in HUVECs. 2D-angiogenesis (<b>A</b>) and migration (<b>B</b>) assays were performed in <span class="html-italic">SDC-1</span>-siRNA- or <span class="html-italic">SDC-4</span>-siRNA- or <span class="html-italic">SNC</span>-siRNA-transfected control cells (<b>B</b>, <b>left panel</b>), or in cells treated with specific anti-SDC-1 or anti-SDC-4 antibodies (<b>B</b>, <b>right panel</b>). Control LMWF induction was arbitrary set at 100% for <span class="html-italic">SNC-</span>siRNA-transfected cells or isotypes. (<b>C</b>) <span class="html-italic">SDC-1</span> and <span class="html-italic">SDC-4</span> mRNA expression was analyzed in <span class="html-italic">EXT2</span>- or <span class="html-italic">HPSE</span>-siRNA-transfected cells. <span class="html-italic">* p &lt;</span> 0.05 <span class="html-italic">versus SNC</span>-siRNA-transfected control cells or isotypes. A.U.: arbitrary unit.</p>
Full article ">
951 KiB  
Article
Salt Effect on the Antioxidant Activity of Red Microalgal Sulfated Polysaccharides in Soy-Bean Formula
by Ariela Burg and Levy-Ontman Oshrat
Mar. Drugs 2015, 13(10), 6425-6439; https://doi.org/10.3390/md13106425 - 20 Oct 2015
Cited by 12 | Viewed by 6361
Abstract
Sulfated polysaccharides produced by microalgae, which are known to exhibit various biological activities, may potentially serve as natural antioxidant sources. To date, only a few studies have examined the antioxidant bioactivity of red microalgal polysaccharides. In this research, the effect of different salts [...] Read more.
Sulfated polysaccharides produced by microalgae, which are known to exhibit various biological activities, may potentially serve as natural antioxidant sources. To date, only a few studies have examined the antioxidant bioactivity of red microalgal polysaccharides. In this research, the effect of different salts on the antioxidant activities of two red microalgal sulfated polysaccharides derived from Porphyridium sp. and Porphyridium aerugineum were studied in a soy bean-based infant milk formula. Salt composition and concentration were both shown to affect the polysaccharides’ antioxidant activity. It can be postulated that the salt ions intefer with the polysaccharide chains’ interactions and alter their structure, leading to a new three-dimensional structure that better exposes antiooxidant sites in comparison to the polysaccharide without salt supplement. Among the cations that were studied, Ca2+ had the strongest enhancement effect on antioxidant activities of both polysaccharides. Understanding the effect of salts on polysaccharides’ stucture, in addition to furthering knowledge on polysaccharide bioactivities, may also shed light on the position of the antioxidant active sites. Full article
Show Figures

Figure 1

Figure 1
<p>Antioxidant activity of PS1 and PS2 in soy bean milk formula following addition of KO<sub>2</sub>. C<sub>MDA</sub> was measured by thiobarbitoric acid (TBA) method. All treatments contained 0.5 g formula, 0.35 mM KO<sub>2</sub>, with or without 0.075/0.15% <span class="html-italic">w</span>/<span class="html-italic">v</span> of either PS1 or PS2 polysaccharides. Values are expressed as mean ± SD, <span class="html-italic">n</span> = 3.</p>
Full article ">Figure 2
<p>Polysaccharide antioxidant activity induction by sodium salts. C<sub>MDA</sub> was measured by TBA method. Experimental setup included 0.5 g formula with 0.35 mM KO<sub>2</sub>, 0.15% <span class="html-italic">w</span>/<span class="html-italic">v</span> polysaccharide and various sodium salts at two different concentrations, in a final volume of 10 mL. Values represent the mean of at least 3 repeats; Maximum standard deviation equals 5%.</p>
Full article ">Figure 3
<p>Polysaccharide antioxidant activity induction by potassium salts. C<sub>MDA</sub> was measured by TBA method. Experimental setup included 0.5 g formula with 0.35 mM KO<sub>2</sub>, 0.15% <span class="html-italic">w</span>/<span class="html-italic">v</span> polysaccharide and various potassium salts at two different concentrations, in a final volume of 10 mL. Values represent the mean of at least 3 repeats; Maximum standard deviation equals 8%.</p>
Full article ">Figure 4
<p>Polysaccharide antioxidant activity induction by divalent cation salts. C<sub>MDA</sub> was measured by TBA test. Experimental setup included 0.5 g formula with 0.35 mM KO<sub>2</sub>, 0.15% <span class="html-italic">w</span>/<span class="html-italic">v</span> polysaccharide, and various divalent salts at two different concentrations, in a final volume of 10 mL. Values represent the mean of at least 3 repeats. Maximum standard deviation equals 8%.</p>
Full article ">
548 KiB  
Communication
Immunostimulative Activity of Low Molecular Weight Chitosans in RAW264.7 Macrophages
by Ning Wu, Zheng-Shun Wen, Xing-Wei Xiang, Yan-Na Huang, Yang Gao and You-Le Qu
Mar. Drugs 2015, 13(10), 6210-6225; https://doi.org/10.3390/md13106210 - 30 Sep 2015
Cited by 53 | Viewed by 7058
Abstract
Chitosan and its derivatives such as low molecular weight chitosans (LMWCs) have been reported to exert many biological activities, such as antioxidant and antitumor effects. However, complex and molecular weight dependent effects of chitosan remain controversial and the mechanisms that mediate these complex [...] Read more.
Chitosan and its derivatives such as low molecular weight chitosans (LMWCs) have been reported to exert many biological activities, such as antioxidant and antitumor effects. However, complex and molecular weight dependent effects of chitosan remain controversial and the mechanisms that mediate these complex effects are still poorly defined. This study was carried out to investigate the immunostimulative effect of different molecular weight chitosan in RAW264.7 macrophages. Our data suggested that two LMWCs (molecular weight of 3 kDa and 50 kDa) both possessed immunostimulative activity, which was dependent on dose and, at the higher doses, also on the molecular weight. LMWCs could significantly enhance the the pinocytic activity, and induce the production of tumor necrosis factor α (TNF-α), interleukin 6 (IL-6), interferon-γ (IFN-γ), nitric oxide (NO) and inducible nitric oxide synthase (iNOS) in a molecular weight and concentration-dependent manner. LMWCs were further showed to promote the expression of the genes including iNOS, TNF-α. Taken together, our findings suggested that LMWCs elicited significantly immunomodulatory response through up-regulating mRNA expression of proinflammatory cytokines and activated RAW264.7 macrophage in a molecular weight- and concentration-dependent manner. Full article
Show Figures

Figure 1

Figure 1
<p>Effects of low molecular weight chitosans (LMWCs) on the cell viability of RAW264.7 macrophage. Each cell population (2 × 10<sup>4</sup> cells/well) was treated with LMWCs (3 kDa and 50 kDa) at the indicated concentrations of 2.5, 10 and 40 μg/mL for 24 h, respectively. Values are means ± SD (<span class="html-italic">n</span> = 3). Bars with no letters are not statistically different (<span class="html-italic">p</span> &gt; 0.05).</p>
Full article ">Figure 2
<p>Effects of LMWCs on the pinocytic activity of RAW264.7 macrophage. Each cell population (1 × 10<sup>4</sup> cells/well) was treated with LMWCs (3 kDa and 50 kDa) at the indicated concentrations of 2.5, 10 and 40 μg/mL for 24 h, respectively. Values are means ± SD (<span class="html-italic">n</span> = 3). Bars with different letters (a, b, c, d, e, f) are statistically different (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>Effects of LMWCs on the production of tumor necrosis factor α (TNF-α) (<b>A</b>), interferon-γ (IFN-γ) (<b>B</b>) and interleukin 6 (IL-6) (<b>C</b>) from RAW264.7 macrophage. Each cell population (1 × 10<sup>6</sup> cells/mL) was treated with LMWCs (3 kDa and 50 kDa) at the indicated concentrations of 2.5, 10 and 40 μg/mL for 24 h, respectively. Values are means ± SD (<span class="html-italic">n</span> = 3). Bars with different letters (a, b, c, d, e, f) are statistically different (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3 Cont.
<p>Effects of LMWCs on the production of tumor necrosis factor α (TNF-α) (<b>A</b>), interferon-γ (IFN-γ) (<b>B</b>) and interleukin 6 (IL-6) (<b>C</b>) from RAW264.7 macrophage. Each cell population (1 × 10<sup>6</sup> cells/mL) was treated with LMWCs (3 kDa and 50 kDa) at the indicated concentrations of 2.5, 10 and 40 μg/mL for 24 h, respectively. Values are means ± SD (<span class="html-italic">n</span> = 3). Bars with different letters (a, b, c, d, e, f) are statistically different (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Effect of LMWCs on Nitric oxide (NO) production and activities of inducible nitric oxide synthase (iNOS) in RAW264.7 macrophage. Each cell population (1 × 10<sup>6</sup> cells/mL) was treated with LMWCs (3 kDa and 50 kDa) at the indicated concentrations of 2.5, 10 and 40 μg/mL for 24 h, respectively. Values are means ± SD (<span class="html-italic">n</span> = 3). Bars with different letters (a, b, c, d, e, f) are statistically different (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Effect of LMWCs on the mRNA expression levels of TNF-α and iNOS in RAW264.7 macrophage. Each cell population (1 × 10<sup>6</sup> cells/mL) was treated with LMWCs (3 kDa and 50 kDa) at the indicated concentrations of 2.5, 10 and 40 μg/mL for 24 h, respectively. Values are means ± SD (<span class="html-italic">n</span> = 3). Bars with different letters (a, b, c, d, e, f) are statistically different (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
2977 KiB  
Article
Alginate-Derived Oligosaccharide Inhibits Neuroinflammation and Promotes Microglial Phagocytosis of β-Amyloid
by Rui Zhou, Xu-Yang Shi, De-Cheng Bi, Wei-Shan Fang, Gao-Bin Wei and Xu Xu
Mar. Drugs 2015, 13(9), 5828-5846; https://doi.org/10.3390/md13095828 - 16 Sep 2015
Cited by 71 | Viewed by 9121
Abstract
Alginate from marine brown algae has been widely applied in biotechnology. In this work, the effects of alginate-derived oligosaccharide (AdO) on lipopolysaccharide (LPS)/β-amyloid (Aβ)-induced neuroinflammation and microglial phagocytosis of Aβ were studied. We found that pretreatment of BV2 microglia with AdO prior to [...] Read more.
Alginate from marine brown algae has been widely applied in biotechnology. In this work, the effects of alginate-derived oligosaccharide (AdO) on lipopolysaccharide (LPS)/β-amyloid (Aβ)-induced neuroinflammation and microglial phagocytosis of Aβ were studied. We found that pretreatment of BV2 microglia with AdO prior to LPS/Aβ stimulation led to a significant inhibition of production of nitric oxide (NO) and prostaglandin E2 (PGE2), expression of inducible nitric oxide synthase (iNOS) and cyclooxygenase-2 (COX-2) and secretion of proinflammatory cytokines. We further demonstrated that AdO remarkably attenuated the LPS-activated overexpression of toll-like receptor 4 (TLR4) and nuclear factor (NF)-κB in BV2 cells. In addition to the impressive inhibitory effect on neuroinflammation, we also found that AdO promoted the phagocytosis of Aβ through its interaction with TLR4 in microglia. Our results suggested that AdO exerted the inhibitory effect on neuroinflammation and the promotion effect on microglial phagocytosis, indicating its potential as a nutraceutical or therapeutic agent for neurodegenerative diseases, particularly Alzheimer’s disease (AD). Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic representation of chemical structures of alginate-derived oligosaccharide (AdO) prepared by oxidative degradation.</p>
Full article ">Figure 2
<p>AdO reduced the lipopolysaccharide (LPS)-activated production of nitric oxide (NO) and prostaglandin E<sub>2</sub> (PGE<sub>2</sub>) and the expression of inducible nitric oxide synthase (iNOS) and cyclooxygenase-2 (COX-2) in BV2 microglial cells. (<b>A</b>) Cell viability was evaluated using the CCK-8 assay for control cells, LPS (0.5 μg/mL) treatment cells, LPS with AdO (100–1000 μg/mL) treatment cells, and AdO (50–1000 μg/mL) treatment cells; (<b>B</b>) The nitrite concentration was measured as an indicator of NO production using the Griess reagent; (<b>C</b>) PGE<sub>2</sub> production was analyzed using ELISA; (<b>D</b>) The expression of the iNOS and COX-2 mRNAs was detected by RT-PCR; (<b>E</b>) The relative mRNA levels of iNOS and COX-2 were analyzed with reference to the control group; (<b>F</b>) The expression of the iNOS and COX-2 proteins was detected using Western blot analysis; (<b>G</b>) The relative levels of the iNOS and COX-2 proteins were analyzed with reference to the control group. The data are presented as the mean ± SD for three independent experiments. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001 indicate significant differences compared with the LPS-treated group.</p>
Full article ">Figure 2 Cont.
<p>AdO reduced the lipopolysaccharide (LPS)-activated production of nitric oxide (NO) and prostaglandin E<sub>2</sub> (PGE<sub>2</sub>) and the expression of inducible nitric oxide synthase (iNOS) and cyclooxygenase-2 (COX-2) in BV2 microglial cells. (<b>A</b>) Cell viability was evaluated using the CCK-8 assay for control cells, LPS (0.5 μg/mL) treatment cells, LPS with AdO (100–1000 μg/mL) treatment cells, and AdO (50–1000 μg/mL) treatment cells; (<b>B</b>) The nitrite concentration was measured as an indicator of NO production using the Griess reagent; (<b>C</b>) PGE<sub>2</sub> production was analyzed using ELISA; (<b>D</b>) The expression of the iNOS and COX-2 mRNAs was detected by RT-PCR; (<b>E</b>) The relative mRNA levels of iNOS and COX-2 were analyzed with reference to the control group; (<b>F</b>) The expression of the iNOS and COX-2 proteins was detected using Western blot analysis; (<b>G</b>) The relative levels of the iNOS and COX-2 proteins were analyzed with reference to the control group. The data are presented as the mean ± SD for three independent experiments. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001 indicate significant differences compared with the LPS-treated group.</p>
Full article ">Figure 3
<p>AdO inhibited the LPS/Aβ-activated secretion of proinflammatory cytokines. (<b>A</b>) Tumor necrosis factor-α (TNF-α), interleukin (IL)-1β and IL-6 expression in the LPS-activated BV2 cells was measured by ELISA; (<b>B</b>) TNF-α, IL-6 and IL-12 production in the Aβ-activated BV2 cells was evaluated by ELISA; (<b>C</b>) The cell viability following AdO (500 μg/mL) pretreatment with or without Aβ (10 μM) stimulation was detected by the CCK-8 assay. The data are presented as the mean ± SD for three independent experiments. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001 indicate significant differences compared with the LPS/Aβ-treated group, and ### <span class="html-italic">p</span> &lt; 0.001 indicates a significant difference between the control group and the Aβ-treated group.</p>
Full article ">Figure 4
<p>AdO suppressed the LPS-induced toll-like receptor 4 (TLR4) expression and nuclear factor (NF)-κB activation. TLR4 expression was evaluated using immunofluorescence analysis (<b>A</b>) and Western blot analysis (<b>B</b>); (<b>C</b>) NF-κB p65 expression was examined using immunofluorescence analysis; (<b>D</b>) The cytoplasm and nuclear proteins were extracted and the NF-κB p65 protein was analyzed by Western blot analysis. The immunofluorescence analysis was carried out by laser scanning confocal microscopy (60×), and the images were processed using ImageJ software. The immunofluorescence analysis and Western blot analysis were performed in three independent experiments.</p>
Full article ">Figure 5
<p>AdO inhibited the morphological changes induced by LPS/β-amyloid (Aβ) in the BV2 cells. (<b>A</b>) The cellular morphology of the control group, AdO (50–500 μg/mL)-treated group, LPS (0.5 μg/mL)-treated group and the group treated with AdO prior to LPS are shown in the dark field images; (<b>B</b>) The percentage of cells exhibiting the activated morphology was statistically analyzed; (<b>C</b>) The cellular morphology of the control group, AdO (500 μg/mL)-treated group, and Aβ (10 μM)-treated group as well as the group treated with AdO prior to Aβ are shown in the dark field images; (<b>D</b>) The percentage of cells exhibiting the activated morphology was statistically analyzed. The cellular morphology was observed using dark-field microscopy (40×), and the images were analyzed using ImageJ software (National Institutes of Health, Bethesda, MD, USA). The normal cell morphology is indicated by white arrows, and the activated cell morphology is indicated by white dotted arrows. Scale bar = 20 μm. The images were from three independent experiments. *** <span class="html-italic">p</span> &lt; 0.001 indicates significant differences between the control group and the LPS/Aβ-treated group; ### <span class="html-italic">p</span> &lt; 0.001 indicates significant differences between the LPS/Aβ-treated group and the group treated with AdO prior to LPS/Aβ.</p>
Full article ">Figure 6
<p>AdO promoted the phagocytosis of gold nanoparticles (AuNPs) and Hilyte Fluo™ 488-labled β-amyloid (1-42) (FL-Aβ) in the BV2 cells. (<b>A</b>) The cells were treated with 1 pM of the AuNPs for indicated incubation times; (<b>B</b>) The phagocytosis of the AuNPs by the BV2 cells was evaluated by counting the number of AuNPs in fifty cells using ImageJ software; (<b>C</b>) The cells were treated with AdO (50–500 μg/mL) for 20 h and then incubated with 1 pM of the AuNPs for 1 h. (<b>D</b>) The average number of phagocytosed AuNPs in fifty cells was analyzed using ImageJ software. The cells with the accumulated AuNPs were examined using dark-field microscopy (40×), scale bar = 20 μm; (<b>E</b>) The cells were treated with AdO (50 μg/mL) for 20 h and then incubated with 500 nM FL-Aβ for 4 h. The cells with the accumulated FL-Aβ were examined using laser scanning confocal microscopy (60×), and the morphology of the cells was shown using differential interference contrast (DIC) images, scale bar = 20 μm; (<b>F</b>) The average fluorescent intensity of fifty cells was evaluated using ImageJ software. The microscopic images were from three independent experiments. *** <span class="html-italic">p</span> &lt; 0.001 indicates significant differences compared with the control group.</p>
Full article ">Figure 6 Cont.
<p>AdO promoted the phagocytosis of gold nanoparticles (AuNPs) and Hilyte Fluo™ 488-labled β-amyloid (1-42) (FL-Aβ) in the BV2 cells. (<b>A</b>) The cells were treated with 1 pM of the AuNPs for indicated incubation times; (<b>B</b>) The phagocytosis of the AuNPs by the BV2 cells was evaluated by counting the number of AuNPs in fifty cells using ImageJ software; (<b>C</b>) The cells were treated with AdO (50–500 μg/mL) for 20 h and then incubated with 1 pM of the AuNPs for 1 h. (<b>D</b>) The average number of phagocytosed AuNPs in fifty cells was analyzed using ImageJ software. The cells with the accumulated AuNPs were examined using dark-field microscopy (40×), scale bar = 20 μm; (<b>E</b>) The cells were treated with AdO (50 μg/mL) for 20 h and then incubated with 500 nM FL-Aβ for 4 h. The cells with the accumulated FL-Aβ were examined using laser scanning confocal microscopy (60×), and the morphology of the cells was shown using differential interference contrast (DIC) images, scale bar = 20 μm; (<b>F</b>) The average fluorescent intensity of fifty cells was evaluated using ImageJ software. The microscopic images were from three independent experiments. *** <span class="html-italic">p</span> &lt; 0.001 indicates significant differences compared with the control group.</p>
Full article ">Figure 7
<p>The involvement of TLR4 in the AdO-promoted phagocytosis of Aβ in the BV2 cells. (<b>A</b>) The cells were treated with anti-TLR4 (10 μg/mL) for 1 h prior to AdO treatment for 20 h and then incubated with 500 nM FL-Aβ for 4 h. The lysosomes and nuclei of the cells were stained with LysoTracker Red DND-99 and 4′,6-diamidino-2-phenylindole (DAPI), respectively. The cells with the accumulated FL-Aβ were observed using laser scanning confocal microscopy (60×), scale bar = 20 μm; (<b>B</b>) The average fluorescence intensity of fifty cells in each group was evaluated using ImageJ software; (<b>C</b>) The phagocytosis of FL-Aβ in untreated cells, AdO-treated cells, and anti-TLR4 and AdO-treated cells was analyzed using flow cytometry. The flow cytometric plots show the M1 region of the phagocytic cell populations based on fluorescence intensity; (<b>D</b>) The mean fluorescence intensity (MFI) of the cells in the M1 region is shown. The images were from three independent experiments. ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001 indicate significant differences between the control group and the AdO-treated group, and ## <span class="html-italic">p</span> &lt; 0.01 and ### <span class="html-italic">p</span> &lt; 0.001 indicate significant differences between the AdO-treated group and the AdO combined with anti-TLR4-treated group.</p>
Full article ">
1074 KiB  
Article
The Mucus of Actinia equina (Anthozoa, Cnidaria): An Unexplored Resource for Potential Applicative Purposes
by Loredana Stabili, Roberto Schirosi, Maria Giovanna Parisi, Stefano Piraino and Matteo Cammarata
Mar. Drugs 2015, 13(8), 5276-5296; https://doi.org/10.3390/md13085276 - 19 Aug 2015
Cited by 56 | Viewed by 9339
Abstract
The mucus produced by many marine organisms is a complex mixture of proteins and polysaccharides forming a weak watery gel. It is essential for vital processes including locomotion, navigation, structural support, heterotrophic feeding and defence against a multitude of environmental stresses, predators, parasites, [...] Read more.
The mucus produced by many marine organisms is a complex mixture of proteins and polysaccharides forming a weak watery gel. It is essential for vital processes including locomotion, navigation, structural support, heterotrophic feeding and defence against a multitude of environmental stresses, predators, parasites, and pathogens. In the present study we focused on mucus produced by a benthic cnidarian, the sea anemone Actinia equina (Linnaeus, 1758) for preventing burial by excess sedimentation and for protection. We investigated some of the physico-chemical properties of this matrix such as viscosity, osmolarity, electrical conductivity, protein, carbohydrate, and total lipid contents. Some biological activities such as hemolytic, cytotoxic, and antibacterial lysozyme-like activities were also studied. The A. equina mucus is mainly composed by water (96.2% ± 0.3%), whereas its dry weight is made of 24.2% ± 1.3% proteins and 7.8% ± 0.2% carbohydrates, with the smallest and largest components referable to lipids (0.9%) and inorganic matter (67.1%). The A. equina mucus matrix exhibited hemolytic activity on rabbit erythrocytes, cytotoxic activity against the tumor cell line K562 (human erythromyeloblastoid leukemia) and antibacterial lysozyme-like activity. The findings from this study improve the available information on the mucus composition in invertebrates and have implications for future investigations related to exploitation of A. equina and other sea anemones’ mucus as a source of bioactive compounds of high pharmaceutical and biotechnological interest. Full article
Show Figures

Figure 1

Figure 1
<p><span class="html-italic">Actinia equina</span> mucus composition: (<b>A</b>) water content and dried weight; (<b>B</b>) organic and inorganic residuals.</p>
Full article ">Figure 2
<p>SDS-PAGE analysis of <span class="html-italic">Actinia equina</span> mucus. Panel <b>A</b>: Molecular weight standards furnished by Fermentas. Molecular weights (kDa) of standard proteins are on left; Panel <b>B</b>: <span class="html-italic">A. equina</span> total mucus; (<b>C</b>) <span class="html-italic">Actinia equina</span> different molecular weight fractions from total mucus extract obtained by membrane filtration system (pore size: 10 kDa). SDS-PAGE 15% acrylamide gel stained with Coomassie Blue R-250. Lane 1: Fraction &gt;10 kDa named “U” (Upper), Lane 2: Standard Low sigma, Lane 3: Fraction &lt;10 kDa. Named “D” (Lower); (<b>D</b>) Micro plate lysis assay carried out against Rabbit erythrocytes (RRBCs) in TBS buffer. Hemolysis is evidenced by free hemoglobin, when the erythrocytes are not lysed a central pellet of erythrocytes is visible on the well center. Lower fraction (<b>D</b>) showing lysis until dilution of 1:64, Upper fraction (U) showing lysis until dilution of 1:2048, Control experiment (Ce) with RRBCs and buffer.</p>
Full article ">Figure 3
<p>Lysozyme-like activity of <span class="html-italic">Actinia equina</span> mucus. (<b>A</b>) Standard assay on Petri dish inoculated with <span class="html-italic">Micrococcus lysodeikticus</span> cell walls to detect the lysozyme-like activity of <span class="html-italic">A. equina</span> mucus; (<b>B</b>) Effect of the pH on the lysozyme-like activity of <span class="html-italic">A. equina</span> mucus. Columns are mean values (<span class="html-italic">n</span> = 20) (vertical bars ± Standard Deviation); (<b>C</b>) Effect of the ionic strength on the lysozyme-like activity of mucus. Columns are mean values (<span class="html-italic">n</span> = 20) (vertical bars ± Standard Deviation); (<b>D</b>) Effect of the incubation temperature on the lysozyme-like activity of mucus. Columns are mean values (<span class="html-italic">n</span> = 20) (vertical bars ± Standard Deviation).</p>
Full article ">Figure 4
<p>(<b>A</b>) Light Microscopic observation of Human erythromyeloblastoid leukemia (K562) cells treated with <span class="html-italic">A. equine</span> mucus crude extract. The target cell lysis was also determined by trypan blue exclusion test. Bar: 25 μm; (<b>B</b>) Control cell observed in the absence of mucus Bar. 25 μm; (<b>C</b>) Colorimetric assay of <span class="html-italic">A. equina</span> mucus extract on human chronic myelogenous leukemia cells K562 (Cytotoxic detection Kit. Boehringer Mannheim, Mannheim, Germany). Lactate dehydrogenase release into the supernatant was used to calculate the percentage of target cell lysis.</p>
Full article ">Figure 5
<p>(<b>A</b>) High performance liquid chromatography separation of <span class="html-italic">A. equina</span> mucus components. The first plot shows profile of HPLC analysis of the crude mucus extract. Green arrows 1 and 2 indicate the isolated peaks at 12.5 and 14.5 min. Insert shows HPLC profiles of bovine serum albumin (BSA-66 kDa), chimotrypsinogen (25 kDa) and ribonuclease (13.7 kDa) used as standards separated on a molecular weight exclusion column BioSuite 250 (10 microns; Waters, Milford, CT, USA). The second plot shows the purification profiles of the high molecular weight fraction (u = upper) and low molecular weight fraction (d = lower) previously separated via centrifugation system on 10 kDa membrane. Red arrow indicate peak 2 detected at 14.5 min post HPLC start running; (<b>B</b>) Lytic activity detected in microplate toward rabbit erythrocytes of peaks 1 and 2 (Ce: Control experiment).</p>
Full article ">
411 KiB  
Review
Chitosan: An Update on Potential Biomedical and Pharmaceutical Applications
by Randy Chi Fai Cheung, Tzi Bun Ng, Jack Ho Wong and Wai Yee Chan
Mar. Drugs 2015, 13(8), 5156-5186; https://doi.org/10.3390/md13085156 - 14 Aug 2015
Cited by 977 | Viewed by 31580
Abstract
Chitosan is a natural polycationic linear polysaccharide derived from chitin. The low solubility of chitosan in neutral and alkaline solution limits its application. Nevertheless, chemical modification into composites or hydrogels brings to it new functional properties for different applications. Chitosans are recognized as [...] Read more.
Chitosan is a natural polycationic linear polysaccharide derived from chitin. The low solubility of chitosan in neutral and alkaline solution limits its application. Nevertheless, chemical modification into composites or hydrogels brings to it new functional properties for different applications. Chitosans are recognized as versatile biomaterials because of their non-toxicity, low allergenicity, biocompatibility and biodegradability. This review presents the recent research, trends and prospects in chitosan. Some special pharmaceutical and biomedical applications are also highlighted. Full article
Show Figures

Figure 1

Figure 1
<p>Scopus indexed publications related to chitosan and its derivatives.</p>
Full article ">Figure 2
<p>A schematic presentation of chitosan preparation from raw materials.</p>
Full article ">Figure 3
<p>Scavenging effects of <span class="html-italic">N</span>-CECS with different degrees of substitution toward ABTS radicals. Reprinted with permission from [<a href="#B75-marinedrugs-13-05156" class="html-bibr">75</a>] Copyright (2015) American Chemical Society.</p>
Full article ">Figure 4
<p>A diagrammatic presentation of how the chitosan-based wound dressing works.</p>
Full article ">
466 KiB  
Article
Laminarin from Irish Brown Seaweeds Ascophyllum nodosum and Laminaria hyperborea: Ultrasound Assisted Extraction, Characterization and Bioactivity
by Shekhar U. Kadam, Colm P. O'Donnell, Dilip K. Rai, Mohammad B. Hossain, Catherine M. Burgess, Des Walsh and Brijesh K. Tiwari
Mar. Drugs 2015, 13(7), 4270-4280; https://doi.org/10.3390/md13074270 - 10 Jul 2015
Cited by 216 | Viewed by 13134
Abstract
Ultrasound assisted extraction (UAE), purification, characterization and antioxidant activity of laminarin from Irish brown seaweeds Ascophyllum nodosum and Laminarina hyperborea were investigated. UAE was carried out using 60% ultrasonic power amplitude and 0.1 M hydrochloric acid for 15 min. Separately, solid-liquid extraction was [...] Read more.
Ultrasound assisted extraction (UAE), purification, characterization and antioxidant activity of laminarin from Irish brown seaweeds Ascophyllum nodosum and Laminarina hyperborea were investigated. UAE was carried out using 60% ultrasonic power amplitude and 0.1 M hydrochloric acid for 15 min. Separately, solid-liquid extraction was carried in an orbital shaker using 0.1 M hydrochloric acid at 70 °C for 2.5 h. UAE with hydrochloric acid resulted in the highest concentration of laminarin, 5.82% and 6.24% on dry weight basis from A. nodosum and L. hyperborea, respectively. Purification of all extracts was carried out using molecular weight cut off dialysis at 10 kDa. Characterization of the laminarin fraction was carried out using matrix assisted laser desorption/ionization time-of-flight mass spectrometry. Antioxidant activity of A. nodosum and L. hyperborea extracts had 2,2-diphenyl-1-picrylhydrazyl (DPPH) inhibition levels of 93.23% and 87.57%, respectively. Moreover, these extracts have shown inihibition of bacterial growth of Staphylcoccus aureus, Listeria monocytogenes, Escherichia coli and Salmonella typhimurium. Full article
Show Figures

Figure 1

Figure 1
<p>Matrix Assisted Laser Desorption Ionization Quadrupole Time-of-Flight Mass Spectrometry (MALDI-Q-TOF-MS) spectra showing the distribution of molecular weights of (<b>A</b>) commercial laminarin in sinapinic acid and (<b>B</b>) <span class="html-italic">Laminaria hyperborea</span> extract obtained by ultrasound assisted extraction using 0.1 M HCl as a solvent.</p>
Full article ">
1879 KiB  
Article
Fucoidan Stimulates Monocyte Migration via ERK/p38 Signaling Pathways and MMP9 Secretion
by Elene Sapharikas, Anna Lokajczyk, Anne-Marie Fischer and Catherine Boisson-Vidal
Mar. Drugs 2015, 13(7), 4156-4170; https://doi.org/10.3390/md13074156 - 30 Jun 2015
Cited by 18 | Viewed by 7798
Abstract
Critical limb ischemia (CLI) induces the secretion of paracrine signals, leading to monocyte recruitment and thereby contributing to the initiation of angiogenesis and tissue healing. We have previously demonstrated that fucoidan, an antithrombotic polysaccharide, promotes the formation of new blood vessels in a [...] Read more.
Critical limb ischemia (CLI) induces the secretion of paracrine signals, leading to monocyte recruitment and thereby contributing to the initiation of angiogenesis and tissue healing. We have previously demonstrated that fucoidan, an antithrombotic polysaccharide, promotes the formation of new blood vessels in a mouse model of hindlimb ischemia. We examined the effect of fucoidan on the capacity of peripheral blood monocytes to adhere and migrate. Monocytes negatively isolated with magnetic beads from peripheral blood of healthy donors were treated with fucoidan. Fucoidan induced a 1.5-fold increase in monocyte adhesion to gelatin (p < 0.05) and a five-fold increase in chemotaxis in Boyden chambers (p < 0.05). Fucoidan also enhanced migration 2.5-fold in a transmigration assay (p < 0.05). MMP9 activity in monocyte supernatants was significantly enhanced by fucoidan (p < 0.05). Finally, Western blot analysis of fucoidan-treated monocytes showed upregulation of ERK/p38 phosphorylation. Inhibition of ERK/p38 phosphorylation abrogated fucoidan enhancement of migration (p < 0.01). Fucoidan displays striking biological effects, notably promoting monocyte adhesion and migration. These effects involve the ERK and p38 pathways, and increased MMP9 activity. Fucoidan could improve critical limb ischemia by promoting monocyte recruitment. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Fucoidan enhances monocyte adhesion to gelatin, and their migration. (<b>A</b>) Representative results obtained with PBMC after 30 min, with or without 24 h of fucoidan pretreatment; (<b>B</b>) Monocyte adhesion (in white, control monocytes, in black, monocytes incubated with 10 μg/mL fucoidan; (<b>C</b>) Representative results for migration of isolated PBMC treated with or without fucoidan (4 h) towards 100 ng/mL MCP-1; (<b>D</b>) Migratory cell numbers in five independent fields; (<b>E</b>) Representative monocyte transmigration (18 h) with or without fucoidan pretreatment (30 min); (<b>F</b>) Transmigratory cell numbers in five independent fields. Three to five independent donors. <b>*</b> <span class="html-italic">p</span> &lt; 0.05; <b>**</b> <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 2
<p>Impact of fucoidan on adhesion molecule and CCR2 receptor expression: PBMC were treated for 30 min or 24 h with (black bars) or without fucoidan (white bars). (<b>A</b>) Percentage of monocytes positive for alpha M expression; (<b>B</b>) Percentage of beta 2-positive cells; (<b>C</b>) Percentage of alpha 4-positive cells; (<b>D</b>) Percentage of beta 1-positive cells; (<b>E</b>) Percentage of CCR2-positive cells (4 independent donors).</p>
Full article ">Figure 3
<p>Impact of fucoidan on monocyte MMP9 expression (gelatinolytic activity): (<b>A</b>) Representative gelatin zymography of culture supernatant of monocytes treated with or without fucoidan for 30 min; (<b>B</b>) MMP9 gelanitolytic activity; (<b>C</b>) MMP2 gelatinolytic activity. Three independent donors; <b>*</b> <span class="html-italic">p</span> &lt; 0.05</p>
Full article ">Figure 4
<p>ERK and p38 signaling pathway involvement in fucoidan-treated monocyte migration: (<b>A</b>) Representative Western blot illustrating phosphorylation of ERK1/2 and P38 when PBMC were treated with or without fucoidan (in the presence or absence of PD98059 or SB203580) for 30 min; (<b>B</b>) Quantitative analysis of ERK phosphorylation; (<b>C</b>) Quantitative analysis of p38 phosphorylation. Results are represented relative to the corresponding control, with with independent donors; (<b>D</b>) Representative fields showing migratory cells treated as in A; (<b>E</b>) Migratory cell numbers in five independent fields. <b>*</b> <span class="html-italic">p</span> &lt; 0.05; <b>**</b> <span class="html-italic">p</span> &lt; 0.01; <b>***</b> <span class="html-italic">p</span> &lt; 0.001 compared to control. Four independent donors.</p>
Full article ">Figure 5
<p>Schematic overview of the effect of fucoidan on monocyte migration. Fucoidan bound to the cell membrane enhances MCP-1 interaction with its receptor CCR2. This interaction leads to phosphorylation of ERK1/2 and p38 and activates MM9 secretion. PD98059 and SB203580 inhibit this phosphorylation, leading to reduced monocyte migration.</p>
Full article ">
619 KiB  
Review
Seaweed Hydrocolloid Production: An Update on Enzyme Assisted Extraction and Modification Technologies
by Nanna Rhein-Knudsen, Marcel Tutor Ale and Anne S. Meyer
Mar. Drugs 2015, 13(6), 3340-3359; https://doi.org/10.3390/md13063340 - 27 May 2015
Cited by 258 | Viewed by 18682
Abstract
Agar, alginate, and carrageenans are high-value seaweed hydrocolloids, which are used as gelation and thickening agents in different food, pharmaceutical, and biotechnological applications. The annual global production of these hydrocolloids has recently reached 100,000 tons with a gross market value just above US$ [...] Read more.
Agar, alginate, and carrageenans are high-value seaweed hydrocolloids, which are used as gelation and thickening agents in different food, pharmaceutical, and biotechnological applications. The annual global production of these hydrocolloids has recently reached 100,000 tons with a gross market value just above US$ 1.1 billion. The techno-functional properties of the seaweed polysaccharides depend strictly on their unique structural make-up, notably degree and position of sulfation and presence of anhydro-bridges. Classical extraction techniques include hot alkali treatments, but recent research has shown promising results with enzymes. Current methods mainly involve use of commercially available enzyme mixtures developed for terrestrial plant material processing. Application of seaweed polysaccharide targeted enzymes allows for selective extraction at mild conditions as well as tailor-made modifications of the hydrocolloids to obtain specific functionalities. This review provides an update of the detailed structural features of κ-, ι-, λ-carrageenans, agars, and alginate, and a thorough discussion of enzyme assisted extraction and processing techniques for these hydrocolloids. Full article
Show Figures

Figure 1

Figure 1
<p>Conversion of the pre cursors μ- and ν-carrageenan into κ- and ι-carrageenan.</p>
Full article ">Figure 2
<p>The gelation mechanism of κ-carrageenan in the presence of potassium ions.</p>
Full article ">
3480 KiB  
Article
Production of Chondroitin Sulphate from Head, Skeleton and Fins of Scyliorhinus canicula By-Products by Combination of Enzymatic, Chemical Precipitation and Ultrafiltration Methodologies
by María Blanco, Javier Fraguas, Carmen G. Sotelo, Ricardo I. Pérez-Martín and José Antonio Vázquez
Mar. Drugs 2015, 13(6), 3287-3308; https://doi.org/10.3390/md13063287 - 27 May 2015
Cited by 37 | Viewed by 7305
Abstract
This study illustrates the optimisation of the experimental conditions of three sequential steps for chondroitin sulphate (CS) recovery from three cartilaginous materials of Scyliorhinus canicula by-products. Optimum conditions of temperature and pH were first obtained for alcalase proteolysis of head cartilage (58 °C/pH [...] Read more.
This study illustrates the optimisation of the experimental conditions of three sequential steps for chondroitin sulphate (CS) recovery from three cartilaginous materials of Scyliorhinus canicula by-products. Optimum conditions of temperature and pH were first obtained for alcalase proteolysis of head cartilage (58 °C/pH 8.5/0.1% (v/w)/10 h of hydrolysis). Then, similar optimal conditions were observed for skeletons and fin materials. Enzymatic hydrolysates were subsequently treated with a combination of alkaline hydroalcoholic saline solutions in order to improve the protein hydrolysis and the selective precipitation of CS. Ranges of 0.53–0.64 M (NaOH) and 1.14–1.20 volumes (EtOH) were the levels for optimal chemical treatment depending on the cartilage origin. Finally, selective purification and concentration of CS and protein elimination of samples obtained from chemical treatment, was assessed by a combination of ultrafiltration and diafiltration (UF-DF) techniques at 30 kDa. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Kinetics of cartilage hydrolysis from <span class="html-italic">Scyliorhinus canicula</span> heads using alcalase in each one of the experimental conditions defined in <a href="#marinedrugs-13-03287-t001" class="html-table">Table 1</a>. The experimental data (symbols) were fitted to the Weibull Equation (4) (continuous line).</p>
Full article ">Figure 2
<p>Predicted response surfaces by empirical equations summarized in <a href="#marinedrugs-13-03287-t003" class="html-table">Table 3</a> corresponding to the combined effect of pH and <span class="html-italic">T</span> on the different dependent variables evaluated for the study of head-cartilages proteolysis by alcalase.</p>
Full article ">Figure 3
<p>Enzymatic hydrolysis at two pH levels for different cartilages from <span class="html-italic">S. canicula</span> wastes (left). To the right, long hydrolysis at the best pH selected are additionally shown. Experimental data were fitted to the Weibull Equation (4). (<b>A</b>) Fins; (<b>B</b>) Heads and (<b>C</b>) Skeletons.</p>
Full article ">Figure 4
<p>Predicted response surfaces by empirical equations summarized in <a href="#marinedrugs-13-03287-t007" class="html-table">Table 7</a> corresponding to the combined effect of NaOH and EtOH on the selective treatment of CS from hydrolysate cartilages of <span class="html-italic">S. canicula.</span></p>
Full article ">Figure 5
<p>UF-DF process for CS purification from <span class="html-italic">S. canicula</span> cartilages of three origins at 30 kDa. Top: Concentration of retained protein (○) and CS (●) in linear relation with the factor of volumetric concentration (fc) showing experimental data (points) and theoretical profiles corresponding to a completely retained solute (discontinuous line). Bottom: Progress of protein (○) and CS (●) retention with the increase of diavolume from DF process (D). Equation (6) was used to fit the experimental data. Error bars are the confidence intervals (α = 0.05; <span class="html-italic">n</span> = 2).</p>
Full article ">
920 KiB  
Article
Antidiabetic Activity of Differently Regioselective Chitosan Sulfates in Alloxan-Induced Diabetic Rats
by Ronge Xing, Xiaofei He, Song Liu, Huahua Yu, Yukun Qin, Xiaolin Chen, Kecheng Li, Rongfeng Li and Pengcheng Li
Mar. Drugs 2015, 13(5), 3072-3090; https://doi.org/10.3390/md13053072 - 15 May 2015
Cited by 22 | Viewed by 5900
Abstract
The present study investigated and compared the hypoglycemic activity of differently regioselective chitosan sulfates in alloxan-induced diabetic rats. Compared with the normal control rats, significantly higher blood glucose levels were observed in the alloxan-induced diabetic rats. The differently regioselective chitosan sulfates exhibited hypoglycemic [...] Read more.
The present study investigated and compared the hypoglycemic activity of differently regioselective chitosan sulfates in alloxan-induced diabetic rats. Compared with the normal control rats, significantly higher blood glucose levels were observed in the alloxan-induced diabetic rats. The differently regioselective chitosan sulfates exhibited hypoglycemic activities at different doses and intervals, especially 3-O-sulfochitosan (3-S). The major results are as follows. First, 3,6-di-O-sulfochitosan and 3-O-sulfochitosan exhibited more significant hypoglycemic activities than 2-N-3, 6-di-O-sulfochitosan and 6-O-sulfochitosan. Moreover, 3-S-treated rats showed a more significant reduction of blood glucose levels than those treated by 3,6-di-O-sulfochitosan. These results indicated that –OSO3 at the C3-position of chitosan is a key active site. Second, 3-S significantly reduced the blood glucose levels and regulated the glucose tolerance effect in the experimental rats. Third, treatment with 3-S significantly increased the plasma insulin levels in the experimental diabetic rats. A noticeable hypoglycemic activity of 3-S in the alloxan-induced diabetic rats was shown. Clinical trials are required in the future to confirm the utility of 3-S. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>FTIR of H2,3,6-S (1) chitosan; (2) H2,3,6-S under dichloroacetic acid; (3) H2,3,6-S under formic acid.</p>
Full article ">Figure 2
<p>FTIR of 2-phthalimido-chitosan under 90 °C; 1: FTIR of 2-phthalimido-chitosan under 3.5 h; 2: FTIR of 2-phthalimido-chitosan under 3.0 h; 3: FTIR of 2-phthalimido-chitosan under 2.0 h.</p>
Full article ">Figure 3
<p>FTIR of 3,6-S; 3: Eliminating the phthalimido group under 3 h; 6: Eliminating the phthalimido group under 6 h; 10: Eliminating the phthalimido group under 10 h; 16: Eliminating the phthalimido group under 16 h.</p>
Full article ">Figure 4
<p>FTIR of 3-S.</p>
Full article ">Figure 5
<p>FTIR of 6-S; 1: Chitosan; 2: Cu-chitosan chelation; 3: Cu- sulfated chitoan chelation under formic acid; 4: Cu- sulfated chitoan chelation without formic acid.</p>
Full article ">Figure 6
<p>FTIR of L2,3,6-S; 1: Chitosan; 2: L2,3,6-S under traditional heating; 3: L2,3,6-S under microwave radiation (800W).</p>
Full article ">Figure 7
<p>Hypoglycemic effects of 3-S on the fasting blood glucose levels of normal rats during GTT, each value shown in mean ± S.E.; <span class="html-italic">n</span> = 10, number of animals in each group.</p>
Full article ">
1425 KiB  
Review
Marine Polysaccharides from Algae with Potential Biomedical Applications
by Maria Filomena De Jesus Raposo, Alcina Maria Bernardo De Morais and Rui Manuel Santos Costa De Morais
Mar. Drugs 2015, 13(5), 2967-3028; https://doi.org/10.3390/md13052967 - 15 May 2015
Cited by 493 | Viewed by 23590
Abstract
There is a current tendency towards bioactive natural products with applications in various industries, such as pharmaceutical, biomedical, cosmetics and food. This has put some emphasis in research on marine organisms, including macroalgae and microalgae, among others. Polysaccharides with marine origin constitute one [...] Read more.
There is a current tendency towards bioactive natural products with applications in various industries, such as pharmaceutical, biomedical, cosmetics and food. This has put some emphasis in research on marine organisms, including macroalgae and microalgae, among others. Polysaccharides with marine origin constitute one type of these biochemical compounds that have already proved to have several important properties, such as anticoagulant and/or antithrombotic, immunomodulatory ability, antitumor and cancer preventive, antilipidaemic and hypoglycaemic, antibiotics and anti-inflammatory and antioxidant, making them promising bioactive products and biomaterials with a wide range of applications. Their properties are mainly due to their structure and physicochemical characteristics, which depend on the organism they are produced by. In the biomedical field, the polysaccharides from algae can be used in controlled drug delivery, wound management, and regenerative medicine. This review will focus on the biomedical applications of marine polysaccharides from algae. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Examples of structures of <b>PS</b> from macro- and microalgae. (<b>A</b>) Repeating units suggested for the structure of alginates [<a href="#B3-marinedrugs-13-02967" class="html-bibr">3</a>]; (<b>B</b>) Repeating units of some carrageenans [<a href="#B3-marinedrugs-13-02967" class="html-bibr">3</a>]; (<b>C</b>) Fucoidan backbone of <span class="html-italic">A. nodosum</span> and three species of <span class="html-italic">Fucus</span>, showing the different distribution pattern of sulphate [<a href="#B75-marinedrugs-13-02967" class="html-bibr">75</a>]; (<b>D</b>) Repeating units, sulphation pattern and gycosidic bounds of the backbone structures of PS of three different brown seaweeds [<a href="#B75-marinedrugs-13-02967" class="html-bibr">75</a>]; (<b>E</b>) Alternative positions and combinations for the repeating units of ulvans. A3s and B3s are aldobiouronic repeating di-units suggested for <span class="html-italic">U. rigida</span> and <span class="html-italic">U. armoricana.</span> U3s and U2s,3s are, respectively, a xyl-(S-rham) and a (S-xyl)-(S-rham) disaccharides [<a href="#B8-marinedrugs-13-02967" class="html-bibr">8</a>]; (<b>F</b>) Galactans of <span class="html-italic">Codium</span> spp. (a) linear (1,3)-β-<span class="html-small-caps">d</span>-galactan, (b) and (c) pyruvylated branched sulphated galactans [<a href="#B8-marinedrugs-13-02967" class="html-bibr">8</a>]; (<b>G</b>) A rare mannan of the <b>PS</b> from <span class="html-italic">C. fragile</span>, with (1,3)-β-man residues and branches at C-2 [<a href="#B8-marinedrugs-13-02967" class="html-bibr">8</a>]. Tabarsa <span class="html-italic">et al.</span> [<a href="#B243-marinedrugs-13-02967" class="html-bibr">243</a>] referred that either branches or sulphates may be bound at the C-2 and/or C-4 positions along the PS backbone); (<b>H</b>) Models 1 or 2 for the possible acidic repeating unit in polysaccharide II, from <span class="html-italic">Porphyridium</span> sp. R = H, SO<sub>2</sub>O, terminal <b>gal</b> or terminal <b>xyl</b>, m = 2 or 3 [<a href="#B14-marinedrugs-13-02967" class="html-bibr">14</a>].</p>
Full article ">Figure 1 Cont.
<p>Examples of structures of <b>PS</b> from macro- and microalgae. (<b>A</b>) Repeating units suggested for the structure of alginates [<a href="#B3-marinedrugs-13-02967" class="html-bibr">3</a>]; (<b>B</b>) Repeating units of some carrageenans [<a href="#B3-marinedrugs-13-02967" class="html-bibr">3</a>]; (<b>C</b>) Fucoidan backbone of <span class="html-italic">A. nodosum</span> and three species of <span class="html-italic">Fucus</span>, showing the different distribution pattern of sulphate [<a href="#B75-marinedrugs-13-02967" class="html-bibr">75</a>]; (<b>D</b>) Repeating units, sulphation pattern and gycosidic bounds of the backbone structures of PS of three different brown seaweeds [<a href="#B75-marinedrugs-13-02967" class="html-bibr">75</a>]; (<b>E</b>) Alternative positions and combinations for the repeating units of ulvans. A3s and B3s are aldobiouronic repeating di-units suggested for <span class="html-italic">U. rigida</span> and <span class="html-italic">U. armoricana.</span> U3s and U2s,3s are, respectively, a xyl-(S-rham) and a (S-xyl)-(S-rham) disaccharides [<a href="#B8-marinedrugs-13-02967" class="html-bibr">8</a>]; (<b>F</b>) Galactans of <span class="html-italic">Codium</span> spp. (a) linear (1,3)-β-<span class="html-small-caps">d</span>-galactan, (b) and (c) pyruvylated branched sulphated galactans [<a href="#B8-marinedrugs-13-02967" class="html-bibr">8</a>]; (<b>G</b>) A rare mannan of the <b>PS</b> from <span class="html-italic">C. fragile</span>, with (1,3)-β-man residues and branches at C-2 [<a href="#B8-marinedrugs-13-02967" class="html-bibr">8</a>]. Tabarsa <span class="html-italic">et al.</span> [<a href="#B243-marinedrugs-13-02967" class="html-bibr">243</a>] referred that either branches or sulphates may be bound at the C-2 and/or C-4 positions along the PS backbone); (<b>H</b>) Models 1 or 2 for the possible acidic repeating unit in polysaccharide II, from <span class="html-italic">Porphyridium</span> sp. R = H, SO<sub>2</sub>O, terminal <b>gal</b> or terminal <b>xyl</b>, m = 2 or 3 [<a href="#B14-marinedrugs-13-02967" class="html-bibr">14</a>].</p>
Full article ">Figure 1 Cont.
<p>Examples of structures of <b>PS</b> from macro- and microalgae. (<b>A</b>) Repeating units suggested for the structure of alginates [<a href="#B3-marinedrugs-13-02967" class="html-bibr">3</a>]; (<b>B</b>) Repeating units of some carrageenans [<a href="#B3-marinedrugs-13-02967" class="html-bibr">3</a>]; (<b>C</b>) Fucoidan backbone of <span class="html-italic">A. nodosum</span> and three species of <span class="html-italic">Fucus</span>, showing the different distribution pattern of sulphate [<a href="#B75-marinedrugs-13-02967" class="html-bibr">75</a>]; (<b>D</b>) Repeating units, sulphation pattern and gycosidic bounds of the backbone structures of PS of three different brown seaweeds [<a href="#B75-marinedrugs-13-02967" class="html-bibr">75</a>]; (<b>E</b>) Alternative positions and combinations for the repeating units of ulvans. A3s and B3s are aldobiouronic repeating di-units suggested for <span class="html-italic">U. rigida</span> and <span class="html-italic">U. armoricana.</span> U3s and U2s,3s are, respectively, a xyl-(S-rham) and a (S-xyl)-(S-rham) disaccharides [<a href="#B8-marinedrugs-13-02967" class="html-bibr">8</a>]; (<b>F</b>) Galactans of <span class="html-italic">Codium</span> spp. (a) linear (1,3)-β-<span class="html-small-caps">d</span>-galactan, (b) and (c) pyruvylated branched sulphated galactans [<a href="#B8-marinedrugs-13-02967" class="html-bibr">8</a>]; (<b>G</b>) A rare mannan of the <b>PS</b> from <span class="html-italic">C. fragile</span>, with (1,3)-β-man residues and branches at C-2 [<a href="#B8-marinedrugs-13-02967" class="html-bibr">8</a>]. Tabarsa <span class="html-italic">et al.</span> [<a href="#B243-marinedrugs-13-02967" class="html-bibr">243</a>] referred that either branches or sulphates may be bound at the C-2 and/or C-4 positions along the PS backbone); (<b>H</b>) Models 1 or 2 for the possible acidic repeating unit in polysaccharide II, from <span class="html-italic">Porphyridium</span> sp. R = H, SO<sub>2</sub>O, terminal <b>gal</b> or terminal <b>xyl</b>, m = 2 or 3 [<a href="#B14-marinedrugs-13-02967" class="html-bibr">14</a>].</p>
Full article ">Figure 1 Cont.
<p>Examples of structures of <b>PS</b> from macro- and microalgae. (<b>A</b>) Repeating units suggested for the structure of alginates [<a href="#B3-marinedrugs-13-02967" class="html-bibr">3</a>]; (<b>B</b>) Repeating units of some carrageenans [<a href="#B3-marinedrugs-13-02967" class="html-bibr">3</a>]; (<b>C</b>) Fucoidan backbone of <span class="html-italic">A. nodosum</span> and three species of <span class="html-italic">Fucus</span>, showing the different distribution pattern of sulphate [<a href="#B75-marinedrugs-13-02967" class="html-bibr">75</a>]; (<b>D</b>) Repeating units, sulphation pattern and gycosidic bounds of the backbone structures of PS of three different brown seaweeds [<a href="#B75-marinedrugs-13-02967" class="html-bibr">75</a>]; (<b>E</b>) Alternative positions and combinations for the repeating units of ulvans. A3s and B3s are aldobiouronic repeating di-units suggested for <span class="html-italic">U. rigida</span> and <span class="html-italic">U. armoricana.</span> U3s and U2s,3s are, respectively, a xyl-(S-rham) and a (S-xyl)-(S-rham) disaccharides [<a href="#B8-marinedrugs-13-02967" class="html-bibr">8</a>]; (<b>F</b>) Galactans of <span class="html-italic">Codium</span> spp. (a) linear (1,3)-β-<span class="html-small-caps">d</span>-galactan, (b) and (c) pyruvylated branched sulphated galactans [<a href="#B8-marinedrugs-13-02967" class="html-bibr">8</a>]; (<b>G</b>) A rare mannan of the <b>PS</b> from <span class="html-italic">C. fragile</span>, with (1,3)-β-man residues and branches at C-2 [<a href="#B8-marinedrugs-13-02967" class="html-bibr">8</a>]. Tabarsa <span class="html-italic">et al.</span> [<a href="#B243-marinedrugs-13-02967" class="html-bibr">243</a>] referred that either branches or sulphates may be bound at the C-2 and/or C-4 positions along the PS backbone); (<b>H</b>) Models 1 or 2 for the possible acidic repeating unit in polysaccharide II, from <span class="html-italic">Porphyridium</span> sp. R = H, SO<sub>2</sub>O, terminal <b>gal</b> or terminal <b>xyl</b>, m = 2 or 3 [<a href="#B14-marinedrugs-13-02967" class="html-bibr">14</a>].</p>
Full article ">
2607 KiB  
Article
Alginate Hydrogels Coated with Chitosan for Wound Dressing
by Maria Cristina Straccia, Giovanna Gomez D'Ayala, Ida Romano, Adriana Oliva and Paola Laurienzo
Mar. Drugs 2015, 13(5), 2890-2908; https://doi.org/10.3390/md13052890 - 11 May 2015
Cited by 128 | Viewed by 13713
Abstract
In this work, a coating of chitosan onto alginate hydrogels was realized using the water-soluble hydrochloride form of chitosan (CH-Cl), with the dual purpose of imparting antibacterial activity and delaying the release of hydrophilic molecules from the alginate matrix. Alginate hydrogels with different [...] Read more.
In this work, a coating of chitosan onto alginate hydrogels was realized using the water-soluble hydrochloride form of chitosan (CH-Cl), with the dual purpose of imparting antibacterial activity and delaying the release of hydrophilic molecules from the alginate matrix. Alginate hydrogels with different calcium contents were prepared by the internal setting method and coated by immersion in a CH-Cl solution. Structural analysis by cryo-scanning electron microscopy was carried out to highlight morphological alterations due to the coating layer. Tests in vitro with human mesenchymal stromal cells (MSC) were assessed to check the absence of toxicity of CH-Cl. Swelling, stability in physiological solution and release characteristics using rhodamine B as the hydrophilic model drug were compared to those of relative uncoated hydrogels. Finally, antibacterial activity against Escherichia coli was tested. Results show that alginate hydrogels coated with chitosan hydrochloride described here can be proposed as a novel medicated dressing by associating intrinsic antimicrobial activity with improved sustained release characteristics. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) FTIR spectrum of chitosan; (<b>b</b>) FTIR spectrum of chitosan hydrochloride (CH-Cl).</p>
Full article ">Figure 2
<p>(<b>a</b>) Homogeneity profiles of Alg hydrogels; (<b>b</b>) homogeneity profiles of c-Alg hydrogels.</p>
Full article ">Figure 3
<p>(<b>a</b>) Swelling curves of Alg hydrogels; (<b>b</b>) swelling curves of c-Alg hydrogels.</p>
Full article ">Figure 4
<p>(<b>a</b>) Stability test in normal saline solution of Alg hydrogels; (<b>b</b>) stability test in normal saline solution of c-Alg hydrogels.</p>
Full article ">Figure 5
<p>(<b>a</b>–<b>c</b>) Cryo-SEM images of Alg1; (<b>d</b>–<b>f</b>) cryo-SEM images of c-Alg1.</p>
Full article ">Figure 5 Cont.
<p>(<b>a</b>–<b>c</b>) Cryo-SEM images of Alg1; (<b>d</b>–<b>f</b>) cryo-SEM images of c-Alg1.</p>
Full article ">Figure 6
<p>Equipment employed to study the release of rhodamine B in physiological conditions. The photo refers to different release times.</p>
Full article ">Figure 7
<p>Release profiles of RhB from: (<b>a</b>) Alg1 and c-Alg1; (<b>b</b>) Alg2 and c-Alg2; (<b>c</b>) Alg3 and c-Alg3; (<b>d</b>) Alg4 and c-Alg4.</p>
Full article ">Figure 8
<p>(<b>a</b>,<b>b</b>) Microscopic images of MSC cultured for seven days in control culture medium after crystal violet staining; (<b>c</b>,<b>d</b>) microscopic images of MSC cultured for seven days in 0.1% CH-Cl culture medium after crystal violet staining.</p>
Full article ">Figure 9
<p>(<b>a</b>) Optical photo of <span class="html-italic">E. coli</span> growing on agar in contact with Alg1; (<b>b</b>) optical photo of <span class="html-italic">E. coli</span> growing on agar in contact with c-Alg1.</p>
Full article ">Figure 10
<p>Antimicrobial activity kinetics of c-Alg hydrogels against <span class="html-italic">E. coli</span>.</p>
Full article ">
1278 KiB  
Article
λ-Carrageenan Suppresses Tomato Chlorotic Dwarf Viroid (TCDVd) Replication and Symptom Expression in Tomatoes
by Jatinder S. Sangha, Saveetha Kandasamy, Wajahatullah Khan, Navratan Singh Bahia, Rudra P. Singh, Alan T. Critchley and Balakrishnan Prithiviraj
Mar. Drugs 2015, 13(5), 2875-2889; https://doi.org/10.3390/md13052875 - 8 May 2015
Cited by 36 | Viewed by 9108
Abstract
The effect of carrageenans on tomato chlorotic dwarf viroid (TCDVd) replication and symptom expression was studied. Three-week-old tomato plants were spray-treated with iota(ɩ)-, lambda(λ)-, and kappa(κ)-carrageenan at 1 g·L−1 and inoculated with TCDVd after 48 h. The λ-carrageenan significantly suppressed viroid symptom [...] Read more.
The effect of carrageenans on tomato chlorotic dwarf viroid (TCDVd) replication and symptom expression was studied. Three-week-old tomato plants were spray-treated with iota(ɩ)-, lambda(λ)-, and kappa(κ)-carrageenan at 1 g·L−1 and inoculated with TCDVd after 48 h. The λ-carrageenan significantly suppressed viroid symptom expression after eight weeks of inoculation, only 28% plants showed distinctive bunchy-top symptoms as compared to the 82% in the control group. Viroid concentration was reduced in the infected shoot cuttings incubated in λ-carrageenan amended growth medium. Proteome analysis revealed that 16 tomato proteins were differentially expressed in the λ-carrageenan treated plants. Jasmonic acid related genes, allene oxide synthase (AOS) and lipoxygenase (LOX), were up-regulated in λ-carrageenan treatment during viroid infection. Taken together, our results suggest that λ-carrageenan induced tomato defense against TCDVd, which was partly jasmonic acid (JA) dependent, and that it could be explored in plant protection against viroid infection. Full article
Show Figures

Figure 1

Figure 1
<p>Symptoms of TCDVd infection in tomatoes. Three-week-old tomato plants were inoculated with 10 μL of TCDVd sap and the symptoms were observed at 28 dpi. Infected plants (RT-PCR confirmed) were stunted with bunchy top symptoms, smaller fruits and fewer roots.</p>
Full article ">Figure 2
<p>Suppressive effect of λ-carrageenan in tomatoes against TCDVd infection. (<b>a</b>) Percentage of plants showing typical TCDVd symptoms in control and λ-carrageenan (λ-Carr) treated plants. Data represent the mean of percent infected plants from three independent trials (Mean ± SEM, <span class="html-italic">n</span> = 36); (<b>b</b>) Relative intensity of TCDVd bands visualized on agarose gel and quantified with ImageJ software in controls and λ-carrageenan (λ-Carr) treated plants at 35 dpi (Mean ± SEM, <span class="html-italic">n</span> = 10).</p>
Full article ">Figure 3
<p>Plant height and last internode length of infected and healthy tomato plants one month after TCDVd inoculation. Bars represent control (black) and λ-carrageenan (λ-Carr, white). Data represent Mean ± SEM (<span class="html-italic">n</span> = 36 for healthy plants and <span class="html-italic">n</span> = 12 plants for infected plants), student’s <span class="html-italic">t</span>-test at <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 4
<p>Effect of carrageenans on TCDVd multiplication in tomato determined by RT-PCR. TCDVd concentration in the leaf samples from infected shoots was determined at 0, 7, 14, and 21 days post carrageenan treatment (λ-carrageenan (λ-Carr), ɩ-carrageenan (ɩ-Carr), κ-carrageenan (κ-Carr)) with RT-PCR using TCDVd specific primers. The bands were visualized on agarose gel.</p>
Full article ">Figure 5
<p>A general proteome map of tomato leaves with TCDVd infection in λ-carrageenan treatment. Proteins were separated first on an IPG strip (pH 4.0–7.0) and then based on molecular weight (kDA) on a 12% SDS-PAGE. Differentially expressed proteins are marked, red circled were increased whereas green circled were decreased.</p>
Full article ">Figure 6
<p>Expression of defense response genes encoding allene oxide synthase (AOS), lipoxygenase (LOX) and pathogenesis related protein 1 (PR1), in tomatoes during TCDVd infection. The expression of genes was analyzed at 0 (AOS-0, LOX-0 and PR1-0) and 7 (AOS-7, LOX-7 and PR1-7) dpi in control and λ-carrageenan (λ-Carr) treated plants. Values with “*” are significantly different (<span class="html-italic">p</span> &lt; 0.05). (Mean ± S.E., <span class="html-italic">n</span> = 3).</p>
Full article ">
1546 KiB  
Review
Perspective on the Use of Sulfated Polysaccharides from Marine Organisms as a Source of New Antithrombotic Drugs
by Paulo A. S. Mourão
Mar. Drugs 2015, 13(5), 2770-2784; https://doi.org/10.3390/md13052770 - 6 May 2015
Cited by 93 | Viewed by 8048
Abstract
Thromboembolic diseases are increasing worldwide and always require anticoagulant therapy. We still need safer and more secure antithrombotic drugs than those presently available. Sulfated polysaccharides from marine organisms may constitute a new source for the development of such drugs. Investigation of these compounds [...] Read more.
Thromboembolic diseases are increasing worldwide and always require anticoagulant therapy. We still need safer and more secure antithrombotic drugs than those presently available. Sulfated polysaccharides from marine organisms may constitute a new source for the development of such drugs. Investigation of these compounds usually attempts to reproduce the therapeutic effects of heparin. However, we may need to follow different routes, focusing particularly in the following aspects: (1) defining precisely the specific structures required for interaction of these sulfated polysaccharides with proteins of the coagulation system; (2) looking for alternative mechanisms of action, distinct from those of heparin; (3) identifying side effects (mostly pro-coagulant action and hypotension rather than bleeding) and preparing derivatives that retain the desired antithrombotic action but are devoid of side effects; (4) considering that sulfated polysaccharides with low anticoagulant action on in vitro assays may display potent effects on animal models of experimental thrombosis; and finally (5) investigating the antithrombotic effect of these sulfated polysaccharides after oral administration or preparing derivatives that may achieve this effect. If these aspects are successfully addressed, sulfated polysaccharides from marine organisms may conquer the frontier of antithrombotic therapy and open new avenues for treatment or prevention of thromboembolic diseases. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Examples of the structures of sulfated galactans from (<b>A</b>) red algae [<a href="#B15-marinedrugs-13-02770" class="html-bibr">15</a>,<a href="#B16-marinedrugs-13-02770" class="html-bibr">16</a>,<a href="#B17-marinedrugs-13-02770" class="html-bibr">17</a>,<a href="#B18-marinedrugs-13-02770" class="html-bibr">18</a>,<a href="#B19-marinedrugs-13-02770" class="html-bibr">19</a>]; (<b>B</b>) sea urchins [<a href="#B20-marinedrugs-13-02770" class="html-bibr">20</a>,<a href="#B21-marinedrugs-13-02770" class="html-bibr">21</a>]; (<b>C</b>) sulfated fucan from brown algae [<a href="#B22-marinedrugs-13-02770" class="html-bibr">22</a>,<a href="#B23-marinedrugs-13-02770" class="html-bibr">23</a>,<a href="#B24-marinedrugs-13-02770" class="html-bibr">24</a>]; and (<b>D</b>) from sea urchins [<a href="#B25-marinedrugs-13-02770" class="html-bibr">25</a>,<a href="#B26-marinedrugs-13-02770" class="html-bibr">26</a>]. Structures of the sulfated polysaccharides from marine algae highlighted in blue and those from sea urchins in red.</p>
Full article ">Figure 2
<p>Structures (<b>A</b>,<b>B</b>) and antithrombotic effect (<b>C</b>) of sulfated polysaccharides rich in 2,4-disulfated α-<span class="html-small-caps">l</span>-Fuc<span class="html-italic">p</span> units. (<b>A</b>) The sulfated fucan is composed of [3-α-<span class="html-small-caps">l</span>-Fuc<span class="html-italic">p</span>-2,4(OSO<sub>3</sub><sup>−</sup>)-1→3-α-<span class="html-small-caps">l</span>-Fuc<span class="html-italic">p</span>-1→3-α-<span class="html-small-caps">l</span>-Fuc<span class="html-italic">p</span>-2(OSO<sub>3</sub><sup>−</sup>)-1→3-α-<span class="html-small-caps">l</span>-Fuc<span class="html-italic">p</span>-2(OSO<sub>3</sub><sup>−</sup>)]<span class="html-italic"><sub>n</sub></span> repeating units; (<b>B</b>) Fucosylated chondroitin sulfate has a chondroitin sulfate-like backbone, but contains branches of 2,4-disulfated α-<span class="html-small-caps">l</span>-Fuc<span class="html-italic">p</span> units linked to the β-<span class="html-small-caps">d</span>-GlcA residues on the polysaccharide core. The 2,4-disulfated α-<span class="html-small-caps">l</span>-Fuc<span class="html-italic">p</span> units in these two polysaccharides are highlighted in green or blue; (<b>C</b>) Antithrombotic effect of sulfated fucan (<math display="inline"> <semantics> <mrow> <mstyle mathcolor="#4F6228"> <mo>●</mo> </mstyle> </mrow> </semantics> </math>), fucosylated chondroitin sulfate (<math display="inline"> <semantics> <mrow> <mstyle mathcolor="#1F497D"> <mo>●</mo> </mstyle> </mrow> </semantics> </math>), unfractionated heparin (■), low-molecular-weight heparin (<math display="inline"> <semantics> <mrow> <mstyle mathcolor="red"> <mo>■</mo> </mstyle> </mrow> </semantics> </math>) and vertebrate chondroitin 6-sulfate (▲) on an arterial thrombosis model induced in carotid artery of rats by laser irradiation. See Ref. [<a href="#B35-marinedrugs-13-02770" class="html-bibr">35</a>] for further details.</p>
Full article ">Figure 3
<p>Structures and anticoagulant effect of a sulfated galactan and a sulfated fucan. (<b>A</b>,<b>B</b>) Structures of the sulfated α-<span class="html-small-caps">l</span>-galactan from the sea urchin <span class="html-italic">E. lucunter</span> (in blue, Panel A) and sulfated α-<span class="html-small-caps">l</span>-fucan from <span class="html-italic">S.</span> <span class="html-italic">franciscanus</span> (in red, Panel B). Both polysaccharides are 3-linked with a regular sulfation at the 2-position, but differ in their constituent monosaccharides; (<b>C</b>) Thrombin inhibition mediated by antithrombin <span class="html-italic">vs.</span> concentrations of the sulfated galactan (<math display="inline"> <semantics> <mrow> <mstyle mathcolor="#1F497D"> <mo>●</mo> </mstyle> </mrow> </semantics> </math>), sulfated fucan (<math display="inline"> <semantics> <mrow> <mstyle mathcolor="red"> <mo>▲</mo> </mstyle> </mrow> </semantics> </math>) and heparin (□). See Ref. [<a href="#B36-marinedrugs-13-02770" class="html-bibr">36</a>] for further details.</p>
Full article ">Figure 4
<p>Major target sites for the sulfated polysaccharides from marine organisms on the coagulation system. Blue and red arrows indicate anticoagulant and pro-coagulant effects, respectively. + indicates activation and − indicates inhibitory effects. Anticoagulant effect: sulfated galactans (SG) from marine red algae and fucosylated chondroitin sulfates (FCS) from sea cucumbers inhibit the intrinsic tenase and prothrombinase complexes [<a href="#B37-marinedrugs-13-02770" class="html-bibr">37</a>,<a href="#B38-marinedrugs-13-02770" class="html-bibr">38</a>]. It is still unclear if sulfated fucans (SF) have similar effects. These polysaccharides also potentiate the inhibitory effect of antithrombin (AT) and/or heparin cofactor II (HCII) on thrombin [<a href="#B16-marinedrugs-13-02770" class="html-bibr">16</a>,<a href="#B32-marinedrugs-13-02770" class="html-bibr">32</a>,<a href="#B33-marinedrugs-13-02770" class="html-bibr">33</a>,<a href="#B36-marinedrugs-13-02770" class="html-bibr">36</a>]. Their effects on factor Xa are very modest. The serpin-independent action preponderates on the plasma system. Pro-coagulant effect: SG and FCS activate factor XII [<a href="#B31-marinedrugs-13-02770" class="html-bibr">31</a>,<a href="#B39-marinedrugs-13-02770" class="html-bibr">39</a>]. This effect may result in severe hypotension (due to bradykinin release) and pro-coagulant (and pro-thrombotic) action. It is unclear if SF activates factor XII. SF inhibits Tissue Factor Protease Inhibitor (TFPI), a specific inhibitor of the extrinsic tenase complex. Consequently, SF has a pro-coagulant effect [<a href="#B40-marinedrugs-13-02770" class="html-bibr">40</a>,<a href="#B41-marinedrugs-13-02770" class="html-bibr">41</a>].</p>
Full article ">Figure 5
<p>Effect of sulfated galactans on anticoagulant assays (<b>A</b>,<b>C</b>,<b>D</b>) and on arterial experimental thrombosis in rats (<b>B</b>). Different concentrations or doses of heparin (■,□), native sulfated galactan from <span class="html-italic">B. occidentalis</span> (<math display="inline"> <semantics> <mrow> <mstyle mathcolor="#4F6228"> <mo>▲</mo> </mstyle> </mrow> </semantics> </math>,<math display="inline"> <semantics> <mrow> <mstyle mathcolor="#4F6228"> <mo>∆</mo> </mstyle> </mrow> </semantics> </math>) or from <span class="html-italic">A. muscoides</span> (<math display="inline"> <semantics> <mrow> <mstyle mathcolor="red"> <mo>♦</mo> </mstyle> </mrow> </semantics> </math>,<math display="inline"> <semantics> <mrow> <mstyle mathcolor="red"> <mo>◊</mo> </mstyle> </mrow> </semantics> </math>) or the sulfated galactan from <span class="html-italic">B. occidentalis</span> with reduced molecular weight (<math display="inline"> <semantics> <mrow> <mstyle mathcolor="#1F497D"> <mo>●</mo> </mstyle> </mrow> </semantics> </math>,<math display="inline"> <semantics> <mrow> <mstyle mathcolor="#1F497D"> <mo>○</mo> </mstyle> </mrow> </semantics> </math>) were tested. In Panels C and D closed and open symbols refer to assays performed using normal or serpin-depleted plasma, respectively. Data from Ref. [<a href="#B19-marinedrugs-13-02770" class="html-bibr">19</a>].</p>
Full article ">Figure 6
<p>Anticoagulant activity based on recalcification time (<b>A</b>) and antithrombotic effect on a stasis thrombosis model in vena cava of rats (<b>B</b>,<b>C</b>). Different concentrations (<b>A</b>) or doses (<b>B</b>,<b>C</b>) of unfractionated heparin (□), low-molecular-weight heparin (■), native sulfated galactan (▲), ~14 kDa (O) and ~5 kDa (●) fragments were tested. In Panel A the anticoagulant activity was expressed as T<sub>1</sub>/T<sub>0</sub>, which is the ratio between the clotting time in the presence or absence of sulfated polysaccharide. The broken line (T<sub>1</sub>/T<sub>0</sub> = 1) indicates no effect from the polysaccharide on coagulation. Data from Ref. [<a href="#B39-marinedrugs-13-02770" class="html-bibr">39</a>].</p>
Full article ">
539 KiB  
Article
Employment of Marine Polysaccharides to Manufacture Functional Biocomposites for Aquaculture Feeding Applications
by Marina Paolucci, Gabriella Fasulo and Maria Grazia Volpe
Mar. Drugs 2015, 13(5), 2680-2693; https://doi.org/10.3390/md13052680 - 29 Apr 2015
Cited by 12 | Viewed by 5289
Abstract
In this study, polysaccharides of marine origin (agar, alginate and κ-carrageenan) were used to embed nutrients to fabricate biocomposites to be employed in animal feeding. The consistency of biocomposites in water has been evaluated up to 14 days, by several methods: swelling, nutrient [...] Read more.
In this study, polysaccharides of marine origin (agar, alginate and κ-carrageenan) were used to embed nutrients to fabricate biocomposites to be employed in animal feeding. The consistency of biocomposites in water has been evaluated up to 14 days, by several methods: swelling, nutrient release and granulometric analysis. Biocomposites were produced with varying percentages of nutrients (5%–25%) and polysaccharides (1%–2%–3%). All possible biopolymer combinations were tested in order to select those with the best network strength. The best performing biocomposites were those manufactured with agar 2% and nutrients 10%, showing the lowest percentage of water absorption and nutrient release. Biocomposites made of agar 2% and nutrients 10% were the most stable in water and were therefore used to analyze their behavior in water with respect to the release of quercetin, a phenolic compound with demonstrated high antibacterial and antioxidant activities. The leaching of such molecules in water was therefore employed as a further indicator of biocomposite water stability. Altogether, our results confirm the suitability of agar as a binder for biocomposites and provide a positive contribution to aquaculture. Full article
Show Figures

Figure 1

Figure 1
<p>Pictures of biocomposites made with 2% polymer and 10% nutrients. (Right) agar; (center) alginate; (left) κ-carrageenan.</p>
Full article ">Figure 2
<p>(<b>a</b>,<b>c</b>,<b>e</b>,<b>g</b>,<b>i</b>) percentage of water adsorbed by biocomposites made with agar 1%–2%–3% and containing nutrients 5%–10%–15%–20%–25%; (<b>b</b>,<b>d</b>,<b>f</b>,<b>h</b>,<b>j</b>) percentage of nutrients released by biocomposites with agar 1%–2%–3% and containing nutrients 5%–10%–15%–20%–25%.</p>
Full article ">Figure 2 Cont.
<p>(<b>a</b>,<b>c</b>,<b>e</b>,<b>g</b>,<b>i</b>) percentage of water adsorbed by biocomposites made with agar 1%–2%–3% and containing nutrients 5%–10%–15%–20%–25%; (<b>b</b>,<b>d</b>,<b>f</b>,<b>h</b>,<b>j</b>) percentage of nutrients released by biocomposites with agar 1%–2%–3% and containing nutrients 5%–10%–15%–20%–25%.</p>
Full article ">Figure 3
<p>Quantity of quercetin (mg QE eq/100 g fresh weight) released by biocomposites made with agar 1%–2%–3% and containing nutrients 5%. Fresh weight refers to biocomposites formulated without any treatment before analysis.</p>
Full article ">
440 KiB  
Article
Seaweed Polysaccharides (Laminarin and Fucoidan) as Functional Ingredients in Pork Meat: An Evaluation of Anti-Oxidative Potential, Thermal Stability and Bioaccessibility
by Natasha C. Moroney, Michael N. O'Grady, Sinéad Lordan, Catherine Stanton and Joseph P. Kerry
Mar. Drugs 2015, 13(4), 2447-2464; https://doi.org/10.3390/md13042447 - 20 Apr 2015
Cited by 66 | Viewed by 9574
Abstract
The anti-oxidative potential of laminarin (L), fucoidan (F) and an L/F seaweed extract was measured using the DPPH free radical scavenging assay, in 25% pork (longissimus thoracis et lumborum (LTL)) homogenates (TBARS) (3 and 6 mg/mL) and in horse heart oxymyoglobin (OxyMb) [...] Read more.
The anti-oxidative potential of laminarin (L), fucoidan (F) and an L/F seaweed extract was measured using the DPPH free radical scavenging assay, in 25% pork (longissimus thoracis et lumborum (LTL)) homogenates (TBARS) (3 and 6 mg/mL) and in horse heart oxymyoglobin (OxyMb) (0.1 and 1 mg/mL). The DPPH activity of fresh and cooked minced LTL containing L (100 mg/g; L100), F100 and L/F100,300, and bioaccessibility post in vitro digestion (L/F300), was assessed. Theoretical cellular uptake of antioxidant compounds was measured in a transwell Caco-2 cell model. Laminarin displayed no activity and fucoidan reduced lipid oxidation but catalysed OxyMb oxidation. Fucoidan activity was lowered by cooking while the L/F extract displayed moderate thermal stability. A decrease in DPPH antioxidant activity of 44.15% and 36.63%, after 4 and 20 h respectively, indicated theoretical uptake of L/F antioxidant compounds. Results highlight the potential use of seaweed extracts as functional ingredients in pork. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Lipid oxidation in 25% <span class="html-italic">longissimus thoracis et lumborum</span> (LTL) pork muscle homogenates following the addition of L, F or L/F and storage for up to 4 h at 4 °C. <b>*</b> Subscripts 3 and 6 denote concentrations in mg/mL. <sup>abcd</sup> Mean values (± standard deviation error bars) bearing different superscripts are significantly different, <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 2
<p>Absorbance spectra of oxymyoglobin (OxyMb) alone and following the addition of F<sub>1</sub> (* Subscript 1 denotes concentration in mg/mL) and storage for up to 8 days at 4 °C.</p>
Full article ">Figure 3
<p>Free radical scavenging activity (DPPH) of L, F or L/F in fresh and cooked minced <span class="html-italic">longissimus thoracis et lumborum</span> (LTL) pork muscle stored for 20 h at ~20 °C. <b>*</b> Subscripts 100 and 300 denote concentrations in mg/g. <sup>ab</sup> Within each treatment, mean values (± standard deviation error bars) bearing different superscripts are significantly different, <span class="html-italic">p</span> &lt; 0.05. Comparing <sup>wx</sup> fresh and <sup>yz</sup> cooked LTL pork muscle treatments to their respective controls, mean values bearing different superscripts are significantly different, <span class="html-italic">p</span> &lt; 0.05. ( <span class="html-fig-inline" id="marinedrugs-13-02447-i001"> <img alt="Marinedrugs 13 02447 i001" src="/marinedrugs/marinedrugs-13-02447/article_deploy/html/images/marinedrugs-13-02447-i001.png"/></span>), fresh; ( <span class="html-fig-inline" id="marinedrugs-13-02447-i002"> <img alt="Marinedrugs 13 02447 i002" src="/marinedrugs/marinedrugs-13-02447/article_deploy/html/images/marinedrugs-13-02447-i002.png"/></span>), cooked.</p>
Full article ">Figure 4
<p>Free radical scavenging activity (DPPH) of L, F or L/F in digested cooked minced <span class="html-italic">longissimus thoracis et lumborum</span> (LTL) pork muscle stored for 20 h at ~20 °C. <b>*</b> Subscripts 100 and 300 denote concentrations in mg/g. <sup>abc</sup> Mean values (± standard deviation error bars) bearing different superscripts are significantly different, <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">
894 KiB  
Article
Chitin-Lignin Material as a Novel Matrix for Enzyme Immobilization
by Jakub Zdarta, Łukasz Klapiszewski, Marcin Wysokowski, Małgorzata Norman, Agnieszka Kołodziejczak-Radzimska, Dariusz Moszyński, Hermann Ehrlich, Hieronim Maciejewski, Allison L. Stelling and Teofil Jesionowski
Mar. Drugs 2015, 13(4), 2424-2446; https://doi.org/10.3390/md13042424 - 20 Apr 2015
Cited by 68 | Viewed by 10190
Abstract
Innovative materials were made via the combination of chitin and lignin, and the immobilization of lipase from Aspergillus niger. Analysis by techniques including FTIR, XPS and 13C CP MAS NMR confirmed the effective immobilization of the enzyme on the surface of [...] Read more.
Innovative materials were made via the combination of chitin and lignin, and the immobilization of lipase from Aspergillus niger. Analysis by techniques including FTIR, XPS and 13C CP MAS NMR confirmed the effective immobilization of the enzyme on the surface of the composite support. The electrokinetic properties of the resulting systems were also determined. Results obtained from elemental analysis and by the Bradford method enabled the determination of optimum parameters for the immobilization process. Based on the hydrolysis reaction of para-nitrophenyl palmitate, a determination was made of the catalytic activity, thermal and pH stability, and reusability. The systems with immobilized enzymes were found to have a hydrolytic activity of 5.72 mU, and increased thermal and pH stability compared with the native lipase. The products were also shown to retain approximately 80% of their initial catalytic activity, even after 20 reaction cycles. The immobilization process, using a cheap, non-toxic matrix of natural origin, leads to systems with potential applications in wastewater remediation processes and in biosensors. Full article
Show Figures

Figure 1

Figure 1
<p>FTIR spectra of chitin-lignin composite and lipase (<b>a</b>) and selected products following 24 h of enzyme immobilization (<b>b</b>), in two different spectral range.</p>
Full article ">Figure 1 Cont.
<p>FTIR spectra of chitin-lignin composite and lipase (<b>a</b>) and selected products following 24 h of enzyme immobilization (<b>b</b>), in two different spectral range.</p>
Full article ">Figure 2
<p><sup>13</sup>C CP MAS NMR spectra of chitin-lignin (<b>a</b>); lipase (<b>b</b>) and chitin-lignin matrix with immobilized enzyme (<b>c</b>).</p>
Full article ">Figure 3
<p>The XPS C 1s spectra for chitin-lignin (<b>a</b>); lipase (<b>b</b>); and the chitin-lignin + lipase product (<b>c</b>).</p>
Full article ">Figure 4
<p>XPS O 1s spectra for lipase, chitin-lignin matrix and the product following enzyme immobilization.</p>
Full article ">Figure 5
<p>The zeta potential, as a function of pH, of the chitin-lignin material and selected products following immobilization.</p>
Full article ">Figure 6
<p>Graph showing changes in the catalytic activity of products depending on the time of immobilization and the concentration of the enzyme solution.</p>
Full article ">Figure 7
<p>Graph of thermal stability of immobilized and native lipase in the temperature range 10–80 °C.</p>
Full article ">Figure 8
<p>Graph showing changes in the catalytic active of immobilized and native lipase over the pH range 3–11.</p>
Full article ">Figure 9
<p>Changes in catalytic activity of immobilized lipase over 20 catalytic cycles.</p>
Full article ">
579 KiB  
Review
The Potential of Chitosan and Its Derivatives in Prevention and Treatment of Age-Related Diseases
by Garry Kerch
Mar. Drugs 2015, 13(4), 2158-2182; https://doi.org/10.3390/md13042158 - 13 Apr 2015
Cited by 98 | Viewed by 14043
Abstract
Age-related, diet-related and protein conformational diseases, such as atherosclerosis, diabetes mellitus, cancer, hypercholesterolemia, cardiovascular and neurodegenerative diseases are common in the elderly population. The potential of chitosan, chitooligosaccharides and their derivatives in prevention and treatment of age-related dysfunctions is reviewed and discussed in [...] Read more.
Age-related, diet-related and protein conformational diseases, such as atherosclerosis, diabetes mellitus, cancer, hypercholesterolemia, cardiovascular and neurodegenerative diseases are common in the elderly population. The potential of chitosan, chitooligosaccharides and their derivatives in prevention and treatment of age-related dysfunctions is reviewed and discussed in this paper. The influence of oxidative stress, low density lipoprotein oxidation, increase of tissue stiffness, protein conformational changes, aging-associated chronic inflammation and their pathobiological significance have been considered. The chitosan-based functional food also has been reviewed. Full article
Show Figures

Figure 1

Figure 1
<p>The potential effect of chitosan on age-related dysfunctions.</p>
Full article ">Figure 2
<p>The possible anticancer effect of chitosan and polyphenols encapsulated in chitosan nanoparticles.</p>
Full article ">Figure 3
<p>The possible effect of chitosan oligosaccharide on Alzheimer disease pathology.</p>
Full article ">Figure 4
<p>The possible influence of chitosan on low density lipoprotein cholesterol content, hydration, protein conformation, and protein conformational diseases.</p>
Full article ">
2239 KiB  
Article
Structural Analysis and Anticoagulant Activities of the Novel Sulfated Fucan Possessing a Regular Well-Defined Repeating Unit from Sea Cucumber
by Mingyi Wu, Li Xu, Longyan Zhao, Chuang Xiao, Na Gao, Lan Luo, Lian Yang, Zi Li, Lingyun Chen and Jinhua Zhao
Mar. Drugs 2015, 13(4), 2063-2084; https://doi.org/10.3390/md13042063 - 13 Apr 2015
Cited by 59 | Viewed by 11964
Abstract
Sulfated fucans, the complex polysaccharides, exhibit various biological activities. Herein, we purified two fucans from the sea cucumbers Holothuria edulis and Ludwigothurea grisea. Their structures were verified by means of HPGPC, FT-IR, GC–MS and NMR. As a result, a novel structural motif [...] Read more.
Sulfated fucans, the complex polysaccharides, exhibit various biological activities. Herein, we purified two fucans from the sea cucumbers Holothuria edulis and Ludwigothurea grisea. Their structures were verified by means of HPGPC, FT-IR, GC–MS and NMR. As a result, a novel structural motif for this type of polymers is reported. The fucans have a unique structure composed of a central core of regular (1→2) and (1→3)-linked tetrasaccharide repeating units. Approximately 50% of the units from L. grisea (100% for H. edulis fucan) contain sides of oligosaccharides formed by nonsulfated fucose units linked to the O-4 position of the central core. Anticoagulant activity assays indicate that the sea cucumber fucans strongly inhibit human blood clotting through the intrinsic pathways of the coagulation cascade. Moreover, the mechanism of anticoagulant action of the fucans is selective inhibition of thrombin activity by heparin cofactor II. The distinctive tetrasaccharide repeating units contribute to the anticoagulant action. Additionally, unlike the fucans from marine alga, although the sea cucumber fucans have great molecular weights and affluent sulfates, they do not induce platelet aggregation. Overall, our results may be helpful in understanding the structure-function relationships of the well-defined polysaccharides from invertebrate as new types of safer anticoagulants. Full article
Show Figures

Figure 1

Figure 1
<p>FT-IR spectrum of the sulfated fucan from sea cucumber.</p>
Full article ">Figure 2
<p><sup>1</sup>H (<b>A</b>,<b>B</b>) and <sup>13</sup>C (<b>C</b>) one-dimensional NMR spectra at 500 MHz of the sulfated fucan from <span class="html-italic">H. edulis</span>. The spectra were recorded at 300 K for samples in D<sub>2</sub>O solution. Chemical shifts are relative to external trimethylsilylpropionic acid at 0 ppm. The residual water has been suppressed by pre-saturation. The anomeric signals assigned by <sup>1</sup>H/<sup>13</sup>C HSQC (see <a href="#marinedrugs-13-02063-f005" class="html-fig">Figure 5</a>) are labeled A–E in the sulfated fucan. Expansion of the 4.9–5.6 ppm region of the <sup>1</sup>H spectrum is shown in the <span class="html-italic">inset</span> in (<b>A</b>). The integrals were listed under the anomeric signals (<b>B</b>).</p>
Full article ">Figure 3
<p><sup>1</sup>H/<sup>1</sup>H COSY spectra of the sulfated fucans from <span class="html-italic">H. edulis</span> (<b>A</b>) and <span class="html-italic">L. grisea</span> (<b>B</b>). The spectra were recorded at 300 K for samples in D<sub>2</sub>O solution. Chemical shifts are relative to external trimethylsilylpropionic acid at 0 ppm. The residual water has been suppressed by pre-saturation. The anomeric signals assigned by <sup>1</sup>H/<sup>13</sup>C HSQC (see <a href="#marinedrugs-13-02063-f005" class="html-fig">Figure 5</a>) are labeled A<span class="html-italic">–</span>E in the sulfated fucans.</p>
Full article ">Figure 4
<p>Expansions of the TOCSY (<b>A</b>,<b>B</b>) and ROESY (<b>C</b>,<b>D</b>) spectra of two sulfated fucans from <span class="html-italic">H. edulis</span> and <span class="html-italic">L. grisea</span> (<b>B</b>,<b>D</b>).The TOCSY spectra (<b>A</b>,<b>B</b>)show some cross-peaks used in the assignment of the fucose residue, especially positions bearing sulfate esters. The ROESY spectra (<b>B</b>,<b>D</b>) show ROEs, the sequence-defining A1–B3, B1–C3, C1–D2, D1–A3 and E1–D4. The five fucose residues in the repeating unit are marked A–E as described in the legend of <a href="#marinedrugs-13-02063-f006" class="html-fig">Figure 6</a>.</p>
Full article ">Figure 5
<p><sup>1</sup>H/<sup>13</sup>C HSQC (<b>A</b>,<b>B</b>) and HMBC (<b>C</b>,<b>D</b>) spectra of two sulfated fucans from two sea cucumbers <span class="html-italic">H. edulis</span> (<b>A</b>,<b>C</b>) and <span class="html-italic">L. grisea</span> (<b>B</b>,<b>D</b>). The assignments were based on TOCSY and COSY spectra. The anomeric signals were identified by the characteristic carbon chemical shifts and are marked A–E. The HMBC spectra (<b>C</b>,<b>D</b>) also show the sequence-defining A1–B3, B1–C3, C1–D2, D/D'1–A3 and E1–D4. The five fucose residues in the repeating unit are marked A–E as described in the legends of <a href="#marinedrugs-13-02063-f006" class="html-fig">Figure 6</a>.</p>
Full article ">Figure 6
<p>Proposed regular repeating units of sulfated fucan isolated from two sea cucumbers <span class="html-italic">H. edulis</span> (<b>A</b>) and <span class="html-italic">L. grisea</span> (<b>B</b>). The five fucose residues in the repeating unit are marked A–E as described in the legends.</p>
Full article ">Figure 7
<p>Inhibitory effects of the sulfated fucans, heparin, low molecular weight heparin (LMWH) and dermatan sulfate (DS) on thrombin mediated by heparin cofactor II. (<b>A</b>) Shows the time course of thrombin inhibition. HCII (~1 μM) was incubated with thrombin (20 NIH/mL) in the presence of 30 μL (625 ng/mL) samples at 37 °C. After 2 min, 30 μL of 4.5 mM CS-01 (38) was added, the residual thrombin activity was recorded by absorbance at 405 nm; (<b>B</b>) Shows the dependence on the sulfated polysaccharide concentration for thrombin inactivation in the presence of HCII. The reaction mixtures were as described in (<b>A</b>), except that different concentrations of sulfated polysaccharides were used. Results are shown as means of duplicates. See <a href="#marinedrugs-13-02063-t004" class="html-table">Table 4</a> for IC<sub>50</sub> values.</p>
Full article ">Figure 8
<p>Inhibitory effects of the sulfated fucans, heparin, LMWH and DS on thrombin in the presence of antithrombin. (<b>A</b>) Shows the time course of thrombin inhibition. Mixed samples of 30 μL of polysaccharides (625 ng/mL) and 30 μL of 0.25 IU/mL AT were incubated at 37 °C for 2 min, and 30 μL of 24 NIH/mL IIa was then added. After incubation for 2 min, 30 μL of 1.25 mM CS-01 (38) was added, the residual factor IIa activity was recorded by absorbance at 405 nm; (<b>B</b>) Shows the dependence on the sulfated polysaccharide concentration for thrombin inactivation mediated by AT. The reaction mixtures were as described in (<b>A</b>), except that different concentrations of sulfated polysaccharides were used. Results are shown as means of duplicates. See <a href="#marinedrugs-13-02063-t004" class="html-table">Table 4</a> for IC<sub>50</sub> values.</p>
Full article ">Figure 9
<p>Inhibitory effects of the sulfated fucans, heparin, LMWH and DS on factor Xa in the presence of antithrombin. (<b>A</b>) Shows the time course of Xa inhibition. Mixed samples of 30 μL of polysaccharides (625 ng/mL) and 30 μL of 1 IU/mL AT were incubated at 37 °C for 2 min, and 30 μL of 8 μg/mL bovine Xa was then added. After incubation for 1 min, 30 μL of 1.20 mM CS-11(65) was added, the residual Xa activity was recorded by absorbance at 405 nm; (<b>B</b>) Shows the dependence on the sulfated polysaccharide concentration for Xa inactivation in the presence of AT. The reaction mixtures were as described in (<b>A</b>), except that different concentrations of sulfated polysaccharides were used. Results are shown as means of duplicates. See <a href="#marinedrugs-13-02063-t004" class="html-table">Table 4</a> for IC<sub>50</sub> values.</p>
Full article ">Figure 10
<p>Profile of the platelet aggregation induced by the sea cucumber polysaccharides: the sulfated fucan from <span class="html-italic">L. grisea</span> (<b>A</b>); the sulfated fucan and fucosylated glycosaminoglycan from <span class="html-italic">H. edulis</span> (<b>B</b>). The profile showed that the sulfated fucans from sea cucumber do not cause platelets to aggregate at several concentrations.</p>
Full article ">
883 KiB  
Article
Prophylactic Administration of Fucoidan Represses Cancer Metastasis by Inhibiting Vascular Endothelial Growth Factor (VEGF) and Matrix Metalloproteinases (MMPs) in Lewis Tumor-Bearing Mice
by Tse-Hung Huang, Yi-Han Chiu, Yi-Lin Chan, Ya-Huang Chiu, Hang Wang, Kuo-Chin Huang, Tsung-Lin Li, Kuang-Hung Hsu and Chang-Jer Wu
Mar. Drugs 2015, 13(4), 1882-1900; https://doi.org/10.3390/md13041882 - 3 Apr 2015
Cited by 85 | Viewed by 12111
Abstract
Fucoidan, a heparin-like sulfated polysaccharide, is rich in brown algae. It has a wide assortment of protective activities against cancer, for example, induction of hepatocellular carcinoma senescence, induction of human breast and colon carcinoma apoptosis, and impediment of lung cancer cells migration and [...] Read more.
Fucoidan, a heparin-like sulfated polysaccharide, is rich in brown algae. It has a wide assortment of protective activities against cancer, for example, induction of hepatocellular carcinoma senescence, induction of human breast and colon carcinoma apoptosis, and impediment of lung cancer cells migration and invasion. However, the anti-metastatic mechanism that fucoidan exploits remains elusive. In this report, we explored the effects of fucoidan on cachectic symptoms, tumor development, lung carcinoma cell spreading and proliferation, as well as expression of metastasis-associated proteins in the Lewis lung carcinoma (LLC) cells-inoculated mice model. We discovered that administration of fucoidan has prophylactic effects on mitigation of cachectic body weight loss and improvement of lung masses in tumor-inoculated mice. These desired effects are attributed to inhibition of LLC spreading and proliferation in lung tissues. Fucoidan also down-regulates expression of matrix metalloproteinases (MMPs), nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB) and vascular endothelial growth factor (VEGF). Moreover, the tumor-bearing mice supplemented with fucoidan indeed benefit from an ensemble of the chemo-phylacticity. The fact is that fucoidan significantly decreases viability, migration, invasion, and MMPs activities of LLC cells. In summary, fucoidan is suitable to act as a chemo-preventative agent for minimizing cachectic symptoms as well as inhibiting lung carcinoma metastasis through down-regulating metastatic factors VEGF and MMPs. Full article
Show Figures

Figure 1

Figure 1
<p>Effects of fucoidan on body weight and lung mass in the LLC xenografted mouse model. At the 25th day, mice were sacrificed and examined for final gains of body weights (<b>A</b>) and lung masses (<b>B</b>); (<b>C</b>) The treatment protocol of fucoidan in tumor-bearing mice. Mice were fed orally with water or low- or high-dose of fucoidan (1 or 3 mg/mice) seven days before tumor implantation. Data are expressed as means ± SD (<span class="html-italic">n</span> = 6 mice per group; two independent experiments). Asterisk (*) stands for a significant difference when compared with the control group (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>Fucoidan reduces growth of lung tumor in mice. (<b>A</b>) Lung, pleural and bronchus tissues. 3 × 10<sup>5</sup> LLC cells were injected by tail vein in mice. Mice were sacrificed at the 25th day. The solid tumors (indicated by arrows) were spotted on multiple sites in mice; (<b>B</b>) Lungs were subjected to histological analysis (H&amp;E stain) for determining metastasis. Six representative samples are shown.</p>
Full article ">Figure 3
<p>Expression of metastatic proteins in sera of tumor bearing mice treated with fucoidan. (<b>A</b>) Western blot analyses of VEGF from representative mice. Expression levels of VEGF normalized to β-actin (<b>B</b>). Asterisk (*) indicates a significant difference (<span class="html-italic">p</span> value &lt; 0.05) when compared to the con group. Pound (#) indicates a significant difference (<span class="html-italic">p</span> value &lt; 0.05) when compared to TB-Con.</p>
Full article ">Figure 4
<p>Expression of metastatic proteins in lung tissues of tumor bearing mice treated with fucoidan. (<b>A</b>) Western blot analyses of MMP-2, NF-κB and VEGF (from representative samples); (<b>B</b>) Protein expression levels that are quantified and expressed as a fold-change relative to the control. Asterisk (*) indicates a significant difference (<span class="html-italic">p</span> value &lt; 0.05) when compared to the con group. Pound (#) indicates a significant difference (<span class="html-italic">p</span> value &lt; 0.05) when compared to TB-Con; (<b>C</b>) Immunofluorescence analysis for lung tumors treated and untreated (TB-Con) with fucoidan. Images are shown at 200× magnification.</p>
Full article ">Figure 4 Cont.
<p>Expression of metastatic proteins in lung tissues of tumor bearing mice treated with fucoidan. (<b>A</b>) Western blot analyses of MMP-2, NF-κB and VEGF (from representative samples); (<b>B</b>) Protein expression levels that are quantified and expressed as a fold-change relative to the control. Asterisk (*) indicates a significant difference (<span class="html-italic">p</span> value &lt; 0.05) when compared to the con group. Pound (#) indicates a significant difference (<span class="html-italic">p</span> value &lt; 0.05) when compared to TB-Con; (<b>C</b>) Immunofluorescence analysis for lung tumors treated and untreated (TB-Con) with fucoidan. Images are shown at 200× magnification.</p>
Full article ">Figure 5
<p>Effects of fucoidan on cell viability in African green monkey kidney Vero and mouse Lewis lung carcinoma cells. Cells were incubated in a culture medium containing various concentrations of fucoidan for 24 h. After the treatment, cell viability was determined by the MTS assay. Values relative to that of vehicle control were determined, in which the cell viability of control is set as 100%. Data (each value is an average of at least three independent experiments (six tests)) are presented as mean ± SEM. Asterisk (*) indicates a significant difference (<span class="html-italic">p</span> value &lt; 0.05) relative to the vehicle-treated cells.</p>
Full article ">Figure 6
<p>Suppression of migration and invasion of lung adenocarcinoma cells by fucoidan. (<b>A</b>) Representative photographs of three independent experiments, showing a dose-dependent inhibition of migration after treatment of fucoidan (24 h). Images of wound closures (10× magnification); (<b>B</b>) Black dotted lines indicate the wound edge. The cell-free areas invaded by cells (across the black dotted lines) were quantified by three random fields as shown in the lower panels; (<b>C</b>) The invasiveness of the LLC cells were quantified by counting the stained cells that invade into the porous polycarbonate membrane; (<b>D</b>) Invasiveness of the LLC cells treated with fucoidan. The LLC cells were pretreated with fucoidan for 24 h and then seeded onto the transwell chamber. Photographs were taken by an inverted microscope with 10× magnification. Data were derived from three independent experiments and presented as mean ± SEM. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.001, *** <span class="html-italic">p</span> &lt; 0.0001 when compared to the vehicle-treated cells.</p>
Full article ">Figure 7
<p>Enzymatic activities of MMP-2 and MMP-9 in the LLC cell lines treated with fucoidan. (<b>A</b>) The activity of MMPs was determined by the gelatinase zymography, in which the bright zones stand for gelatin digested; (<b>B</b>) The MMPs activity was quantified by measuring the band intensity in the zymography. Data were derived from three independent experiments and presented as mean ± SEM. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.001, *** <span class="html-italic">p</span> &lt; 0.0001 when compared to the vehicle-treated cells.</p>
Full article ">
425 KiB  
Review
Recent Advances in Exopolysaccharides from Paenibacillus spp.: Production, Isolation, Structure, and Bioactivities
by Tzu-Wen Liang and San-Lang Wang
Mar. Drugs 2015, 13(4), 1847-1863; https://doi.org/10.3390/md13041847 - 1 Apr 2015
Cited by 89 | Viewed by 9331
Abstract
This review provides a comprehensive summary of the most recent developments of various aspects (i.e., production, purification, structure, and bioactivity) of the exopolysaccharides (EPSs) from Paenibacillus spp. For the production, in particular, squid pen waste was first utilized successfully to produce [...] Read more.
This review provides a comprehensive summary of the most recent developments of various aspects (i.e., production, purification, structure, and bioactivity) of the exopolysaccharides (EPSs) from Paenibacillus spp. For the production, in particular, squid pen waste was first utilized successfully to produce a high yield of inexpensive EPSs from Paenibacillus sp. TKU023 and P. macerans TKU029. In addition, this technology for EPS production is prevailing because it is more environmentally friendly. The Paenibacillus spp. EPSs reported from various references constitute a structurally diverse class of biological macromolecules with different applications in the broad fields of pharmacy, cosmetics and bioremediation. The EPS produced by P. macerans TKU029 can increase in vivo skin hydration and may be a new source of natural moisturizers with potential value in cosmetics. However, the relationships between the structures and activities of these EPSs in many studies are not well established. The contents and data in this review will serve as useful references for further investigation, production, structure and application of Paenibacillus spp. EPSs in various fields. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic diagram for the isolation of EPS from <span class="html-italic">Paenibacillus</span> spp.</p>
Full article ">
746 KiB  
Review
Stability of Chitosan—A Challenge for Pharmaceutical and Biomedical Applications
by Emilia Szymańska and Katarzyna Winnicka
Mar. Drugs 2015, 13(4), 1819-1846; https://doi.org/10.3390/md13041819 - 1 Apr 2015
Cited by 621 | Viewed by 18058
Abstract
Chitosan—one of the natural multifunctional polymers—due to its unique and versatile biological properties is regarded as a useful compound in medical and pharmaceutical technology. Recently, considerable research effort has been made in order to develop safe and efficient chitosan products. However, the problem [...] Read more.
Chitosan—one of the natural multifunctional polymers—due to its unique and versatile biological properties is regarded as a useful compound in medical and pharmaceutical technology. Recently, considerable research effort has been made in order to develop safe and efficient chitosan products. However, the problem of poor stability of chitosan-based systems restricts its practical applicability; thus, it has become a great challenge to establish sufficient shelf-life for chitosan formulations. Improved stability can be assessed by controlling the environmental factors, manipulating processing conditions (e.g., temperature), introducing a proper stabilizing compound, developing chitosan blends with another polymer, or modifying the chitosan structure using chemical or ionic agents. This review covers the influence of internal, environmental, and processing factors on the long-term stability of chitosan products. The aim of this paper is also to highlight the latest developments which enable the physicochemical properties of chitosan-based applications to be preserved upon storage. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Factors affecting stability of chitosan-based products.</p>
Full article ">Figure 2
<p>Possible degradation mechanisms of chitosan’s structure (adapted from [<a href="#B32-marinedrugs-13-01819" class="html-bibr">32</a>] with modifications).</p>
Full article ">Figure 3
<p>Strategies to improve the stability of chitosan-based products.</p>
Full article ">
1375 KiB  
Article
Morphological and Proteomic Analyses Reveal that Unsaturated Guluronate Oligosaccharide Modulates Multiple Functional Pathways in Murine Macrophage RAW264.7 Cells
by Xu Xu, De-Cheng Bi, Chao Li, Wei-Shan Fang, Rui Zhou, Shui-Ming Li, Lian-Li Chi, Min Wan and Li-Ming Shen
Mar. Drugs 2015, 13(4), 1798-1818; https://doi.org/10.3390/md13041798 - 30 Mar 2015
Cited by 34 | Viewed by 7908
Abstract
Alginate is a natural polysaccharide extracted from various species of marine brown algae. Alginate-derived guluronate oligosaccharide (GOS) obtained by enzymatic depolymerization has various pharmacological functions. Previous studies have demonstrated that GOS can trigger the production of inducible nitric oxide synthase (iNOS)/nitric oxide (NO), [...] Read more.
Alginate is a natural polysaccharide extracted from various species of marine brown algae. Alginate-derived guluronate oligosaccharide (GOS) obtained by enzymatic depolymerization has various pharmacological functions. Previous studies have demonstrated that GOS can trigger the production of inducible nitric oxide synthase (iNOS)/nitric oxide (NO), reactive oxygen species (ROS) and tumor necrosis factor (TNF)-α by macrophages and that it is involved in the nuclear factor (NF)-κB and mitogen-activated protein (MAP) kinase signaling pathways. To expand upon the current knowledge regarding the molecular mechanisms associated with the GOS-induced immune response in macrophages, comparative proteomic analysis was employed together with two-dimensional electrophoresis (2-DE), matrix-assisted laser desorption/ionization time-of-flight mass spectrometry (MALDI-TOF/TOF MS) and Western blot verification. Proteins showing significant differences in expression in GOS-treated cells were categorized into multiple functional pathways, including the NF-κB signaling pathway and pathways involved in inflammation, antioxidant activity, glycolysis, cytoskeletal processes and translational elongation. Moreover, GOS-stimulated changes in the morphologies and actin cytoskeleton organization of RAW264.7 cells were also investigated as possible adaptations to GOS. This study is the first to reveal GOS as a promising agent that can modulate the proper balance between the pro- and anti-inflammatory immune responses, and it provides new insights into pharmaceutical applications of polysaccharides. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The infrared (IR) spectrum of polyguluronic acid (PG). The spectrum was run in KBr pellets with 1 mg of sample and 200 mg of KBr.</p>
Full article ">Figure 2
<p>The electrospray ionization mass spectrometry (ESI-MS) of guluronate oligosaccharide (GOS). The spectrum was acquired in the negative ion mode with a high-resolution hybrid time-of-flight mass spectrometer. The ions were present in the form of [M + <span class="html-italic">x</span>Na(K) − (<span class="html-italic">x</span> + <span class="html-italic">n</span>)H]<sup><span class="html-italic">n</span>−</sup>. The corresponding molecular weights and DP were then calculated using the monoisotopic peaks and charge states of each group of ions.</p>
Full article ">Figure 3
<p>(<b>A</b>) A representative two-dimensional electrophoresis (2-DE) gel showing of total proteins from untreated and GOS-treated RAW264.7 cells. Each gel is representative of three independent replicates. Differentially regulated proteins (≥2.0-fold) are indicated by arrows and numbers and are listed in <a href="#marinedrugs-13-01798-t002" class="html-table">Table 2</a>; (<b>B</b>) Magnified image of an identified protein spot.</p>
Full article ">Figure 4
<p>Western blot analysis of altered proteins in GOS-treated RAW264.7 cells. A representative result of three independent experiments is shown. Typical experiment conducted three times with similar results. β-tubulin was used as an internal control.</p>
Full article ">Figure 5
<p>Effects of GOS on the morphology and actin organization of RAW264.7 cells. RAW264.7 cell morphology was observed by dark-field and confocal microscopy (×40). The cells were treated with or without 1 mg/mL GOS for 24 h. Representative dark-field images and analysis results show that GOS induced morphological changes in RAW264.7 cells, including an increase in the cell number (<b>A</b>,<b>B</b>), larger cell sizes (<b>C</b>,<b>D</b>), extended nucleus areas (<b>E</b>,<b>F</b>), and a greater number of dual-nuclei cells (<b>G</b>,<b>H</b>), compared with untreated cells (control). F-actin was stained with FITC-phalloidin. GOS-induced F-actin organization was examined using fluorescence images (<b>I</b>) and quantitative analysis (<b>J</b>). Scale bar, 20 µm. All images were analyzed with ImageJ software. ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001 indicate significant differences between the control group and the GOS-treated group.</p>
Full article ">Figure 5 Cont.
<p>Effects of GOS on the morphology and actin organization of RAW264.7 cells. RAW264.7 cell morphology was observed by dark-field and confocal microscopy (×40). The cells were treated with or without 1 mg/mL GOS for 24 h. Representative dark-field images and analysis results show that GOS induced morphological changes in RAW264.7 cells, including an increase in the cell number (<b>A</b>,<b>B</b>), larger cell sizes (<b>C</b>,<b>D</b>), extended nucleus areas (<b>E</b>,<b>F</b>), and a greater number of dual-nuclei cells (<b>G</b>,<b>H</b>), compared with untreated cells (control). F-actin was stained with FITC-phalloidin. GOS-induced F-actin organization was examined using fluorescence images (<b>I</b>) and quantitative analysis (<b>J</b>). Scale bar, 20 µm. All images were analyzed with ImageJ software. ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001 indicate significant differences between the control group and the GOS-treated group.</p>
Full article ">Figure 6
<p>Effects of GOS on the lipopolysaccharide (LPS)-activated morphological changes in RAW264.7 cells. The morphologies of RAW264.7 cells were visualized with a phase contrast microscope (×40). The cells were pretreated with 1 mg/mL GOS for 2 h before incubation with 1 μg/mL LPS for 24 h. (<b>A</b>) Control; (<b>B</b>) LPS-treated only; and (<b>C</b>) LPS-treated with GOS. Scale bar, 20 µm.</p>
Full article ">
1273 KiB  
Article
Structural Analysis and Anti-Complement Activity of Polysaccharides from Kjellmaniella crsaaifolia
by Wenjing Zhang, Weihua Jin, Delin Sun, Luyu Zhao, Jing Wang, Delin Duan and Quanbin Zhang
Mar. Drugs 2015, 13(3), 1360-1374; https://doi.org/10.3390/md13031360 - 16 Mar 2015
Cited by 17 | Viewed by 6489
Abstract
Two polysaccharides, named KCA and KCW, were extracted from Kjellmaniella crassifolia using dilute hydrochloric acid and water, respectively. Composition analysis showed that these polysaccharides predominantly consisted of fucose, with galactose, mannose and glucuronic acid as minor components. After degradation and partial desulfation, electrospray [...] Read more.
Two polysaccharides, named KCA and KCW, were extracted from Kjellmaniella crassifolia using dilute hydrochloric acid and water, respectively. Composition analysis showed that these polysaccharides predominantly consisted of fucose, with galactose, mannose and glucuronic acid as minor components. After degradation and partial desulfation, electrospray ionization mass spectrometry (ESI-MS) was performed, which showed that the polysaccharides consisted of sulfated fucooligosaccharides, sulfated galactofucooligosaccharides and methyl glycosides of mono-sulfated/multi-sulfated fucooligosaccharides. The structures of the oligomeric fragments were further characterized by electrospray ionization collision-induced dissociation tandem mass spectrometry (ESI-CID-MS2 and ESI-CID-MS3). Moreover, the activity of KCA and KCW against the hemolytic activity of both the classical and alternative complement pathways was determined. The activity of KCA was found to be similar to KCW, suggesting that the method of extraction did not influence the activity. In addition, the degraded polysaccharides (DKCA and DKCW) displayed lower activity levels than the crude polysaccharides (KCA and KCW), indicating that molecular weight had an effect on activity. Moreover, the desulfated fractions (ds-DKCA and ds-DKCW) showed less or no activity, which confirmed that sulfate was important for activity. In conclusion, polysaccharides from K. crassifolia may be good candidates for the treatment of diseases involving the complement pathway. Full article
Show Figures

Figure 1

Figure 1
<p>The IR spectra of polysaccharides.</p>
Full article ">Figure 2
<p>Negative ion mode ESI-MS spectra of ds-DKCW (<b>a</b>) and ds-DKCA (<b>b</b>).</p>
Full article ">Figure 3
<p>Negative ion mode ESI-CID-MS<sup>2</sup> spectrum of the ion [MeFuc<sub>7</sub>SO<sub>3</sub>Na-Na]<sup>−</sup> at <span class="html-italic">m/z</span> 1133.415 (−1).</p>
Full article ">Figure 4
<p>Negative ion mode ESI-CID-MS<sup>2</sup> spectrum of the ion at <span class="html-italic">m/z</span> 421.129 (−2) (<b>a</b>) and negative ion mode ESI-CID-MS<sup>3</sup> spectra of the ions at <span class="html-italic">m/z</span> 681.208(−1) (<b>b</b>) and 383.103 (−2) (<b>c</b>).</p>
Full article ">Figure 5
<p>Negative ion mode ESI-CID-MS<sup>2</sup> spectrum of the ion [MeFuc<sub>5</sub>(SO<sub>3</sub>Na)<sub>2</sub>-2Na]<sup>2−</sup> at <span class="html-italic">m/z</span> 460.120 (−2).</p>
Full article ">Figure 6
<p>Negative ion mode ESI-MS<sup>2</sup> spectrum of ion [Fuc<sub>6</sub>(SO<sub>3</sub>Na)<sub>2</sub>-2Na]<sup>2−</sup> at <span class="html-italic">m/z</span> 526.141 (−2).</p>
Full article ">Figure 7
<p>Inhibition of the classical pathway-mediated hemolysis of EA (<b>a</b> and <b>b</b>) and alternative pathway-mediated hemolysis of ER (<b>c</b> and <b>d</b>) in 1:10-diluted NHS in the presence of increasing amounts of the polysaccharides. Heparin was used as the reference. The results are expressed as percent inhibition of hemolysis. Data are the means from 3 determinations ± S.E.M.</p>
Full article ">
914 KiB  
Article
Acetylated Chitosan Oligosaccharides Act as Antagonists against Glutamate-Induced PC12 Cell Death via Bcl-2/Bax Signal Pathway
by Cui Hao, Lixia Gao, Yiran Zhang, Wei Wang, Guangli Yu, Huashi Guan, Lijuan Zhang and Chunxia Li
Mar. Drugs 2015, 13(3), 1267-1289; https://doi.org/10.3390/md13031267 - 12 Mar 2015
Cited by 40 | Viewed by 9007
Abstract
Chitosan oligosaccharides (COSs), depolymerized products of chitosan composed of β-(1→4) d-glucosamine units, have broad range of biological activities such as antitumour, antifungal, and antioxidant activities. In this study, peracetylated chitosan oligosaccharides (PACOs) and N-acetylated chitosan oligosaccharides (NACOs) were prepared from the COSs [...] Read more.
Chitosan oligosaccharides (COSs), depolymerized products of chitosan composed of β-(1→4) d-glucosamine units, have broad range of biological activities such as antitumour, antifungal, and antioxidant activities. In this study, peracetylated chitosan oligosaccharides (PACOs) and N-acetylated chitosan oligosaccharides (NACOs) were prepared from the COSs by chemcal modification. The structures of these monomers were identified using NMR and ESI-MS spectra. Their antagonist effects against glutamate-induced PC12 cell death were investigated. The results showed that pretreatment of PC12 cells with the PACOs markedly inhibited glutamate-induced cell death in a concentration-dependent manner. The PACOs were better glutamate antagonists compared to the COSs and the NACOs, suggesting the peracetylation is essential for the neuroprotective effects of chitosan oligosaccharides. In addition, the PACOs pretreatment significantly reduced lactate dehydrogenase release and reactive oxygen species production. It also attenuated the loss of mitochondrial membrane potential. Further studies indicated that the PACOs inhibited glutamate-induced cell death by preventing apoptosis through depressing the elevation of Bax/Bcl-2 ratio and caspase-3 activation. These results suggest that PACOs might be promising antagonists against glutamate-induced neural cell death. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of the repeating saccharide units in chitosan oligosaccharide and its acetylated derivatives. (A–C) Chemical structure of chitosan oligosaccharides (COS) (<b>A</b>), <span class="html-italic">N</span>-acetylated chitosan oligosaccharides (NACO) (<b>B</b>), and peracetylated chitosan oligosaccharides (PACO) (<b>C</b>). The degree of polymerization (DP) of COS, NACO, and PACO is 2~4.</p>
Full article ">Figure 2
<p>Effect of glutamate on undifferentiated and fully-differentiated PC12 cells. (A–B) The undifferentiated (<b>A</b>) and fully-differentiated (<b>B</b>) PC12 cells were plated on the cell culture plates at a density of 1 × 10<sup>5</sup> cells/mL, and then treated with glutamate at different concentrations (0.25, 0.5, 1, 2, 4, 8, 16 and 32 mM) for 24 h. Then the cell viability was evaluated by resazurin assay. The results were presented as a percentage of the normal control group. Values are the mean ± SD (<span class="html-italic">n</span> = 3). Significance: * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 <span class="html-italic">vs.</span> normal control group without glutamate.</p>
Full article ">Figure 3
<p>The cytotoxicity of different chitosan oligosaccharides in PC12 cells. PC12 cells were incubated with different chitosan oligosaccharides COS-2~4 (<b>A</b>), peracetylated chitosan oligosaccharides Q-2~4 (<b>B</b>) and <span class="html-italic">N</span>-acetylated chitosan oligosaccharides N-2~4 (<b>C</b>) at indicated concentrations (100, 200, 400 μg/mL) for 24 h. Then the cell viability was evaluated by resazurin assay. The results were presented as a percentage of non drug treated normal control group. Values are the mean ± SD (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 4
<p>The effect of different chitosan oligosaccharides on glutamate-induced PC12 cell damage. PC12 cells were treated with or without different chitosan oligosaccharides COS-2~4 (<b>A</b>) and peracetylated chitosan oligosaccharides Q-2~4 (<b>B</b>) at indicated concentrations for 2 h. Then cells were treated with glutamate for another 24 h before performing a resazurin assay. The untreated normal cells (control) were assigned values of 100 and the results presented as mean ± SD (<span class="html-italic">n</span> = 4). Significance: # <span class="html-italic">p</span> &lt; 0.05 <span class="html-italic">vs.</span> normal control group; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 <span class="html-italic">vs.</span> glutamate treated control group.</p>
Full article ">Figure 5
<p>The neuroprotective effects of different acetylated neutral oligosaccharides in PC12 cells. (<b>A</b>) PC12 cells were treated with or without different N-acetylated chitosan oligosaccharides N-2~4 at indicated concentrations for 2 h. Then cells were treated with glutamate for another 24 h before performing a resazurin assay. The untreated normal cells (control) were assigned values of 100 and the results presented as mean ± SD (<span class="html-italic">n</span> = 4). Significance: # <span class="html-italic">p</span> &lt; 0.05 <span class="html-italic">vs.</span> normal control group; * <span class="html-italic">p</span> &lt; 0.05 <span class="html-italic">vs.</span> glutamate treated control group. (<b>B</b>) PC12 cells were treated with or without peracetylated chitobiose Q-2, lactose, acetylated lactose, cellobiose, or acetylated cellobiose at indicated concentrations for 2 h. Then cells were treated with glutamate for another 24 h. The untreated normal cells (control) were assigned values of 100 and the results presented as mean ± SD (<span class="html-italic">n</span> = 3). Significance: ## <span class="html-italic">p</span> &lt; 0.01 <span class="html-italic">vs.</span> normal control group; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 <span class="html-italic">vs.</span> glutamate treated control group.</p>
Full article ">Figure 5 Cont.
<p>The neuroprotective effects of different acetylated neutral oligosaccharides in PC12 cells. (<b>A</b>) PC12 cells were treated with or without different N-acetylated chitosan oligosaccharides N-2~4 at indicated concentrations for 2 h. Then cells were treated with glutamate for another 24 h before performing a resazurin assay. The untreated normal cells (control) were assigned values of 100 and the results presented as mean ± SD (<span class="html-italic">n</span> = 4). Significance: # <span class="html-italic">p</span> &lt; 0.05 <span class="html-italic">vs.</span> normal control group; * <span class="html-italic">p</span> &lt; 0.05 <span class="html-italic">vs.</span> glutamate treated control group. (<b>B</b>) PC12 cells were treated with or without peracetylated chitobiose Q-2, lactose, acetylated lactose, cellobiose, or acetylated cellobiose at indicated concentrations for 2 h. Then cells were treated with glutamate for another 24 h. The untreated normal cells (control) were assigned values of 100 and the results presented as mean ± SD (<span class="html-italic">n</span> = 3). Significance: ## <span class="html-italic">p</span> &lt; 0.01 <span class="html-italic">vs.</span> normal control group; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 <span class="html-italic">vs.</span> glutamate treated control group.</p>
Full article ">Figure 6
<p>Effect of peracetylated chitosan oligosaccharides on glutamate-induced LDH release and ROS overproduction. (<b>A</b>) After the treatment of cells with different PACO monomers Q-2, Q-3 and Q-4 at a concentration of 200 μg/mL for 2 h and 4 mM of glutamate for 24 h, the level of LDH in the culture media was measured using an LDH assay kit. The data were normalized to the activity of LDH released from control cells. Values are the mean ± SD (<span class="html-italic">n</span> = 4). Significance: ## <span class="html-italic">P</span> &lt; 0.01 <span class="html-italic">vs.</span> normal control group; * <span class="html-italic">P</span> &lt; 0.05, ** <span class="html-italic">P</span> &lt; 0.01 <span class="html-italic">vs.</span> glutamate treated control group; (<b>B</b>) After the treatment of cells with different PACO monomers Q-2, Q-3 and Q-4 at the concentration of 200 μg/mL for 2 h and 4 mM of glutamate for another 24 h. The fluorescence intensity of DCF was measured in a microplate-reader. Data were expressed as a percentage of non-treated control. Values are the mean ± SD (<span class="html-italic">n</span> = 4). Significance: ## <span class="html-italic">P</span> &lt; 0.01 <span class="html-italic">vs.</span> normal control group; ** <span class="html-italic">P</span> &lt; 0.01 <span class="html-italic">vs.</span> glutamate treated control group.</p>
Full article ">Figure 7
<p>Peracetylated chitosan oligosaccharides protect PC12 cells against glutamate-induced loss of MMP and the activation of Caspase-3 and Caspase-9. (<b>A</b>) After being pretreated with 200 μg/mL of different PACO monomers Q-2, Q-3 and Q-4 for 2 h and 4 mM of glutamate for another 24 h, the mitochondrial membrane potential (MMP) in PC12 cells were evaluated with the probe JC-1. Data were expressed as a percentage of non-treated control. Values are the mean ± SD (<span class="html-italic">n</span> = 3). Significance: # <span class="html-italic">p</span> &lt; 0.05 <span class="html-italic">vs.</span> normal control group; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 <span class="html-italic">vs.</span> glutamate treated control group; (<b>B</b>) After being pretreated with 200 μg/mL of different PACO monomers Q-2, Q-3 and Q-4 for 2 h and 4 mM of glutamate for another 24 h, the levels of cleaved caspase-3 were measured by western blot. Blots were also probed for β-actin protein as loading controls. The result shown is a representative of three separate experiments with similar results. (<b>C</b>) Quantification of immunoblot for the ratio of caspase-3 to β-actin. The ratio for non-treated normal control cells was assigned values of 1.0 and the data presented as mean ± SD (<span class="html-italic">n</span> = 3). Significance: ## <span class="html-italic">p</span> &lt; 0.01 <span class="html-italic">vs.</span> normal control group; ** <span class="html-italic">p</span> &lt; 0.01 <span class="html-italic">vs.</span> glutamate treated control group. (D–E) After treatment, the activities of caspase-3 (<b>D</b>) and caspase-9 (<b>E</b>) were measured using an ELISA assay kit (Beyotime, China). Data were expressed as a percentage of non-treated control. Values are the mean ± SD (<span class="html-italic">n</span> = 3). Significance: # <span class="html-italic">p</span> &lt; 0.05 <span class="html-italic">vs.</span> normal control group; * <span class="html-italic">p</span> &lt; 0.05 <span class="html-italic">vs.</span> glutamate treated control group.</p>
Full article ">Figure 7 Cont.
<p>Peracetylated chitosan oligosaccharides protect PC12 cells against glutamate-induced loss of MMP and the activation of Caspase-3 and Caspase-9. (<b>A</b>) After being pretreated with 200 μg/mL of different PACO monomers Q-2, Q-3 and Q-4 for 2 h and 4 mM of glutamate for another 24 h, the mitochondrial membrane potential (MMP) in PC12 cells were evaluated with the probe JC-1. Data were expressed as a percentage of non-treated control. Values are the mean ± SD (<span class="html-italic">n</span> = 3). Significance: # <span class="html-italic">p</span> &lt; 0.05 <span class="html-italic">vs.</span> normal control group; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 <span class="html-italic">vs.</span> glutamate treated control group; (<b>B</b>) After being pretreated with 200 μg/mL of different PACO monomers Q-2, Q-3 and Q-4 for 2 h and 4 mM of glutamate for another 24 h, the levels of cleaved caspase-3 were measured by western blot. Blots were also probed for β-actin protein as loading controls. The result shown is a representative of three separate experiments with similar results. (<b>C</b>) Quantification of immunoblot for the ratio of caspase-3 to β-actin. The ratio for non-treated normal control cells was assigned values of 1.0 and the data presented as mean ± SD (<span class="html-italic">n</span> = 3). Significance: ## <span class="html-italic">p</span> &lt; 0.01 <span class="html-italic">vs.</span> normal control group; ** <span class="html-italic">p</span> &lt; 0.01 <span class="html-italic">vs.</span> glutamate treated control group. (D–E) After treatment, the activities of caspase-3 (<b>D</b>) and caspase-9 (<b>E</b>) were measured using an ELISA assay kit (Beyotime, China). Data were expressed as a percentage of non-treated control. Values are the mean ± SD (<span class="html-italic">n</span> = 3). Significance: # <span class="html-italic">p</span> &lt; 0.05 <span class="html-italic">vs.</span> normal control group; * <span class="html-italic">p</span> &lt; 0.05 <span class="html-italic">vs.</span> glutamate treated control group.</p>
Full article ">Figure 8
<p>Peracetylated chitosan oligosaccharides protect PC12 cells against glutamate-induced Cyto C release from mitochondria. (<b>A</b>) After being pretreated with 200 μg/mL of different PACO monomers Q-2, Q-3 and Q-4 for 2 h and 4 mM of glutamate for another 24 h, the protein levels of Cyto C were evaluated by western blot. Blots were also probed for β-actin as loading controls. The result shown is a representative of three separate experiments with similar results; (<b>B</b>) Quantification of immunoblot for the ratio of Cyto C to β-actin. The ratio for non-treated control cells was assigned values of 1.0 and the data presented as mean ± SD (<span class="html-italic">n</span> = 3). Significance: ## <span class="html-italic">p</span> &lt; 0.01 <span class="html-italic">vs.</span> normal control group; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 <span class="html-italic">vs.</span> glutamate treated control group. (<b>C</b>–<b>H</b>) After being pretreated with 200 μg/mL of PACO monomers Q-2, Q-3 and Q-4 for 2 h, PC12 cells were exposed to 4 mM glutamate for 24 h. Then the levels of Cyto C in the cytoplasm were detected by immunofluorescence assay using anti-Cyto C antibody. <b>C</b>: Normal control, <b>D</b>: Glu, <b>E</b>: Glu + HupA, <b>F</b>: Glu + Q2, G: Glu + Q3, <b>H</b>: Glu + Q4. Scale bar represents 20 μm.</p>
Full article ">Figure 9
<p>Effect of PACOs on the expression of Bax and Bcl-2 in PC12 cells. (A–B) After being pretreated with 200 μg/mL of different PACO monomers Q-2, Q-3 and Q-4 for 2 h and 4 mM of glutamate for another 24 h, the levels of Bax (<b>A</b>) and Bcl-2 (<b>B</b>) were measured by western blot. Blots were also probed for β-actin as loading controls. The result shown is a representative of three separate experiments with similar results; (<b>C</b>) Quantification of immunoblot for the ratio of Bax and Bcl-2. The ratio for non-treated normal control cells was assigned values of 1.0. Values are the mean ± SD (<span class="html-italic">n</span> = 3). Significance: ## <span class="html-italic">p</span> &lt; 0.01 <span class="html-italic">vs.</span> normal control group; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 <span class="html-italic">vs.</span> glutamate treated control group.</p>
Full article ">Figure 10
<p>Proposed schemas of the mechanisms by which PACO suppressed glutamate-induced apoptosis in PC12 cells. Glutamate induces intracellular ROS generation in PC12 cells. The resultant oxidative stress triggers the activation of Bax protein, thereby reducing ΔΨm, releasing cytochrome c, activating caspase 3, and finally inducing apoptosis. PACO inhibits ROS and suppresses the activity of downstream molecules, such as Bax, Cyto c, and caspase 3, for promoting neuronal survival.</p>
Full article ">
1198 KiB  
Article
Conformational Analysis of the Oligosaccharides Related to Side Chains of Holothurian Fucosylated Chondroitin Sulfates
by Alexey G. Gerbst, Andrey S. Dmitrenok, Nadezhda E. Ustyuzhanina and Nikolay E. Nifantiev
Mar. Drugs 2015, 13(2), 936-947; https://doi.org/10.3390/md13020936 - 12 Feb 2015
Cited by 7 | Viewed by 5992
Abstract
Anionic polysaccharides fucosylated chondroitin sulfates (FCS) from holothurian species were shown to affect various biological processes, such as metastasis, angiogenesis, clot formation, thrombosis, inflammation, and some others. To understand the mechanism of FCSs action, knowledge about their spatial arrangement is required. We have [...] Read more.
Anionic polysaccharides fucosylated chondroitin sulfates (FCS) from holothurian species were shown to affect various biological processes, such as metastasis, angiogenesis, clot formation, thrombosis, inflammation, and some others. To understand the mechanism of FCSs action, knowledge about their spatial arrangement is required. We have started the systematic synthesis, conformational analysis, and study of biological activity of the oligosaccharides related to various fragments of these types of natural polysaccharides. In this communication, five molecules representing distinct structural fragments of chondroitin sulfate have been studied by means of molecular modeling and NMR. These are three disaccharides and two trisaccharides containing fucose and glucuronic acid residues with one sulfate group per each fucose residue or without it. Long-range C–H coupling constants were used for the verification of the theoretical models. The presence of two conformers for both linkage types was revealed. For the Fuc–GlA linkage, the dominant conformer was the same as described previously in a literature as the molecular dynamics (MD) average in a dodechasaccharide FCS fragment representing the backbone chain of the polysaccharide including GalNAc residues. This shows that the studied oligosaccharides, in addition to larger ones, may be considered as reliable models for Quantitative Structure-Activity Relationship (QSAR) studies to reveal pharmacophore fragments of FCS. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Branched fragments of fucosylated chondroitin sulfates and synthetic oligosaccharides related to the knots.</p>
Full article ">Figure 2
<p>Torsional angles describing a glycosidic linkage and the Karplus equation [<a href="#B17-marinedrugs-13-00936" class="html-bibr">17</a>]. Angles φ and ψ are defined as H<sub>1</sub>–C<sub>1</sub>–O–C<sub>x</sub> and C<sub>1</sub>–O–C<sub>x</sub>–H<sub>x</sub> correspondingly.</p>
Full article ">Figure 3
<p>Sample confomational maps obtained by means of torsion scanning for Fuc–GlA (<b>A</b>) and Fuc–Fuc (<b>B</b>) linkages.</p>
Full article ">Figure 4
<p>Molecular dynamics (MD) graphs and conformational plots obtained in the Solvent Accessible Surface Area (SASA) approximation for disaccharides <b>1</b>–<b>3</b> ordered from (<b>A</b>) to (<b>C</b>). Angles φ and ψ are defined as H<sub>1</sub>–C<sub>1</sub>–O–C<sub>x</sub> and C<sub>1</sub>–O–C<sub>x</sub>–H<sub>x</sub>, respectively, for the sake of compatibility with the Karplus equation.</p>
Full article ">Figure 5
<p>Conformational plots obtained in the SASA approximation for glycosidic linkages in structures <b>4</b> ((<b>A</b>), Fuc–GlA; (<b>B</b>), Fuc–Fuc) and <b>5</b> ((<b>C</b>), Fuc–GlA, (<b>D</b>), Fuc–Fuc). Angles φ and ψ are defined as H<sub>1</sub>–C<sub>1</sub>–O–C<sub>x</sub> and C<sub>1</sub>–O–C<sub>x</sub>–H<sub>x</sub>, respectively, for the sake of compatibility with the Karplus equation.</p>
Full article ">Figure 6
<p>Principal conformers of the Fuc(1→3)GlA linkage.</p>
Full article ">Figure 7
<p>Torsional angles graph and conformational plots obtained using explicit water approximation for the fucosyl–glucuronide (<b>A</b>) and difucoside (<b>B</b>) fragments. Data are extracted from simulations of compound <b>4</b>. Angles φ and ψ are defined as H<sub>1</sub>–C<sub>1</sub>–O–C<sub>x</sub> and C<sub>1</sub>–O–C<sub>x</sub>–H<sub>x</sub> correspondingly for the sake of compatibility with the Karplus equation.</p>
Full article ">Figure 8
<p>The dominant conformer of the Fuc(1→3)GlA linkage with the position of GalNAc residue introduction.</p>
Full article ">Figure 9
<p>The NOE build-up graph at different mixing times for compound <b>4</b>. The measured NOE values are for H1(Fuc)/H3(GlA) interaction.</p>
Full article ">

2014

Jump to: 2024, 2022, 2021, 2020, 2019, 2018, 2017, 2016, 2015, 2013, 2012, 2011, 2010

708 KiB  
Article
Preliminary Characterization, Antioxidant Properties and Production of Chrysolaminarin from Marine Diatom Odontella aurita
by Song Xia, Baoyan Gao, Aifen Li, Jihai Xiong, Ziqiang Ao and Chengwu Zhang
Mar. Drugs 2014, 12(9), 4883-4897; https://doi.org/10.3390/md12094883 - 23 Sep 2014
Cited by 93 | Viewed by 9681
Abstract
A new chrysolaminarin, named CL2, with a molecular mass of 7.75 kDa, was purified from the marine diatom, Odontella aurita, using DEAE-52 cellulose anion-exchange chromatography and Sephadex G-200 gel-filtration chromatography. The monosaccharide and structural analysis revealed that CL2 was a glucan mainly [...] Read more.
A new chrysolaminarin, named CL2, with a molecular mass of 7.75 kDa, was purified from the marine diatom, Odontella aurita, using DEAE-52 cellulose anion-exchange chromatography and Sephadex G-200 gel-filtration chromatography. The monosaccharide and structural analysis revealed that CL2 was a glucan mainly composed of glucose, which was linked by the β-d-(1→3) (main chain) and β-d-(1→6) (side chain) glycosidic bond, demonstrated by infrared spectroscopy (IR) and nuclear magnetic resonance (NMR). The antioxidant activity tests revealed that the CL2 presented stronger hydroxyl radical scavenging activity with increasing concentrations, but less was effective on reducing power analysis and scavenging 1,1-diphenyl-2-picrylhydrazyl (DPPH) radical. The influences of nitrogen concentration and light intensity on chrysolaminarin production of O. aurita were further investigated in a glass column photobioreactor, and a record high chrysolaminarin productivity of 306 mg L−1 day−1 was achieved. In conclusion, the chrysolaminarin CL2 from O. aurita may be explored as a natural antioxidant agent for application in aquaculture, food and pharmaceutical areas. Full article
Show Figures

Figure 1

Figure 1
<p>DEAE-cellulose column elution profile of crude polysaccharide from <span class="html-italic">O. aurita</span>.</p>
Full article ">Figure 2
<p>Sephadex G-200 gel-filtration chromatogram of the fraction CL1 (chrysolaminarin 1) obtained from DEAE-cellulose column elution.</p>
Full article ">Figure 3
<p>The FTIR spectra of CL2 from <span class="html-italic">O. aurita</span>.</p>
Full article ">Figure 4
<p>(<b>a</b>) <sup>1</sup>H-NMR and (<b>b</b>) <sup>13</sup>C-NMR spectrum of CL2 from <span class="html-italic">O. aurita</span> (NA: not assigned).</p>
Full article ">Figure 5
<p>Antioxidant assays for the chrysolaminarin CL2 from <span class="html-italic">O. aurita</span>. (<b>a</b>) Reducing power; (<b>b</b>) scavenging of DPPH radicals; (<b>c</b>) scavenging of hydroxyl radicals. Values are the means ± SD (<span class="html-italic">n</span> = 3). When error bars cannot be seen, the error is less than the size of the symbol.</p>
Full article ">Figure 6
<p>The biomass (<b>a</b>,<b>b</b>) and chrysolaminarin content (<b>c</b>,<b>d</b>) of <span class="html-italic">O. aurita</span> cultivated in the column photobioreactor under 100 (<b>a</b>,<b>c</b>) and 300 (<b>b</b>,<b>d</b>) μmol photons m<sup>−2</sup> s<sup>−1</sup> with a replete (18 mM) and deficient (6 mM) nitrate supply. Values are the means ± SD (<span class="html-italic">n</span> = 3). When error bars cannot be seen, the error is less than the size of the symbol.</p>
Full article ">Figure 7
<p>Isolation and purification procedure of chrysolaminarin from <span class="html-italic">O. aurita</span>.</p>
Full article ">

2013

Jump to: 2024, 2022, 2021, 2020, 2019, 2018, 2017, 2016, 2015, 2014, 2012, 2011, 2010

964 KiB  
Article
Isolation and Structural Characterization of a Novel Antioxidant Mannoglucan from a Marine Bubble Snail, Bullacta exarata (Philippi)
by Donghong Liu, Ningbo Liao, Xingqian Ye, Yaqin Hu, Dan Wu, Xin Guo, Jianjun Zhong, Jianyong Wu and Shiguo Chen
Mar. Drugs 2013, 11(11), 4464-4477; https://doi.org/10.3390/md11114464 - 11 Nov 2013
Cited by 17 | Viewed by 8223
Abstract
Bullacta exarata is one of the most economically important aquatic species in China, noted for not only its delicious taste and nutritional value, but also for its pharmacological activities. In order to explore its potential in medical applications, a mannoglucan designated as BEPS-IB [...] Read more.
Bullacta exarata is one of the most economically important aquatic species in China, noted for not only its delicious taste and nutritional value, but also for its pharmacological activities. In order to explore its potential in medical applications, a mannoglucan designated as BEPS-IB was isolated and purified from the foot muscle of B. exarata after papain digestion. Chemical composition analysis indicated BEPS-IB contained mainly d-glucose and d-mannose in a molar ratio of 1:0.52, with an average molecular weight of about 94 kDa. The linkage information was determined by methylation analysis, and the anomeric configuration and chain linkage were confirmed by IR and 2D NMR. The results indicated BEPS-IB was composed of Glcp6Manp heptasaccharide repeating unit in the backbone, with occasional branch chains of mannose residues (14%) occurring in the backbone mannose. Further antioxidant assay indicated BEPS-IB exhibited positive antioxidant activity in scavenging superoxide radicals and reducing power. This is the first report on the structure and bioactivity of the mannoglucan from the B. exarata. Full article
Show Figures

Figure 1

Figure 1
<p>Isolation of the polysaccharides present in the aqueous extract of <span class="html-italic">B. exarata</span>. The crude extract was fractionated by ion-exchange chromatography on a DEAE ion-exchange column (<b>a</b>) and the collected fraction was further purified by gel filtration chromatography on a Sephacryl S-300 HR column (<b>b</b>). Solid bars indicate the fractions collected. The molecular weight of polysaccharide fraction BEPS-IB (Mw = 94 kDa) was determined by HPLC on a TSK-Gel G4000 PWXL column, eluted with 0.2 mol/L NaCl at 0.5 mL/min (<b>c</b>). Range of molecular weight in kDa: I = 500; II = 66.9; III = 40.</p>
Full article ">Figure 2
<p>Infrared spectra of polysaccharide (BEPS-IB) from <span class="html-italic">B. exarata.</span></p>
Full article ">Figure 3
<p><sup>1</sup>H NMR (<b>a</b>) and <sup>13</sup>C NMR (<b>b</b>) spectrum (600 MHz, D<sub>2</sub>O, 60 °C) of BEPS-IB isolated from <span class="html-italic">B. exarata</span>.</p>
Full article ">Figure 4
<p>HMQC spectrum of BEPS-IB isolated from <span class="html-italic">B. exarata</span>.</p>
Full article ">Figure 5
<p>Proposed structural features of the BEPS-IB isolated from <span class="html-italic">B. exarata</span>.</p>
Full article ">Figure 6
<p>Antioxidant activity of the BEPS-IB. (<b>a</b>) Scavenging effects of BEPS-IB on superoxide radical (O<sub>2</sub><sup>•</sup>); (<b>b</b>) Reducing power. Values are means ± SD (<span class="html-italic">n</span> = 3). Significant differences from the control were evaluated using Student’s <span class="html-italic">t</span>-test: * <span class="html-italic">p</span> &lt; 0.05. Reducing power was expressed as a percentage of the activity shown by vitamin C.</p>
Full article ">
768 KiB  
Review
Chondroitin Sulfate, Hyaluronic Acid and Chitin/Chitosan Production Using Marine Waste Sources: Characteristics, Applications and Eco-Friendly Processes: A Review
by José Antonio Vázquez, Isabel Rodríguez-Amado, María Ignacia Montemayor, Javier Fraguas, María Del Pilar González and Miguel Anxo Murado
Mar. Drugs 2013, 11(3), 747-774; https://doi.org/10.3390/md11030747 - 11 Mar 2013
Cited by 210 | Viewed by 25412
Abstract
In the last decade, an increasing number of glycosaminoglycans (GAGs), chitin and chitosan applications have been reported. Their commercial demands have been extended to different markets, such as cosmetics, medicine, biotechnology, food and textiles. Marine wastes from fisheries and aquaculture are susceptible sources [...] Read more.
In the last decade, an increasing number of glycosaminoglycans (GAGs), chitin and chitosan applications have been reported. Their commercial demands have been extended to different markets, such as cosmetics, medicine, biotechnology, food and textiles. Marine wastes from fisheries and aquaculture are susceptible sources for polymers but optimized processes for their recovery and production must be developed to satisfy such necessities. In the present work, we have reviewed different alternatives reported in the literature to produce and purify chondroitin sulfate (CS), hyaluronic acid (HA) and chitin/chitosan (CH/CHs) with the aim of proposing environmentally friendly processes by combination of various microbial, chemical, enzymatic and membranes strategies and technologies. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Overview of chondroitin sulfate (CS) recovery and purification processes from marine cartilage by-products. SED: sediment, SUP: supernatant, PER: permeate and RET: retentate.</p>
Full article ">Figure 2
<p>Flowchart of purification methods to extract hyaluronic acid (HA) from vitreous humor (VH) of fish eyeball. SED: sediment, SUP: supernatant, PER: permeate and RET: retentate.</p>
Full article ">Figure 3
<p>Scheme of CH and CHs preparation from crustacean shell waste using chemical methods.</p>
Full article ">Figure 4
<p>Scheme of chitin and chitosan preparation from crustacean shell waste using eco-friendly methods.</p>
Full article ">

2012

Jump to: 2024, 2022, 2021, 2020, 2019, 2018, 2017, 2016, 2015, 2014, 2013, 2011, 2010

631 KiB  
Article
Effect of Fucoidan Extracted from Mozuku on Experimental Cartilaginous Tissue Injury
by Tomohiro Osaki, Koudai Kitahara, Yoshiharu Okamoto, Tomohiro Imagawa, Takeshi Tsuka, Yasunari Miki, Hitoshi Kawamoto, Hiroyuki Saimoto and Saburo Minami
Mar. Drugs 2012, 10(11), 2560-2570; https://doi.org/10.3390/md10112560 - 13 Nov 2012
Cited by 9 | Viewed by 7607
Abstract
We investigated the effect of fucoidan, a sulfated polysaccharide, on acceleration of healing of experimental cartilage injury in a rabbit model. An injured cartilage model was surgically created by introduction of three holes, one in the articular cartilage of the medial trochlea and [...] Read more.
We investigated the effect of fucoidan, a sulfated polysaccharide, on acceleration of healing of experimental cartilage injury in a rabbit model. An injured cartilage model was surgically created by introduction of three holes, one in the articular cartilage of the medial trochlea and two in the trochlear sulcus of the distal femur. Rabbits in three experimental groups (F groups) were orally administered fucoidan of seven different molecular weights (8, 50, 146, 239, 330, 400, or 1000 kD) for 3 weeks by screening. Control (C group) rabbits were provided water ad libitum. After the experimental period, macroscopic examination showed that the degree of filling in the fucoidan group was higher than that in the C group. Histologically, the holes were filled by collagen fiber and fibroblasts in the C group, and by chondroblasts and fibroblasts in the F groups. Image analysis of Alcian blue- and safranin O-stained F-group specimens showed increased production of glycosaminoglycans (GAGs) and proteoglycans (PGs), respectively. Some injured holes were well repaired both macroscopically and microscopically and were filled with cartilage tissues; cartilage matrices such as PGs and GAGs were produced in groups F 50, F 146, and F 239. Thus, fucoidan administration enhanced morphologically healing of cartilage injury. Full article
Show Figures

Figure 1

Figure 1
<p>Average defect restoration scores. Data are expressed as the average ± standard deviation. * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 2
<p>Light microscopy of tissues from each group (HE stain; 40×). (<b>A</b>) C group; (<b>B</b>) F 8 group; (<b>C</b>) F 50 group; (<b>D</b>) F 146 group; (<b>E</b>) F 239 group; (<b>F</b>) F 330 group; (<b>G</b>) F 400 group; (<b>H</b>) F 1000 group. (a) cartilage layer; (b) superficial layer of cancellous bone; (c) deeper layer of cancellous bone. * joint cavity.</p>
Full article ">Figure 3
<p>Light microscopy of tissues from each group (Alcian blue stain; ×40). (<b>A</b>) C group; (<b>B</b>) F 8 group; (<b>C</b>) F 50 group; (<b>D</b>) F 146 group; (<b>E</b>) F 239 group; (<b>F</b>) F 330 group; (<b>G</b>) F 400 group; (<b>H</b>) F 1000 group.</p>
Full article ">Figure 4
<p>Light microscopy of tissues from each group (safranin O stain; 40×). (<b>A</b>) C group; (<b>B</b>) F 8 group; (<b>C</b>) F 50 group; (<b>D</b>) F 146 group; (<b>E</b>) F 239 group, (<b>F</b>) F 330 group; (<b>G</b>) F 400 group; (<b>H</b>) F 1000 group.</p>
Full article ">Figure 5
<p>Image analysis of Alcian blue-stained specimens. Data are expressed as the average ± standard deviation. *<span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 6
<p>Image analysis of safranin O-stained specimens. The data are expressed as the average ± standard deviation. * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">
1964 KiB  
Article
Structural Analysis of a Heteropolysaccharide from Saccharina japonica by Electrospray Mass Spectrometry in Tandem with Collision-Induced Dissociation Tandem Mass Spectrometry (ESI-CID-MS/MS)
by Weihua Jin, Jing Wang, Sumei Ren, Ni Song and Quanbin Zhang
Mar. Drugs 2012, 10(10), 2138-2152; https://doi.org/10.3390/md10102138 - 25 Sep 2012
Cited by 49 | Viewed by 9448
Abstract
A fucoidan extracted from Saccharina japonica was fractionated by anion exchange chromatography. The most complex fraction F0.5 was degraded by dilute sulphuric acid and then separated by use of an activated carbon column. Fraction Y1 was fractionated by anion exchange and gel filtration [...] Read more.
A fucoidan extracted from Saccharina japonica was fractionated by anion exchange chromatography. The most complex fraction F0.5 was degraded by dilute sulphuric acid and then separated by use of an activated carbon column. Fraction Y1 was fractionated by anion exchange and gel filtration chromatography while Fraction Y2 was fractionated by gel filtration chromatography. The fractions were determined by ESI-MS and analyzed by ESI-CID-MS/MS. It was concluded that F0.5 had a backbone of alternating 4-linked GlcA and 2-linked Man with the first Man residue from the nonreducing end accidentally sulfated at C6. In addition, F0.5 had a 3-linked glucuronan, in accordance with a previous report by NMR. Some other structural characteristics included GlcA 1→3 Man 1→4 GlcA, Man 1→3 GlcA 1→4 GlcA, Fuc 1→4 GlcA and Fuc 1→3 Fuc. Finally, it was shown that fucose was sulfated at C2 or C4 while galactose was sulfated at C2, C4 or C6. Full article
Show Figures

Figure 1

Figure 1
<p>Gel filtration chromatography of Y2 oligosaccharide on a Bio-Gel P-4 Gel column.</p>
Full article ">Figure 2
<p>Negative-ion mode electrospray mass spectrometry (ESI-MS) spectra of YF (<b>a</b>), YD-1 (<b>b</b>), YD-2 (<b>c</b>) and YT (<b>d</b>).</p>
Full article ">Figure 3
<p>Negative-ion mode ESI-MS spectra and HPLC spectra of G1 (<b>d</b>,<b>D-1</b>), G2 (<b>c</b>,<b>C-1</b>), G3 (<b>b</b>,<b>B-1</b>) and G4 (<b>a</b>,<b>A-1</b>).</p>
Full article ">Figure 4
<p>Negative-ion mode electrospray mass spectrometry in tandem with collision-induced dissociation tandem mass spectrometry (ESI-CID-MS/MS) spectrum of the ion at <span class="html-italic">m/z</span> 309.112.</p>
Full article ">Figure 5
<p>Negative-ion mode ESI-CID-MS/MS spectrum of the ion at <span class="html-italic">m/z</span> 339.230.</p>
Full article ">Figure 6
<p>Negative-ion mode ESI-CID-MS/MS spectra of the ions at <span class="html-italic">m/z</span> 545.096 (<b>a</b>) and 531.117 (<b>b</b>).</p>
Full article ">Figure 7
<p>Negative-ion mode ESI-CID-MS/MS spectra of the ions at <span class="html-italic">m/z</span> 1031.250 (<b>a</b>) and 555.100 (−2) (<b>b</b>).</p>
Full article ">

2011

Jump to: 2024, 2022, 2021, 2020, 2019, 2018, 2017, 2016, 2015, 2014, 2013, 2012, 2010

852 KiB  
Review
Marine Polysaccharides in Microencapsulation and Application to Aquaculture: “From Sea to Sea”
by Massimiliano Borgogna, Barbara Bellich and Attilio Cesàro
Mar. Drugs 2011, 9(12), 2572-2604; https://doi.org/10.3390/md9122572 - 8 Dec 2011
Cited by 38 | Viewed by 12179
Abstract
This review’s main objective is to discuss some physico-chemical features of polysaccharides as intrinsic determinants for the supramolecular structures that can efficiently provide encapsulation of drugs and other biological entities. Thus, the general characteristics of some basic polysaccharides are outlined in terms of [...] Read more.
This review’s main objective is to discuss some physico-chemical features of polysaccharides as intrinsic determinants for the supramolecular structures that can efficiently provide encapsulation of drugs and other biological entities. Thus, the general characteristics of some basic polysaccharides are outlined in terms of their conformational, dynamic and thermodynamic properties. The analysis of some polysaccharide gelling properties is also provided, including the peculiarity of the charged polysaccharides. Then, the way the basic physical chemistry of polymer self-assembly is made in practice through the laboratory methods is highlighted. A description of the several literature procedures used to influence molecular interactions into the macroscopic goal of the encapsulation is given with an attempt at classification. Finally, a practical case study of specific interest, the use of marine polysaccharide matrices for encapsulation of vaccines in aquaculture, is reported. Full article
Show Figures

Figure 1

Figure 1
<p>General mechanism (nucleation and growth) of gel formation as macroscopically measured by viscoelastic, optical and chiroptical properties. The insets show sketch of the polymer solution microstructure at the various stages.</p>
Full article ">Figure 2
<p>Schematic representation of the Gibbs free-energy curve <span class="html-italic">vs</span>. temperature for the liquid phase (as a function of composition, blue lines); and for the crystalline phase as a function of decreasing crystal size (red lines). The practical effect is a decrease in the measured melting temperature.</p>
Full article ">Figure 3
<p>Structural representation of four disaccharides from glucose: (<b>a</b>) maltose (α-(1-4)-linked); (<b>b</b>) nigerose (α-(1-3)-linked); (<b>c</b>) cellobiose (β-(1-4)-linked); (<b>d</b>) laminarabiose (β-(1-3)-linked). The related polysaccharide structures of amylose (<b>A</b>), nigeran (<b>B</b>), cellulose (<b>C</b>) and laminaran (<b>D</b>) are shown in the ordered helical conformations as measured by X-ray fiber diffraction studies (left) and as snapshots of random chains modeled by Monte Carlo calculation (right).</p>
Full article ">Figure 4
<p>Correlation between structural conformational data obtained by scattering experiments (top) and structural data obtained by molecular thermodynamic theories (bottom). The example refers to the chain mass-per-unit length (from light scattering) and to the average distance between charged groups (from potentiometry or calorimetry).</p>
Full article ">Figure 5
<p>Schematic representation of the methods described for gel particle technology.</p>
Full article ">Figure 6
<p>Graphical classification of the encapsulation procedures and techniques in polysaccharide hydrogels.</p>
Full article ">Figure 7
<p>(<b>a</b>) Comparison of lysozyme release from alginate beads in phosphate buffer (●) and in Tris/NaCl (■) at pH = 7.4; (<b>b</b>) Comparison of lysozyme release in Tris/NaCl from alginate beads (■) and from alginate/chitosan beads (▲); (<b>c</b>) Experimental integral heat of water evaporation from alginate beads (curve a); and from alginate/chitosan beads (curve b). Polymer samples: alginate (F<sub>G</sub> = 0.4; Mv = 86 kDa); chitosan (Mv = 492 kDa; DA = 11%). Other experimental details in refs [<a href="#B182-marinedrugs-09-02572" class="html-bibr">182</a>,<a href="#B183-marinedrugs-09-02572" class="html-bibr">183</a>].</p>
Full article ">
499 KiB  
Article
Sulfated-Polysaccharide Fraction from Red Algae Gracilaria caudata Protects Mice Gut Against Ethanol-Induced Damage
by Renan Oliveira Silva, Geice Maria Pereira dos Santos, Lucas Antonio Duarte Nicolau, Larisse Tavares Lucetti, Ana Paula Macedo Santana, Luciano de Souza Chaves, Francisco Clark Nogueira Barros, Ana Lúcia Ponte Freitas, Marcellus Henrique Loiola Ponte Souza and Jand-Venes Rolim Medeiros
Mar. Drugs 2011, 9(11), 2188-2200; https://doi.org/10.3390/md9112188 - 2 Nov 2011
Cited by 48 | Viewed by 9773
Abstract
The aim of the present study was to investigate the gastroprotective activity of a sulfated-polysaccharide (PLS) fraction extracted from the marine red algae Gracilaria caudata and the mechanism underlying the gastroprotective activity. Male Swiss mice were treated with PLS (3, 10, 30 and [...] Read more.
The aim of the present study was to investigate the gastroprotective activity of a sulfated-polysaccharide (PLS) fraction extracted from the marine red algae Gracilaria caudata and the mechanism underlying the gastroprotective activity. Male Swiss mice were treated with PLS (3, 10, 30 and 90 mg·kg−1, p.o.), and after 30 min, they were administered 50% ethanol (0.5 mL/25 g−1, p.o.). One hour later, gastric damage was measured using a planimeter. Samples of the stomach tissue were also obtained for histopathological assessment and for assays of glutathione (GSH) and malondialdehyde (MDA). Other groups were pretreated with l-NAME (10 mg·kg−1, i.p.), dl-propargylglycine (PAG, 50 mg·kg−1, p.o.) or glibenclamide (5 mg·kg−1, i.p.). After 1 h, PLS (30 mg·kg−1, p.o.) was administered. After 30 min, ethanol 50% was administered (0.5 mL/25g−1, p.o.), followed by sacrifice after 60 min. PLS prevented-ethanol-induced macroscopic and microscopic gastric injury in a dose-dependent manner. However, treatment with l-NAME or glibenclamide reversed this gastroprotective effect. Administration of propargylglycine did not influence the effect of PLS. Our results suggest that PLS has a protective effect against ethanol-induced gastric damage in mice via activation of the NO/KATP pathway. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">
<p>The effect of PLS on ethanol-induced gastric damage. Mice were treated by gavage with either saline or PLS (3, 10, 30 and 90·mg·kg<sup>−1</sup>). Thirty minutes later, mice in experimental groups were administered 50% ethanol (0.5 mL/25 g<sup>−1</sup>); the negative control group was administered saline. The total area of macroscopic gastric lesions was determined after 1 h. The results are expressed as mean ± SEM of a minimum of 5 animals per group. <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05 <span class="html-italic">vs.</span> saline group; * <span class="html-italic">p</span> &lt; 0.05 <span class="html-italic">vs.</span> ethanol group; ANOVA and Newman-Keuls test.</p>
Full article ">
<p>Photomicrographs of gastric mucosa (Magnification, 100×): (<b>A</b>) saline control; (<b>B</b>) animals treated with 50% ethanol, showing disruption of the superficial region of the gastric gland with epithelial cell loss and intense hemorrhage; (<b>C</b>) animals treated with 50% ethanol + polysaccharide (30 mg·kg<sup>−1</sup>), showing preservation of the gastric mucosa. Quantitative results from these assessments are shown in <a href="#t1-marinedrugs-09-02188" class="html-table">Table 1</a>.</p>
Full article ">
<p>The effect of PLS on glutathione (GSH) levels in the gastric mucosa of mice treated with ethanol. Mice were treated by gavage with saline or PLS (30 mg·kg<sup>−1</sup>). Thirty minutes later, 50% ethanol (0.5 mL 25 g<sup>−1</sup>) was administered to the experimental groups, while the control group was administered saline. Ethanol administration promoted a reduction in the GSH gastric levels. This effect was partially reverted when the animals were treated with PLS. The results are expressed as mean ± SEM of 5 animals per group. <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05 <span class="html-italic">vs.</span> saline group; * <span class="html-italic">p</span> &lt; 0.05 <span class="html-italic">vs.</span> ethanol group; ANOVA and Newman-Keuls test.</p>
Full article ">
<p>The effect of PLS on malondialdehyde (MDA) concentration in the gastric mucosa of mice treated with ethanol. Mice were treated by gavage with PLS (30 mg·kg<sup>−1</sup>). Thirty minutes later, mice in the experimental group were administered 50% ethanol (0.5 mL 25 g<sup>−1</sup>), and the control group was administered saline. The ethanol evidently promoted an increase in MDA gastric levels. When the animals were pre-treated with PLS, this effect was reverted. The results are expressed as the Means ± SEM of 5 animals per group. <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05 <span class="html-italic">vs.</span> saline group; * <span class="html-italic">p</span> &lt; 0.05 <span class="html-italic">vs.</span> ethanol group; ANOVA and Newman-Keuls test.</p>
Full article ">
<p>The effect of <span class="html-small-caps">l</span>-NAME, PAG, and glibenclamide on PLS-mediated protection against macroscopically visible ethanol-induced gastric damage. Mice were initially treated with <span class="html-small-caps">l</span>-NAME (10 mg·kg<sup>−1</sup>, <span class="html-italic">i.p.</span>), <span class="html-small-caps">dl</span>-propargylglycine (PAG, 50 mg·kg<sup>−1</sup>, <span class="html-italic">p.o.</span>), or glibenclamide (5 mg·kg<sup>−1</sup>, <span class="html-italic">i.p.</span>). After 1 h, PLS (30 mg·kg<sup>−1</sup>, <span class="html-italic">p.o.</span>) was administered. Thirty minutes later, 50% ethanol was administered to the experimental groups, and the control group was administered saline. The total area of the macroscopic gastric lesions was determined after 1 h. The results are expressed as mean ± SEM of a minimum of 5 animals per group. * <span class="html-italic">p</span> &lt; 0.05 <span class="html-italic">vs.</span> ethanol group; <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05 <span class="html-italic">vs.</span> PLS + ethanol group; ANOVA and Newman-Keuls test.</p>
Full article ">
1943 KiB  
Review
The Structural Diversity of Carbohydrate Antigens of Selected Gram-Negative Marine Bacteria
by Evgeny L. Nazarenko, Russell J. Crawford and Elena P. Ivanova
Mar. Drugs 2011, 9(10), 1914-1954; https://doi.org/10.3390/md9101914 - 14 Oct 2011
Cited by 39 | Viewed by 10677
Abstract
Marine microorganisms have evolved for millions of years to survive in the environments characterized by one or more extreme physical or chemical parameters, e.g., high pressure, low temperature or high salinity. Marine bacteria have the ability to produce a range of biologically active [...] Read more.
Marine microorganisms have evolved for millions of years to survive in the environments characterized by one or more extreme physical or chemical parameters, e.g., high pressure, low temperature or high salinity. Marine bacteria have the ability to produce a range of biologically active molecules, such as antibiotics, toxins and antitoxins, antitumor and antimicrobial agents, and as a result, they have been a topic of research interest for many years. Among these biologically active molecules, the carbohydrate antigens, lipopolysaccharides (LPSs, O-antigens) found in cell walls of Gram-negative marine bacteria, show great potential as candidates in the development of drugs to prevent septic shock due to their low virulence. The structural diversity of LPSs is thought to be a reflection of the ability for these bacteria to adapt to an array of habitats, protecting the cell from being compromised by exposure to harsh environmental stress factors. Over the last few years, the variety of structures of core oligosaccharides and O-specific polysaccharides from LPSs of marine microrganisms has been discovered. In this review, we discuss the most recently encountered structures that have been identified from bacteria belonging to the genera Aeromonas, Alteromonas, Idiomarina, Microbulbifer, Pseudoalteromonas, Plesiomonas and Shewanella of the Gammaproteobacteria phylum; Sulfitobacter and Loktanella of the Alphaproteobactera phylum and to the genera Arenibacter, Cellulophaga, Chryseobacterium, Flavobacterium, Flexibacter of the Cytophaga-Flavobacterium-Bacteroides phylum. Particular attention is paid to the particular chemical features of the LPSs, such as the monosaccharide type, non-sugar substituents and phosphate groups, together with some of the typifying traits of LPSs obtained from marine bacteria. A possible correlation is then made between such features and the environmental adaptations undertaken by marine bacteria. Full article
Show Figures


Full article ">
Full article ">
Full article ">
Full article ">
Full article ">
Full article ">
Full article ">
Full article ">
Full article ">
<p><span class="html-italic">Pseudoalteromonas rubra</span> ATCC 29570<sup>T</sup> [<a href="#b48-marinedrugs-09-01914" class="html-bibr">48</a>]. Reprinted with permission from Elsevier. M—malic acid residue.</p>
Full article ">
<p>LPS structure of <span class="html-italic">Flexibacter maritimus</span> [<a href="#b153-marinedrugs-09-01914" class="html-bibr">153</a>]. Reproduced with permission from Wiley-Blackwell.</p>
Full article ">
<p>Carbohydrate structure of <span class="html-italic">Loktanella rosea</span> [<a href="#b172-marinedrugs-09-01914" class="html-bibr">172</a>]. Reproduced with permission from Oxford Univeristy Press.</p>
Full article ">
266 KiB  
Review
Therapies from Fucoidan; Multifunctional Marine Polymers
by Janet Helen Fitton
Mar. Drugs 2011, 9(10), 1731-1760; https://doi.org/10.3390/md9101731 - 30 Sep 2011
Cited by 296 | Viewed by 20777
Abstract
Published research on fucoidans increased three fold between 2000 and 2010. These algal derived marine carbohydrate polymers present numerous valuable bioactivities. This review discusses the role for fucoidan in the control of acute and chronic inflammation via selectin blockade, enzyme inhibition and inhibiting [...] Read more.
Published research on fucoidans increased three fold between 2000 and 2010. These algal derived marine carbohydrate polymers present numerous valuable bioactivities. This review discusses the role for fucoidan in the control of acute and chronic inflammation via selectin blockade, enzyme inhibition and inhibiting the complement cascade. The recent data on toxicology and uptake of fucoidan is detailed together with a discussion on the comparative activities of fractions of fucoidan from different sources. Recent in vivo, in vitro and clinical research related to diverse clinical needs is discussed. Targets include osteoarthritis, kidney and liver disease, neglected infectious diseases, hemopoietic stem cell modulation, protection from radiation damage and treatments for snake envenomation. In recent years, the production of well characterized reproducible fucoidan fractions on a commercial scale has become possible making therapies from fucoidan a realizable goal. Full article
6037 KiB  
Review
Marine Polysaccharides: A Source of Bioactive Molecules for Cell Therapy and Tissue Engineering
by Karim Senni, Jessica Pereira, Farida Gueniche, Christine Delbarre-Ladrat, Corinne Sinquin, Jacqueline Ratiskol, Gaston Godeau, Anne-Marie Fischer, Dominique Helley and Sylvia Colliec-Jouault
Mar. Drugs 2011, 9(9), 1664-1681; https://doi.org/10.3390/md9091664 - 23 Sep 2011
Cited by 235 | Viewed by 17068
Abstract
The therapeutic potential of natural bioactive compounds such as polysaccharides, especially glycosaminoglycans, is now well documented, and this activity combined with natural biodiversity will allow the development of a new generation of therapeutics. Advances in our understanding of the biosynthesis, structure and function [...] Read more.
The therapeutic potential of natural bioactive compounds such as polysaccharides, especially glycosaminoglycans, is now well documented, and this activity combined with natural biodiversity will allow the development of a new generation of therapeutics. Advances in our understanding of the biosynthesis, structure and function of complex glycans from mammalian origin have shown the crucial role of this class of molecules to modulate disease processes and the importance of a deeper knowledge of structure-activity relationships. Marine environment offers a tremendous biodiversity and original polysaccharides have been discovered presenting a great chemical diversity that is largely species specific. The study of the biological properties of the polysaccharides from marine eukaryotes and marine prokaryotes revealed that the polysaccharides from the marine environment could provide a valid alternative to traditional polysaccharides such as glycosaminoglycans. Marine polysaccharides present a real potential for natural product drug discovery and for the delivery of new marine derived products for therapeutic applications. Full article
Show Figures


<p>Cell surface GAGs and cell behavior. (<b>a</b>) GAGs and cell adhesion; (<b>b</b>) GAGs and growth factor promotion.</p>
Full article ">
<p>Cell surface GAGs and cell behavior. (<b>a</b>) GAGs and cell adhesion; (<b>b</b>) GAGs and growth factor promotion.</p>
Full article ">
<p>Structure of sulfated oligofucoidan constitutive of algal fucoidan from <span class="html-italic">Ascophyllum nodosum</span> [<a href="#b32-marinedrugs-09-01664" class="html-bibr">32</a>].</p>
Full article ">
<p>Angiographies of hind limbs from rabbits, 3 days after apoptosis induction. (<b>a</b>) Rabbit receiving LMWF; (<b>b</b>) Rabbit receiving placebo.</p>
Full article ">
<p>Repeating unit of the marine bacterial polysaccharide (HE800 EPS) produced by <span class="html-italic">Vibrio diabolicus</span> [<a href="#b68-marinedrugs-09-01664" class="html-bibr">68</a>].</p>
Full article ">
<p>Effect of the DRS HE800 derivative on vascular tube formation on Matrigel from endothelial progenitor cells (EPCs). Photographs show vascular tube formation by EPCs previously treated (<b>a</b>) with 5% of fetal calf serum (control); (<b>b</b>) with proangiogenic factor VEGF (40 ng/mL); and (<b>c</b>) with proangiogenic factor VEGF (40 ng/mL) and DRS HE800 derivative (10 μg/mL).</p>
Full article ">
<p>Repeating unit of marine bacterial polysaccharide (GY785 EPS) produced by <span class="html-italic">Alteromonas infernus</span> [<a href="#b74-marinedrugs-09-01664" class="html-bibr">74</a>,<a href="#b75-marinedrugs-09-01664" class="html-bibr">75</a>].</p>
Full article ">
644 KiB  
Review
Biomedical Exploitation of Chitin and Chitosan via Mechano-Chemical Disassembly, Electrospinning, Dissolution in Imidazolium Ionic Liquids, and Supercritical Drying
by Riccardo A. A. Muzzarelli
Mar. Drugs 2011, 9(9), 1510-1533; https://doi.org/10.3390/md9091510 - 9 Sep 2011
Cited by 179 | Viewed by 15235
Abstract
Recently developed technology permits to optimize simultaneously surface area, porosity, density, rigidity and surface morphology of chitin-derived materials of biomedical interest. Safe and ecofriendly disassembly of chitin has superseded the dangerous acid hydrolysis and provides higher yields and scaling-up possibilities: the chitosan nanofibrils [...] Read more.
Recently developed technology permits to optimize simultaneously surface area, porosity, density, rigidity and surface morphology of chitin-derived materials of biomedical interest. Safe and ecofriendly disassembly of chitin has superseded the dangerous acid hydrolysis and provides higher yields and scaling-up possibilities: the chitosan nanofibrils are finding applications in reinforced bone scaffolds and composite dressings for dermal wounds. Electrospun chitosan nanofibers, in the form of biocompatible thin mats and non-wovens, are being actively studied: composites of gelatin + chitosan + polyurethane have been proposed for cardiac valves and for nerve conduits; fibers are also manufactured from electrospun particles that self-assemble during subsequent freeze-drying. Ionic liquids (salts of alkylated imidazolium) are suitable as non-aqueous solvents that permit desirable reactions to occur for drug delivery purposes. Gel drying with supercritical CO2 leads to structures most similar to the extracellular matrix, even when the chitosan is crosslinked, or in combination with metal oxides of interest in orthopedics. Full article
Show Figures


<p>FTIR spectrum of spray-dried α-chitin nanofibrils ready for incorporation in a chitin + chitosan composite used for wound dressing. This spectrum showed for the first time unmatched resolution of all typical chitin bands. Reprinted from [<a href="#b41-marinedrugs-09-01510" class="html-bibr">41</a>]. Copyright (2007) with permission from Elsevier.</p>
Full article ">
<p>SEM micrograph of chitosan nonwoven fabrics obtained by electrospraying and subsequent freeze drying. Reprinted from [<a href="#b78-marinedrugs-09-01510" class="html-bibr">78</a>]. Copyright (2011) with permission from Elsevier.</p>
Full article ">
<p>The chitosan fibers are replicated by titanium oxide: the SEM image shows the combined chitosan + titania fibers. Reprinted from [<a href="#b107-marinedrugs-09-01510" class="html-bibr">107</a>]. Copyright (2011) with permission from Elsevier.</p>
Full article ">
<p>Removal of chitosan leads to pure titanium oxide with filamentous structure. Reprinted from [<a href="#b107-marinedrugs-09-01510" class="html-bibr">107</a>]. Copyright (2011) with permission from Elsevier.</p>
Full article ">
558 KiB  
Article
Chitosan Nanoparticles Act as an Adjuvant to Promote both Th1 and Th2 Immune Responses Induced by Ovalbumin in Mice
by Zheng-Shun Wen, Ying-Lei Xu, Xiao-Ting Zou and Zi-Rong Xu
Mar. Drugs 2011, 9(6), 1038-1055; https://doi.org/10.3390/md9061038 - 14 Jun 2011
Cited by 176 | Viewed by 14105
Abstract
The study was conducted to investigate the promoted immune response to ovalbumin in mice by chitosan nanoparticles (CNP) and its toxicity. CNP did not cause any mortality or side effects when mice were administered subcutaneously twice with a dose of 1.5 mg at [...] Read more.
The study was conducted to investigate the promoted immune response to ovalbumin in mice by chitosan nanoparticles (CNP) and its toxicity. CNP did not cause any mortality or side effects when mice were administered subcutaneously twice with a dose of 1.5 mg at 7-day intervals. Institute of Cancer Research (ICR) mice were immunized subcutaneously with 25 µg ovalbumin (OVA) alone or with 25 µg OVA dissolved in saline containing Quil A (10 µg), chitosan (CS) (50 µg) or CNP (12.5, 50 or 200 µg) on days 1 and 15. Two weeks after the secondary immunization, serum OVA-specific antibody titers, splenocyte proliferation, natural killer (NK) cell activity, and production and mRNA expression of cytokines from splenocytes were measured. The serum OVA-specific IgG, IgG1, IgG2a, and IgG2b antibody titers and Con A-, LPS-, and OVA-induced splenocyte proliferation were significantly enhanced by CNP (P < 0.05) as compared with OVA and CS groups. CNP also significantly promoted the production of Th1 (IL-2 and IFN-γ) and Th2 (IL-10) cytokines and up-regulated the mRNA expression of IL-2, IFN-γ and IL-10 cytokines in splenocytes from the immunized mice compared with OVA and CS groups. Besides, CNP remarkably increased the killing activities of NK cells activity (P < 0.05). The results suggested that CNP had a strong potential to increase both cellular and humoral immune responses and elicited a balanced Th1/Th2 response, and that CNP may be a safe and efficacious adjuvant candidate suitable for a wide spectrum of prophylactic and therapeutic vaccines. Full article
Show Figures


<p>Morphology of chitosan nanoparticles (CNP). (<b>A</b>, <b>B</b>) atomic force micrographs (AFMs) of CNP; (<b>C</b>) the size distribution by intendity of CNP, the size of CNP ranges from 63.16 to 101.70 nm, and the mean of size is about 83.66 nm; (<b>D</b>) Zeta potential distribution of CNP, CNP exhibit a zeta potential range from 20.04 to 51.13 mV and have a mean charge with 35.43 mV.</p>
Full article ">
<p>Effect of CNP on OVA-specific IgG, IgG1, IgG2a and IgG2b antibody titers in OVA-immunized mice. Mice (<span class="html-italic">n</span> = 6/group) were subcutaneously injected with OVA (25 μg) alone or with OVA 25 (μg) dissolved in saline containing CNP (12.5, 50 or 200 μg), CS (50 μg) or QuilA (10 μg) on days 1 and 15. Sera were collected 2 weeks after the secondary immunizations for analysis these OVA-specific antibodies using indirect ELISA. Bar with different letters are statistically different (<span class="html-italic">P</span> &lt; 0.05). Abbreviations: QuilA: mixture of triterpene saponins from the bark of <span class="html-italic">Quillaja saponaria</span> Molina.</p>
Full article ">
<p>Effect of CNP on NK cell activity in mice immunized with OVA. Mice (<span class="html-italic">n</span> = 6/group) were subcutaneously injected with OVA (25 μg) alone or with OVA 25 μg dissolved in saline containing CNP (12.5, 50 or 200 μg), CS (50 μg) or QuilA (10 μg) on days 1 and 15. Bars with different letters are statistically different (<span class="html-italic">P</span> &lt; 0.05).</p>
Full article ">
<p>Effect of CNP on mitogen- and OVA-stimulated splenocyte proliferation in the mice immunized with OVA. Bars with different letters are statistically different (<span class="html-italic">P</span> &lt; 0.05).</p>
Full article ">
<p>Effect of CNP on mitogen- and OVA-stimulated splenocyte proliferation in the mice immunized with OVA. Bars with different letters are statistically different (<span class="html-italic">P</span> &lt; 0.05).</p>
Full article ">
<p>Effects of CNP on cytokine production in splenocytes from the OVA-immunized mice. Splenocytes were prepared and cultured with Con A for 48 h. The levels of IL-2, IFN-γ and IL-10 in the culture supernatants were determined by ELISA as described in the text. Values are expressed as means ± S.D. of six animals. Bars with different letters are statistically different (<span class="html-italic">P</span> &lt; 0.05).</p>
Full article ">
<p>Eeffect of CNP on the mRNA expression of cytokines and GAPDH in splenocytes from the OVA-immunized mice. Lane M, DNA marker; lane 1, OVA; lane 2, QuilA; lane 3, OVA-CNP (12.5 μg); lane 4, OVA-CNP (50 μg); lane 5, OVA-CNP (200 μg); lane 6, OVA-CS (50 μg).</p>
Full article ">

2010

Jump to: 2024, 2022, 2021, 2020, 2019, 2018, 2017, 2016, 2015, 2014, 2013, 2012, 2011

1311 KiB  
Article
Nature and Lability of Northern Adriatic Macroaggregates
by Jadran Faganeli, Bojana Mohar, Romina Kofol, Vesna Pavlica, Tjaša Marinšek, Ajda Rozman, Nives Kovač and Angela Šurca Vuk
Mar. Drugs 2010, 8(9), 2480-2492; https://doi.org/10.3390/md8092480 - 6 Sep 2010
Cited by 8 | Viewed by 9118
Abstract
The key organic constituents of marine macroaggregates (macrogels) of prevalently phytoplankton origin, periodically occurring in the northern Adriatic Sea, are proteins, lipids and especially polysaccharides. In this article, the reactivity of various macroaggregate fractions in relation to their composition in order to decode [...] Read more.
The key organic constituents of marine macroaggregates (macrogels) of prevalently phytoplankton origin, periodically occurring in the northern Adriatic Sea, are proteins, lipids and especially polysaccharides. In this article, the reactivity of various macroaggregate fractions in relation to their composition in order to decode the potentially »bioavailable« fractions is summarized and discussed. The enzymatic hydrolysis of the macroaggregate matrix, using α-amylase, β-glucosidase, protease, proteinase and lipase, revealed the simultaneous degradation of polysaccharides and proteins, while lipids seem largely preserved. In the fresh surface macroaggregate samples, a pronounced degradation of the α-glycosidic bond compared to β-linkages. Degradation of the colloidal fraction proceeded faster in the higher molecular weight (MW) fractions. N-containing polysaccharides can be important constituents of the higher MW fraction while the lower MW constituents can mostly be composed of poly- and oligosaccharides. Since the polysaccharide component in the higher MW fraction is more degradable compared to N‑containing polysaccharides, the higher MW fraction represents a possible path of organic nitrogen preservation. Enzymatic hydrolysis, using α-amylase and β-glucosidase, revealed the presence of α- and β-glycosidic linkages in all fractions with similar decomposition kinetics. Our results indicate that different fractions of macroaggregates are subjected to compositional selective reactivity with important implications for macroaggregate persistence in the seawater column and deposition. Full article
Show Figures


<p>FT-IR spectra of (<b>A</b>) macroaggregate matrix, surface sample (black bold line) and water column sample (black dotted line); and (<b>B</b>) macroaggregate interstitial water colloids, surface sample (blue bold line) and water column sample (blue dotted line).</p>
Full article ">
<p>C<sub>org.</sub>/N<sub>tot.</sub> (<b>A</b>) and C<sub>org.</sub>/P<sub>tot.</sub> (<b>B</b>) molar ratios ultrafiltrate retentates (UF/0) and permeates (UF/F) using a nominal molecular weight cutoff of 30–10 (UF1), 10–5 (UF2) and &lt;5 kDa (UF3) at the start, after 1 week and after 4 weeks of the degradation experiment.</p>
Full article ">
<p>Concentration changes of carbohydrates and proteins during various enzyme hydrolyses (6 hours at 26 °C).</p>
Full article ">
<p>FT-IR spectra of the (<b>A</b>) surface and (<b>B</b>) water column macroaggregate matrices, and after (i) α-amylase + β-glucosidase, (ii) protease + proteinase K hydrolysis: carbohydrate bands (region ~1150–900 cm<sup>−1</sup>), protein bands (region 1654–1635 cm<sup>−1</sup>), lipid bands (region 2950–2850 cm<sup>−1</sup>) and inorganic (mineral) components (region &lt;1000 cm<sup>−1</sup>).</p>
Full article ">
<p>FT-IR spectra of the surface macroaggregate matrix and aqueous phase of experimental slurries after α-amylase + β-glucosidase and protease + proteinase K hydrolysis.</p>
Full article ">
<p>FT-IR spectra of the surface macroaggregate matrix (sampled at the beginning of the degradation experiment; t<sub>1</sub>—black line), and after lipase (after 3 weeks at 26 °C, t<sub>1</sub>—blue line) hydrolysis: carbohydrate bands (region ~1150–900 cm<sup>−1</sup>), protein bands (region 1654–1635 cm<sup>−1</sup>), lipid bands (region 2950–2850 cm<sup>−1</sup>) and inorganic (mineral) components (region &lt;1000 cm<sup>−1</sup>).</p>
Full article ">
481 KiB  
Review
Marine Polysaccharides in Pharmaceutical Applications: An Overview
by Paola Laurienzo
Mar. Drugs 2010, 8(9), 2435-2465; https://doi.org/10.3390/md8092435 - 2 Sep 2010
Cited by 486 | Viewed by 23322
Abstract
The enormous variety of polysaccharides that can be extracted from marine plants and animal organisms or produced by marine bacteria means that the field of marine polysaccharides is constantly evolving. Recent advances in biological techniques allow high levels of polysaccharides of interest to [...] Read more.
The enormous variety of polysaccharides that can be extracted from marine plants and animal organisms or produced by marine bacteria means that the field of marine polysaccharides is constantly evolving. Recent advances in biological techniques allow high levels of polysaccharides of interest to be produced in vitro. Biotechnology is a powerful tool to obtain polysaccharides from a variety of micro-organisms, by controlling the growth conditions in a bioreactor while tailoring the production of biologically active compounds. Following an overview of the current knowledge on marine polysaccharides, with special attention to potential pharmaceutical applications and to more recent progress on the discovering of new polysaccharides with biological appealing characteristics, this review will focus on possible strategies for chemical or physical modification aimed to tailor the final properties of interest. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">
<p>The synthetic strategy for preparing alginic acid sulfated derivatives.</p>
Full article ">
<p>The synthetic strategy for preparing chitin and chitosan sulfated derivatives [<a href="#b50-marinedrugs-08-02435" class="html-bibr">50</a>].</p>
Full article ">
188 KiB  
Article
Characterization of the Exopolysaccharide Produced by Salipiger mucosus A3T, a Halophilic Species Belonging to the Alphaproteobacteria, Isolated on the Spanish Mediterranean Seaboard
by Inmaculada Llamas, Juan Antonio Mata, Richard Tallon, Philippe Bressollier, María C. Urdaci, Emilia Quesada and Victoria Béjar
Mar. Drugs 2010, 8(8), 2240-2251; https://doi.org/10.3390/md8082240 - 30 Jul 2010
Cited by 62 | Viewed by 13288
Abstract
We have studied the exopolysaccharide produced by the type strain of Salipiger mucosus, a species of halophilic, EPS-producing (exopolysaccharide-producing) bacterium belonging to the Alphaproteobacteria. The strain, isolated on the Mediterranean seaboard, produced a polysaccharide, mainly during its exponential growth phase but [...] Read more.
We have studied the exopolysaccharide produced by the type strain of Salipiger mucosus, a species of halophilic, EPS-producing (exopolysaccharide-producing) bacterium belonging to the Alphaproteobacteria. The strain, isolated on the Mediterranean seaboard, produced a polysaccharide, mainly during its exponential growth phase but also to a lesser extent during the stationary phase. Culture parameters influenced bacterial growth and EPS production. Yield was always directly related to the quantity of biomass in the culture. The polymer is a heteropolysaccharide with a molecular mass of 250 kDa and its components are glucose (19.7%, w/w), mannose (34%, w/w), galactose (32.9%, w/w) and fucose (13.4%, w/w). Fucose and fucose-rich oligosaccharides have applications in the fields of medicine and cosmetics. The chemical or enzymatic hydrolysis of fucose-rich polysaccharides offers a new efficient way to process fucose. The exopolysaccharide in question produces a solution of very low viscosity that shows pseudoplastic behavior and emulsifying activity on several hydrophobic substrates. It also has a high capacity for binding cations and incorporating considerable quantities of sulfates, this latter feature being very unusual in bacterial polysaccharides. Full article
Show Figures


<p>Growth profile and EPS production by <span class="html-italic">S. mucosus</span> strain A3<sup>T</sup> in MY medium at 7.5% total salts <span class="html-italic">versus</span> consumption of glucose. 100% of residual glucose corresponds to 10 g/L of glucose.</p>
Full article ">
<p>Transmission electronic microscopy photograph of <span class="html-italic">S. mucosus</span> strain A3<sup>T</sup> stained with the specific stain for polysaccharide ruthenium red. Bar: 1 μm. Arrow indicates the EPS of the strain.</p>
Full article ">
<p>Viscosity of a 0.5% (w/v) solution of EPSs produced by <span class="html-italic">S. mucosus</span> A3<sup>T</sup> in MY medium at 2.5% total salts.</p>
Full article ">
326 KiB  
Review
Prebiotics from Marine Macroalgae for Human and Animal Health Applications
by Laurie O’Sullivan, Brian Murphy, Peter McLoughlin, Patrick Duggan, Peadar G. Lawlor, Helen Hughes and Gillian E. Gardiner
Mar. Drugs 2010, 8(7), 2038-2064; https://doi.org/10.3390/md8072038 - 1 Jul 2010
Cited by 323 | Viewed by 26444
Abstract
The marine environment is an untapped source of bioactive compounds. Specifically, marine macroalgae (seaweeds) are rich in polysaccharides that could potentially be exploited as prebiotic functional ingredients for both human and animal health applications. Prebiotics are non-digestible, selectively fermented compounds that stimulate the [...] Read more.
The marine environment is an untapped source of bioactive compounds. Specifically, marine macroalgae (seaweeds) are rich in polysaccharides that could potentially be exploited as prebiotic functional ingredients for both human and animal health applications. Prebiotics are non-digestible, selectively fermented compounds that stimulate the growth and/or activity of beneficial gut microbiota which, in turn, confer health benefits on the host. This review will introduce the concept and potential applications of prebiotics, followed by an outline of the chemistry of seaweed polysaccharides. Their potential for use as prebiotics for both humans and animals will be highlighted by reviewing data from both in vitro and in vivo studies conducted to date. Full article
Show Figures


<p>Distribution of the dominant, sub-dominant and minor components of human fecal microflora. Major dominant phyla are denoted. *: Other components are at the family or genus level (adapted from reference [<a href="#b16-marinedrugs-08-02038" class="html-bibr">16</a>]).</p>
Full article ">
<p>Mode of action of prebiotics and purported health benefits in humans and animals.</p>
Full article ">
<p>Green seaweed constituents: <b>(a)</b> α-L-rhamnose and <b>(b)</b> glucuronic acid [<a href="#b31-marinedrugs-08-02038" class="html-bibr">31</a>].</p>
Full article ">
<p>Constituent acids of alginic acid, where <b>(a)</b> is β-D-mannuronic acid and <b>(b)</b> is α-L-guluronic acid [<a href="#b43-marinedrugs-08-02038" class="html-bibr">43</a>].</p>
Full article ">
<p>Fucoidan: Branched polysaccharide sulfate ester with L-fucose building blocks as the major component with predominantly α-(1,2) linkages [<a href="#b43-marinedrugs-08-02038" class="html-bibr">43</a>].</p>
Full article ">
<p>Basic chemical units of laminarin, made up of β-(1,3) and β-(1,6) linked glucose.</p>
Full article ">
<p>Agar constituents [<a href="#b31-marinedrugs-08-02038" class="html-bibr">31</a>] (R=H or CH<sub>3</sub>).</p>
Full article ">
<p>Basic structure of kappa-, iota-, and lambda-carrageenan [<a href="#b60-marinedrugs-08-02038" class="html-bibr">60</a>].</p>
Full article ">
350 KiB  
Review
Bacterial Exopolysaccharides from Extreme Marine Habitats: Production, Characterization and Biological Activities
by Annarita Poli, Gianluca Anzelmo and Barbara Nicolaus
Mar. Drugs 2010, 8(6), 1779-1802; https://doi.org/10.3390/md8061779 - 3 Jun 2010
Cited by 312 | Viewed by 22262
Abstract
Many marine bacteria produce exopolysaccharides (EPS) as a strategy for growth, adhering to solid surfaces, and to survive adverse conditions. There is growing interest in isolating new EPS producing bacteria from marine environments, particularly from extreme marine environments such as deep-sea hydrothermal vents [...] Read more.
Many marine bacteria produce exopolysaccharides (EPS) as a strategy for growth, adhering to solid surfaces, and to survive adverse conditions. There is growing interest in isolating new EPS producing bacteria from marine environments, particularly from extreme marine environments such as deep-sea hydrothermal vents characterized by high pressure and temperature and heavy metal presence. Marine EPS-producing microorganisms have been also isolated from several extreme niches such as the cold marine environments typically of Arctic and Antarctic sea ice, characterized by low temperature and low nutrient concentration, and the hypersaline marine environment found in a wide variety of aquatic and terrestrial ecosystems such as salt lakes and salterns. Most of their EPSs are heteropolysaccharides containing three or four different monosaccharides arranged in groups of 10 or less to form the repeating units. These polymers are often linear with an average molecular weight ranging from 1 × 105 to 3 × 105 Da. Some EPS are neutral macromolecules, but the majority of them are polyanionic for the presence of uronic acids or ketal-linked pyruvate or inorganic residues such as phosphate or sulfate. EPSs, forming a layer surrounding the cell, provide an effective protection against high or low temperature and salinity, or against possible predators. By examining their structure and chemical-physical characteristics it is possible to gain insight into their commercial application, and they are employed in several industries. Indeed EPSs produced by microorganisms from extreme habitats show biotechnological promise ranging from pharmaceutical industries, for their immunomodulatory and antiviral effects, bone regeneration and cicatrizing capacity, to food-processing industries for their peculiar gelling and thickening properties. Moreover, some EPSs are employed as biosurfactants and in detoxification mechanisms of petrochemical oil-polluted areas. The aim of this paper is to give an overview of current knowledge on EPSs produced by marine bacteria including symbiotic marine EPS-producing bacteria isolated from some marine annelid worms that live in extreme niches. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">
<p>The repeating unit of EPS secreted by <span class="html-italic">Pseudoalteromonas</span> strain 721 [<a href="#b39-marinedrugs-08-01779" class="html-bibr">39</a>].</p>
Full article ">
<p>Schematic steps involved in the studies of <span class="html-italic">Geobacillus</span> strain 4004 EPS [<a href="#b48-marinedrugs-08-01779" class="html-bibr">48</a>].</p>
Full article ">
133 KiB  
Communication
Anti-Inflammatory Activity of Chitooligosaccharides in Vivo
by João C. Fernandes, Humberto Spindola, Vanessa De Sousa, Alice Santos-Silva, Manuela E. Pintado, Francisco Xavier Malcata and João E. Carvalho
Mar. Drugs 2010, 8(6), 1763-1768; https://doi.org/10.3390/md8061763 - 28 May 2010
Cited by 102 | Viewed by 13455
Abstract
All the reports to date on the anti-inflammatory activity of chitooligosaccharides (COS) are mostly based on in vitro methods. In this work, the anti-inflammatory activity of two COS mixtures is characterized in vivo (using balb/c mice), following the carrageenan-induced paw edema method. This [...] Read more.
All the reports to date on the anti-inflammatory activity of chitooligosaccharides (COS) are mostly based on in vitro methods. In this work, the anti-inflammatory activity of two COS mixtures is characterized in vivo (using balb/c mice), following the carrageenan-induced paw edema method. This is a widely accepted animal model of acute inflammation to evaluate the anti-inflammatory effect of drugs. Our data suggest that COS possess anti-inflammatoryactivity, which is dependent on dose and, at higher doses, also on the molecular weight. A single dose of 500 mg/kg b.w. weight may be suitable to treat acute inflammation cases; however, further studies are needed to ascertain the effect upon longer inflammation periods as well as studies upon the bioavailability of these compounds. Full article
Show Figures


<p>Effect of various doses of both COS, administered orally 60 min prior to injection of carrageenan, on mice paw edema volume (mL), after 3 and 6 h. (Average ± S.E.M.). Legend: (a) statistically different from all other compounds tested (<span class="html-italic">p</span> &lt; 0.05), except b; (b) statistically different from all other compounds tested (<span class="html-italic">p</span> &lt; 0.05), except a; (c) statistically different (<span class="html-italic">p</span> &lt; 0.05) from COS3–500 mg and INN at 3 and 6 h; (d) statistically different from COS3–500 mg and INN at 3 and 6 h; (e) statistically different (<span class="html-italic">p</span> &lt; 0.05) from other COS3 concentrations; (e*) statistically different from COS5–500 mg at 6 h (<span class="html-italic">p</span> &lt; 0.05); (f) statistically different from other COS5 concentrations; (g) statistically different from COS3 and COS5 concentrations, except 500 mg values (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
<p>Effect of 500 mg/kg b.w. of both COS, administered orally 60 min prior to injection of carrageenan, on mice paw edema volume (mL), along the time (Average ± S.E.M.).</p>
Full article ">
Back to TopTop