[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (38)

Search Parameters:
Keywords = deep-sea sponge

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
25 pages, 29385 KiB  
Article
Porifera Associated with Deep-Water Stylasterids (Cnidaria, Hydrozoa): New Species and Records from the Ross Sea (Antarctica)
by Barbara Calcinai, Teo Marrocco, Camilla Roveta, Stefania Puce, Paolo Montagna, Claudio Mazzoli, Simonepietro Canese, Carlo Vultaggio and Marco Bertolino
J. Mar. Sci. Eng. 2024, 12(12), 2317; https://doi.org/10.3390/jmse12122317 - 17 Dec 2024
Viewed by 635
Abstract
Stylasterid corals are known to be fundamental habitat-formers in both deep and shallow waters. Their tridimensional structure enhances habitat complexity by creating refuges for a variety of organisms and by acting as basibionts for many other invertebrates, including sponges. Porifera represent crucial components [...] Read more.
Stylasterid corals are known to be fundamental habitat-formers in both deep and shallow waters. Their tridimensional structure enhances habitat complexity by creating refuges for a variety of organisms and by acting as basibionts for many other invertebrates, including sponges. Porifera represent crucial components of marine benthic assemblages and, in Antarctica, they often dominate benthic communities. Here, we explore the sponge community associated with thanatocoenosis, mostly composed of dead stylasterid skeletons, collected along the Western and Northern edges of the Ross Sea continental shelf. Overall, 37 sponge species were identified from 278 fragments of the stylasterid Inferiolabiata labiata, of which 7 are first records for the Ross Sea, 1 is first record for Antarctic waters and 2 are proposed as new species. Despite the high biodiversity recorded in this and previous studies on Antarctic deep-sea communities, we are still far from capturing the true richness of Antarctic benthic assemblages. Long-term research programs designed to improve the knowledge of the deep-sea fauna inhabiting Antarctic waters are needed to support successful management and conservation plans, especially in this area, considered one of the main marine diversity hotspots worldwide. Full article
(This article belongs to the Section Marine Biology)
Show Figures

Figure 1

Figure 1
<p>Map of the location of the sampling stations in the Ross Sea continental shelf.</p>
Full article ">Figure 2
<p>(<b>A</b>) Bar plots representing the total number of specimens for sponge species with more than 2 samples. (<b>B</b>) Donut charts showing the percentage of sponge species and specimens with an encrusting (Ec) or massive erect (ME) habit, or both (ME/Ec).</p>
Full article ">Figure 3
<p>Some sponge specimens with larger sizes: (<b>A</b>) <span class="html-italic">Acanthascus</span> (<span class="html-italic">Rhabdocalyptus</span>) <span class="html-italic">australis</span> (MNA 16005, GCR-02-223 D), (<b>B</b>) <span class="html-italic">Iophon unicorne</span> (MNA 16007, GRC-08-023 DC), (<b>C</b>) <span class="html-italic">Haliclona</span> (<span class="html-italic">Gellius</span>) <span class="html-italic">rudis</span> (MNA 16006, GRC-08-023 DE).</p>
Full article ">Figure 4
<p>Bar plots showing (<b>A</b>) the total number of specimens in relation to the area covered on the stylasterid <span class="html-italic">Inferiolabiata labiata</span> and (<b>B</b>) the sponge species with the highest percentage cover on <span class="html-italic">I. labiata</span> (expressed as average ± standard deviation).</p>
Full article ">Figure 5
<p>(<b>A</b>) <span class="html-italic">Lissodendoryx</span> (<span class="html-italic">Lissodendoryx</span>) <span class="html-italic">stylosa</span> sp. nov. (holotype MNA 15959, GRC-02-223 O1); (<b>B</b>) <span class="html-italic">L.</span> (<span class="html-italic">L.</span>) <span class="html-italic">styloderma</span> (MNA 13340, GRC-02-223 (1) sp. 3); (<b>C</b>) <span class="html-italic">Crella</span> (<span class="html-italic">Crella</span>) <span class="html-italic">tubifex</span> (MNA 15968, GRC-02-223 BO1); (<b>D</b>) <span class="html-italic">Esperiopsis flagellata</span> sp. nov. (holotype MNA 15962, GRC-02-223 (8) sp. 6); (<b>E</b>) <span class="html-italic">Artemisina plumosa</span> (MNA 15989, GRC-02-223 AN2); (<b>F</b>) <span class="html-italic">Mycale</span> (<span class="html-italic">Anomomycale</span>) cf. <span class="html-italic">titubans</span> (MNA 15994, GRC-02-223 AH1); (<b>G</b>) <span class="html-italic">Tetilla coronida</span> (MNA 15999, GRC-TR17-007 CP1); (<b>H</b>) <span class="html-italic">Poecillastra antarctica</span> comb. nov. (MNA 13302, GRC-02-223 (26) sp. 1). White arrows indicate the position of the sponge specimens on <span class="html-italic">Inferiolabiata labiata</span> fragments.</p>
Full article ">Figure 6
<p>SEM pictures of <span class="html-italic">Lissodendoryx</span> (<span class="html-italic">Lissodendoryx</span>) <span class="html-italic">stylosa</span> sp. nov.: (<b>A</b>) style I; (<b>B</b>) style II; (<b>C</b>) arcuate isochelae; (<b>D</b>) sigmas.</p>
Full article ">Figure 7
<p>SEM pictures of <span class="html-italic">Lissodendoryx</span> (<span class="html-italic">Lissodendoryx</span>) <span class="html-italic">styloderma</span>: (<b>A</b>) subtylostyle; (<b>B</b>) magnification of the head and the pointed tip of a subtylostyle; (<b>C</b>) tornotes; (<b>D</b>) magnification of a pointed end of a tornote; (<b>E</b>) arcuate isochelae.</p>
Full article ">Figure 8
<p>SEM pictures of <span class="html-italic">Crella</span> (<span class="html-italic">Crella</span>) <span class="html-italic">tubifex</span>: (<b>A</b>) acanthostyle; (<b>B</b>) anisostrongyle; (<b>C</b>) acanthostrongyle; (<b>D</b>) magnification of the central portion and extremities of an acanthostrongyle.</p>
Full article ">Figure 9
<p>Optical microscope pictures of <span class="html-italic">Esperiopsis flagellata</span> sp. nov.: (<b>A</b>) style; (<b>B</b>) end of a style; (<b>C</b>) isochelae I; (<b>D</b>) isochelae II; (<b>E</b>) C-shaped sigma; (<b>F</b>) flagellated sigma.</p>
Full article ">Figure 10
<p>SEM pictures of <span class="html-italic">Artemisina plumosa</span>: (<b>A</b>) style I; (<b>B</b>) style II; (<b>C</b>) tylote; (<b>D</b>) magnification of the spined head of a tylote; (<b>E</b>) isochelae; (<b>F</b>) toxas.</p>
Full article ">Figure 11
<p>SEM pictures of <span class="html-italic">Mycale</span> (<span class="html-italic">Anomomycale</span>) cf. <span class="html-italic">titubans</span>: (<b>A</b>) large mycalostyle; (<b>B</b>) thin mycalostyle; (<b>C</b>) anomochelae; (<b>D</b>) sigmas.</p>
Full article ">Figure 12
<p><span class="html-italic">Hymeniacidon fragilis</span> (Koltun, 1964) comb. nov.: (<b>A</b>) massively encrusting specimen MNA 13377 (GRC-02-223 (22) sp. 1); (<b>B</b>) laminar specimen MNA 16004 (GRC-02-223 (37) D); (<b>C</b>,<b>D</b>) choanosomal skeleton; (<b>E</b>) ectosomal skeleton; (<b>F</b>) spicules. White arrows indicate the position of the sponge attachment to the stylasterid.</p>
Full article ">Figure 13
<p>SEM pictures of <span class="html-italic">Tetilla coronida</span>: (<b>A</b>) oxeas; (<b>B</b>) extremity of a protriaene; (<b>C</b>) extremities of anatrienes; (<b>D</b>) extremities of anamonaenes; (<b>E</b>) sigmaspires.</p>
Full article ">Figure 14
<p>Optical microscope pictures of <span class="html-italic">Poecillastra antarctica</span>: (<b>A</b>) oxeas; (<b>B</b>) calthrops and orthotrianes; (<b>C</b>) plesiasters; (<b>D</b>) amphiasters; (<b>E</b>) spirasters.</p>
Full article ">
21 pages, 3060 KiB  
Article
Shedding Light on the Italian Mesophotic Spongofauna
by Margherita Toma, Marzia Bo, Marco Bertolino, Martina Canessa, Michela Angiolillo, Alessandro Cau, Franco Andaloro, Simonepietro Canese, Silvestro Greco and Giorgio Bavestrello
J. Mar. Sci. Eng. 2024, 12(11), 2110; https://doi.org/10.3390/jmse12112110 - 20 Nov 2024
Cited by 1 | Viewed by 642
Abstract
An analysis of 483 remotely operated vehicle (ROV) dives carried out along the Italian coast on hard substrata at mesophotic depths (40–200 m) allowed an overview of the rich sponge diversity (53 taxa) of the deep continental platform to be obtained for the [...] Read more.
An analysis of 483 remotely operated vehicle (ROV) dives carried out along the Italian coast on hard substrata at mesophotic depths (40–200 m) allowed an overview of the rich sponge diversity (53 taxa) of the deep continental platform to be obtained for the first time. About 40% of the potential actual species diversity was recognisable using ROV, suggesting that this group is among the richest yet underestimated using this technology in contrast to other megabenthic taxa. Additionally, the study allowed us to gather data on the current basin-scale distribution and bathymetric limits of five common and easily identifiable demosponges with up to 55% occurrence in the explored sites: Aplysina cavernicola, the group Axinella damicornis/verrucosa, Chondrosia reniformis, Foraminospongia spp., and Hexadella racovitzai. Four of these latitudinal distributions were characterised by high occurrence in the Ligurian Sea and a progressive decrease towards the south Tyrrhenian Sea, with an occasional second minor peak of occurrence in the Sicily Channel. In contrast, Foraminospongia spp. showed a maximum occurrence on the offshore reliefs and a second one in the North–central Tyrrhenian Sea, while it was almost absent in the Ligurian Sea. Trophic and biogeographic reasons were discussed as possible causes of the double-peak distributions. The vertical distributions support a more consistent occurrence of all considered taxa in deeper waters than previously known. This suggests that they may more typically belong to the mesophotic realm than the shallow waters, owing to a more extensive sampling effort in the deeper depth range. The five target taxa are typical or associated species of seven reference habitats in the recently revised UNEP/SPA-RAC classification. However, they may create such dense aggregations that they should be listed as new facies in the abovementioned classification. Full article
Show Figures

Figure 1

Figure 1
<p>Non-metric Multi-Dimensional Scaling (nMDS) plots for the (<b>A</b>) upper (&lt;100 m) and (<b>B</b>) lower (&gt;100 m) mesophotic zone. Bray–Curtis similarity Index; Shepard plot stress plots: 0.015. LIG, Ligurian Sea; NCT, North–central Tyrrhenian Sea; ST, South Tyrrhenian Sea; SC, Sicily Channel.</p>
Full article ">Figure 2
<p>(<b>A</b>) Similarity percentage of each macro-area, and (<b>B</b>) pair-wise comparisons among macro-areas. White bars, upper (&lt;100 m) mesophotic zone; black bars, lower (&gt;100 m) mesophotic zone. LIG, Ligurian Sea; NCT, North–central Tyrrhenian Sea; ST, South Tyrrhenian Sea; SC, Sicily Channel.</p>
Full article ">Figure 3
<p>(<b>A</b>) Average number of sponge taxa (± SE) in each site in the five considered macro-areas, and (<b>B</b>) percentage frequency of occurrence of the five target sponge taxa in each macro-area. Legend: LIG, Ligurian Sea; NCT, North–central Tyrrhenian Sea; ST, South Tyrrhenian Sea, SC, Sicily Channel; OR, offshore reliefs.</p>
Full article ">Figure 4
<p>(<b>A</b>–<b>E</b>) Distribution maps of the populations of five target taxa, indicating the densities (number of individuals m<sup>−2</sup>, percentage of coverage).</p>
Full article ">Figure 5
<p>Bathymetric distributions of the five target taxa observed in the present work (yellow) and reported in the available literature (grey).</p>
Full article ">Figure 6
<p>Aggregations of the five target taxa in different habitats: (<b>A</b>) numerous colonies of <span class="html-italic">Aplysina cavernicola</span> on the coralligenous outcrops in Diano Marina (western Ligurian Sea, 55 m); (<b>B</b>) <span class="html-italic">A. cavernicola</span> and <span class="html-italic">Hexadella racovitzai</span> dominating the basal layer of the facies with Alcyonacea on Ischia Island (South Tyrrhenian Sea, 75 m); (<b>C</b>) dense aggregation of <span class="html-italic">Axinella damicornis/verrucosa</span> on the detritic bottom of Santo Stefano (western Ligurian Sea, 50 m); (<b>D</b>,<b>E</b>) numerous individuals of <span class="html-italic">A. damicornis/verrucosa</span> beneath gorgonian and black coral forests, respectively (Bordighera Canyon, western Ligurian Sea, 75 m; Favazzina, South Tyrrhenian Sea, 85 m); (<b>F</b>) the densest aggregation of <span class="html-italic">Chondrosia reniformis</span> observed in the present study on the circalittoral coralligenous outcrops in the Gulf of Gioiosa (South Tyrrhenian Sea, 78 m); (<b>G</b>) <span class="html-italic">A. cavernicola</span> and <span class="html-italic">C. reniformis</span> together with other benthic organisms completely covering an iron shipwreck in Lampedusa (Sicily Channel, 65 m); (<b>H</b>) <span class="html-italic">Foraminospongia</span> spp. on a rhodolith bed in Porto Corallo (North–central Tyrrhenian Sea, 125 m); (<b>I</b>) dense aggregation of <span class="html-italic">Foraminospongia</span> spp. on coralligenous accretions in the Pontine Archipelago (North–central Tyrrhenian Sea, 105 m); (<b>J</b>) numerous specimens of <span class="html-italic">Foraminospongia</span> spp. together with structuring anthozoans on the offshore circalittoral rocks off Orosei Canyon (North–central Tyrrhenian Sea, 170 m); (<b>K</b>) bright reddish patches of <span class="html-italic">H. racovitzai</span> on a coralligenous cliff of Elba Island (North–central Tyrrhenian Sea, 85 m); (<b>L</b>) the silted deep banks of Santo Stefano covered by contracted <span class="html-italic">H. racovitzai</span> (western Ligurian Sea, 90 m). Scale bar: 10 cm.</p>
Full article ">Figure 7
<p>Percentage preference for the considered types of substrates (<b>left</b>) and slope (<b>right</b>) of the target taxa. Substrate categories’ codes: D, detritic including rhodoliths; R, outcropping or sub-outcropping rocks; CCA, rocks covered by crustose coralline algae. Inclination category codes: H, horizontal or sub-horizontal; S, sloping; V, vertical or sub-vertical.</p>
Full article ">
24 pages, 8726 KiB  
Article
Tricoma (Tricoma) disparseta sp. nov. (Nematoda: Desmoscolecidae), a New Free-Living Marine Nematode from a Seamount in the Northwest Pacific Ocean, with a New Record of T. (T.) longirostris (Southern, 1914)
by Hyo Jin Lee, Heegab Lee, Ji-Hoon Kihm and Hyun Soo Rho
Diversity 2024, 16(10), 648; https://doi.org/10.3390/d16100648 - 20 Oct 2024
Cited by 2 | Viewed by 887
Abstract
During a survey of marine biodiversity in the deep sea off northeastern Guam, two marine desmoscolecid nematodes belonging to the subgenus Tricoma were discovered. Tricoma (Tricoma) disparseta sp. nov. was described based on specimens collected from sponge and starfish habitats on [...] Read more.
During a survey of marine biodiversity in the deep sea off northeastern Guam, two marine desmoscolecid nematodes belonging to the subgenus Tricoma were discovered. Tricoma (Tricoma) disparseta sp. nov. was described based on specimens collected from sponge and starfish habitats on a seamount at depths ranging from 1300 to 1500 m. Tricoma (Tricoma) disparseta sp. nov. is distinguished by having 59 to 62 main rings, 9 to 10 subdorsal setae, and 14 to 18 subventral setae on each side. Notable features include the differentiation in length and insertion between the subdorsal and subventral setae, as well as the amphid extending to the second or third main ring. Additionally, the spicules have a relatively small capitulum at the proximal end, while the gubernaculum is bent into a hooked shape. The specimen of T. (T.) longirostris observed in this study closely resembles previously reported specimens, characterized by 78 main rings, a long and narrow head shape, eight to nine subdorsal setae, 14 to 15 subventral setae, and a gubernaculum with a knobbed apophysis. Two Tricoma species from the Northwest Pacific Ocean are described in detail, and pictorial keys and comparative tables for species identification are provided for groups with 50 to 64 main rings. Full article
(This article belongs to the Special Issue Dynamics of Marine Communities)
Show Figures

Figure 1

Figure 1
<p>A map depicting the locations where samples were collected.</p>
Full article ">Figure 2
<p><span class="html-italic">Tricoma</span> (<span class="html-italic">Tricoma</span>) <span class="html-italic">disparseta</span> sp. nov. Holotype male (MNB001). (<b>A</b>) Entire view of the male body, lateral view; (<b>B</b>) head region, left side; (<b>C</b>) head region, ventral view (Paratype MNB006); (<b>D</b>) head region, right side (Paratype MNB005); (<b>E</b>) spicules and tail region, right side (Paratype MNB005); (<b>F</b>) a posterior region showing sightly laterally inserted subventral setae (arrow). Scale bars: 50 µm in (<b>A</b>); 10 µm in (<b>B</b>–<b>E</b>); 20 µm in (<b>F</b>).</p>
Full article ">Figure 3
<p><span class="html-italic">Tricoma</span> (<span class="html-italic">Tricoma</span>) <span class="html-italic">disparseta</span> sp. nov. Allotype female (MNB007). (<b>A</b>) Entire view of the male body, lateral view; (<b>B</b>) head region, right side; (<b>C</b>) head region, right side (Paratype MNB008); (<b>D</b>) reproductive systems; (<b>E</b>) tail region, right side. Scale bars: 50 µm in (<b>A</b>); 10 µm in (<b>B</b>,<b>C</b>); 20 µm in (<b>D</b>,<b>E</b>).</p>
Full article ">Figure 4
<p><span class="html-italic">Tricoma</span> (<span class="html-italic">Tricoma</span>) <span class="html-italic">disparseta</span> sp. nov. SEM photomicrographs, male. (<b>A</b>) Entire view of the body, lateral view; (<b>B</b>) anterior region showing laterally inserted subventral setae (white arrow); (<b>C</b>) head showing the amphideal fovea, with a white arrow indicating the amphideal pore, dorsal view; (<b>D</b>) head region, anterior view; (<b>E</b>) cephalic setae enclosed by a thin membrane; (<b>F</b>) the distal end of cephalic setae, split and enclosed by a thin membrane. Scale bars: 100 µm in (<b>A</b>); 10 µm in (<b>B</b>); 5 µm in (<b>C</b>–<b>E</b>); 500 nm in (<b>F</b>).</p>
Full article ">Figure 5
<p><span class="html-italic">Tricoma</span> (<span class="html-italic">Tricoma</span>) <span class="html-italic">disparseta</span> sp. nov. SEM photomicrographs, male. (<b>A</b>) Cuticular layer showing the height difference between the dorsal and ventral peduncles; (<b>B</b>) subdorsal setae; (<b>C</b>) subventral setae appearing with curved tip; (<b>D</b>) spicules region, lateral view; (<b>E</b>) spicules region, anterior view; (<b>F</b>) terminal ring. Scale bars: 10 µm in (<b>A</b>,<b>C</b>,<b>D</b>); 5 µm in (<b>B</b>,<b>E</b>,<b>F</b>).</p>
Full article ">Figure 6
<p><span class="html-italic">Tricoma</span> (<span class="html-italic">Tricoma</span>) <span class="html-italic">disparseta</span> sp. nov., DIC photomicrographs, holotype male (MNB001). (<b>A</b>) Entire view of the body; (<b>B</b>) anterior region; (<b>C</b>) amphideal fovea; (<b>D</b>) posterior region; (<b>E</b>) head region, ventral view (paratype MNB006); (<b>F</b>) somatic setae; (<b>G</b>) spicules and gubernaculum of the specimen treated with lactic acid, which was additionally used to increase transparency for the observation of internal reproductive organs, and showing spicules capitulum (arrow) and hooked gubernaculum (white arrow) (paratype MNB003); (<b>H</b>) terminal ring showing phasmata (paratype MNB004). Scale bars: 50 µm in (<b>A</b>); 10 µm in (<b>B</b>–<b>H</b>).</p>
Full article ">Figure 7
<p><span class="html-italic">Tricoma</span> (<span class="html-italic">Tricoma</span>) <span class="html-italic">disparseta</span> sp. nov., DIC photomicrographs, allotype female (MNB007). (<b>A</b>) Entire view of the body; (<b>B</b>) head region; (<b>C</b>) amphideal fovea showing anterior margin (arrow); (<b>D</b>) reproductive system showing naked vulva (arrow), left side; (<b>E</b>) posterior region; (<b>F</b>) entire view of the body showing somatic setae (paratype MNB008). Scale bars: 50 µm in (<b>A</b>,<b>F</b>); 10 µm in (<b>B</b>–<b>E</b>).</p>
Full article ">Figure 8
<p>Pictorial key to the species group with 50 to 64 main rings in the subgenus <span class="html-italic">Tricoma</span>. Sources of the figures: (<b>A</b>) Timm (1970); (<b>B</b>) Decraemer (1979); (<b>C</b>) Freudenhammer (1975); (<b>D</b>) Decraemer (1987); (<b>E</b>) Decraemer (1987); (<b>F</b>) Decraemer (1987); (<b>G</b>) Timm (1970); (<b>H</b>) Decraemer (1978); (<b>I</b>) Decraemer (1987); (<b>J</b>) <span class="html-italic">T</span>. (<span class="html-italic">T</span>.) <span class="html-italic">disparseta</span> sp. nov.; (<b>K</b>) Timm (1970); (<b>L</b>) Decraemer (1978); (<b>M</b>) Decraemer (1983); (<b>N</b>) Blome (1982); (<b>O</b>) Decraemer (1987); (<b>P</b>) Timm (1970); (<b>Q</b>) Blome (1982); (<b>R</b>) Chitwood (1951); (<b>S</b>) Decraemer (1986); (<b>T</b>) Decraemer (1979); (<b>U</b>) Lee, Lee and Rho (2023).</p>
Full article ">Figure 8 Cont.
<p>Pictorial key to the species group with 50 to 64 main rings in the subgenus <span class="html-italic">Tricoma</span>. Sources of the figures: (<b>A</b>) Timm (1970); (<b>B</b>) Decraemer (1979); (<b>C</b>) Freudenhammer (1975); (<b>D</b>) Decraemer (1987); (<b>E</b>) Decraemer (1987); (<b>F</b>) Decraemer (1987); (<b>G</b>) Timm (1970); (<b>H</b>) Decraemer (1978); (<b>I</b>) Decraemer (1987); (<b>J</b>) <span class="html-italic">T</span>. (<span class="html-italic">T</span>.) <span class="html-italic">disparseta</span> sp. nov.; (<b>K</b>) Timm (1970); (<b>L</b>) Decraemer (1978); (<b>M</b>) Decraemer (1983); (<b>N</b>) Blome (1982); (<b>O</b>) Decraemer (1987); (<b>P</b>) Timm (1970); (<b>Q</b>) Blome (1982); (<b>R</b>) Chitwood (1951); (<b>S</b>) Decraemer (1986); (<b>T</b>) Decraemer (1979); (<b>U</b>) Lee, Lee and Rho (2023).</p>
Full article ">Figure 9
<p><span class="html-italic">Tricoma</span> (<span class="html-italic">Tricoma</span>) <span class="html-italic">longirostris</span> (Southern, 1914), newly discovered in this study, male (<b>A</b>–<b>D</b>). (<b>A</b>) Entire view of the male body, lateral view; (<b>B</b>) head region, lateral view; (<b>C</b>) spicules and gubernaculum; (<b>D</b>) spicule and tail region. <span class="html-italic">T.</span> (<span class="html-italic">T.</span>) <span class="html-italic">longirostris</span> (Southern, 1914) from the original description, male (<b>E</b>–<b>G</b>). (<b>E</b>) head, dorsal view; (<b>F</b>) spicules and gubernaculum; (<b>G</b>) tail region (after Southern, 1914). <span class="html-italic">T.</span> (<span class="html-italic">T.</span>) <span class="html-italic">glutinosa</span> Steiner, 1916, male (<b>H</b>). (<b>H</b>) Anterior end (after Steiner, 1916). <span class="html-italic">T.</span> (<span class="html-italic">T.</span>) <span class="html-italic">septentrionalis</span> Timm, 1978, male (<b>I</b>,<b>J</b>). (<b>I</b>) Head region; (<b>J</b>) specular apparatus (after Timm, 1978). <span class="html-italic">T.</span> (<span class="html-italic">T.</span>) <span class="html-italic">longirostris</span> (Southern, 1914), male (<b>K</b>,<b>L</b>). (<b>K</b>) Head, surface view; (<b>L</b>) copulatory apparatus and tail (after Decraemer, 1983). Scale bars: 50 µm in (<b>A</b>); 30 µm in (<b>L</b>); 15 µm in (<b>K</b>); 12 µm in (<b>I</b>,<b>J</b>); 10 µm in (<b>B</b>–<b>D</b>).</p>
Full article ">Figure 10
<p><span class="html-italic">Tricoma</span> (<span class="html-italic">Tricoma</span>) <span class="html-italic">longirostris</span> (Southern, 1914), DIC photomicrographs, male. (<b>A</b>) Entire body view; (<b>B</b>) head region; (<b>C</b>) cephalic setae; (<b>D</b>) amphideal fovea; (<b>E</b>) spicules and gubernaculum showing a knobbed apophysis (arrow); (<b>F</b>) common forms of somatic setae; (<b>G</b>) tail region. Scale bars: 50 µm in (<b>A</b>); 10 µm in (<b>B</b>–<b>G</b>).</p>
Full article ">
15 pages, 3980 KiB  
Article
Taxonomic Investigations on Cladorhizidae (Carnivorous Sponges) of the East Scotia Ridge (Antarctica) with the Description of Three New Species
by Camino Eck, Xiaoyu Kröner and Dorte Janussen
J. Mar. Sci. Eng. 2024, 12(4), 612; https://doi.org/10.3390/jmse12040612 - 31 Mar 2024
Viewed by 1600
Abstract
This study investigates taxonomic characteristics of carnivorous sponges from the Southern Ocean. The specimens were collected in 2010 from deep-sea hydrothermal vents of the East Scotia Ridge during the RRS James Cook Cruise JC42. All the investigated sponges are new to science. They [...] Read more.
This study investigates taxonomic characteristics of carnivorous sponges from the Southern Ocean. The specimens were collected in 2010 from deep-sea hydrothermal vents of the East Scotia Ridge during the RRS James Cook Cruise JC42. All the investigated sponges are new to science. They belong to the genera Abyssocladia and Cladorhiza within the family Cladorhizidae. This study provides descriptions and remarks for the three new species Abyssocladia truespacemeni, Abyssocladia hendrixii and Cladorhiza elsaae. Comparative faunistic and ecological aspects of these sponge genera within the Southern Ocean sponge fauna are discussed. The genera Abyssocladia and Cladorhiza are recorded here for the first time from the ecosystem around hydrothermal vents in the Antarctic deep sea. The descriptions of new species contribute to and expand the current knowledge of the Cladorhizidae and consequently support future taxonomic identifications and descriptions of Antarctic deep-sea carnivorous sponges. The appearance of these newly discovered species underlines the hypothesis that Cladorhizidae is the second most species-rich family of Demospongiae in the Southern Ocean, and many new species of this deep-sea sponge family are still to be discovered. Full article
Show Figures

Figure 1

Figure 1
<p>Location of the East Scotia Ridge hydrothermal vents in the Southern Ocean and the three sites Kemp Caldera, E2 and E9 (black dots) surveyed during the CHESSO Cruise JC 42 in 2010 [<a href="#B25-jmse-12-00612" class="html-bibr">25</a>]. E2 and E9 are the sites where the ROV <span class="html-italic">Isis</span> collected the studied sponges during the expedition [<a href="#B25-jmse-12-00612" class="html-bibr">25</a>]. The main map originated from Arango and Linse [<a href="#B24-jmse-12-00612" class="html-bibr">24</a>], whose study describes three new species of <span class="html-italic">Sericosura</span>, also collected on the RRS James Cook Cruise JC 42 [<a href="#B24-jmse-12-00612" class="html-bibr">24</a>]. The overview map in the upper corner, whose red frame broadly indicates the research area, was created using OpenStreetMap.</p>
Full article ">Figure 2
<p><span class="html-italic">Abyssocladia truespacemeni</span>. (<b>a</b>) In situ picture of specimen SMF-12213. (<b>b</b>) Holotype with a disciform body, the right picture shows eggs on the underside of the body. (<b>c</b>) Mycalostyle. (<b>d</b>) Oxea. (<b>e</b>) Isochela I. (<b>f</b>) Sigmancistra I. (<b>g</b>) Sigmancistra II. (<b>h</b>) Isochela II, front and back view.</p>
Full article ">Figure 3
<p><span class="html-italic">Abyssocladia hendrixii</span>. (<b>a</b>) Left picture shows paratype SMF-12215, right picture shows holotype SMF-12214. (<b>b</b>) Arcuate isochelae. (<b>c</b>) Mycalostyle. (<b>d</b>) Sigmancistra I. (<b>e</b>) Sigmancistra II.</p>
Full article ">Figure 4
<p><span class="html-italic">Cladorhiza elsaae</span>. (<b>a</b>) The white holotype SMF-12210-a and the reddish paratype 1 SMF-12211. (<b>b</b>) Anchorate anisochelae front and side view. (<b>c</b>) Sigma. (<b>d</b>) Developing anisochela side view. (<b>e</b>) Mycalostyle and enlargement of ends. (<b>f</b>) Sigmancistras I. (<b>g</b>) Sigmancistras II.</p>
Full article ">Figure 5
<p>Bayesian phylogenetic analyses of <span class="html-italic">Cladorhiza</span> by Georgieva et al. [<a href="#B29-jmse-12-00612" class="html-bibr">29</a>] based on a concatenated dataset of COI, 28S and ALG11 [<a href="#B29-jmse-12-00612" class="html-bibr">29</a>] and modified with the new species <span class="html-italic">Cladorhiza elsaae</span>. The yellow marker and the red font indicate the samples taken from the E2 site. The specimens SMF 12210 (holotype) and SMF12211 (paratype) are particularly relevant and were investigated for this study.</p>
Full article ">
19 pages, 4777 KiB  
Article
A Bio-Inspired Approach to Improve the Toughness of Brittle Bast Fibre-Reinforced Composites Using Cellulose Acetate Foils
by Nina Graupner and Jörg Müssig
Biomimetics 2024, 9(3), 131; https://doi.org/10.3390/biomimetics9030131 - 21 Feb 2024
Cited by 1 | Viewed by 1520
Abstract
Bast fibre-reinforced plastics are characterised by good strength and stiffness but are often brittle due to the stiff and less ductile fibres. This study uses a biomimetic approach to improve impact strength. Based on the structure of the spicules of a deep-sea glass [...] Read more.
Bast fibre-reinforced plastics are characterised by good strength and stiffness but are often brittle due to the stiff and less ductile fibres. This study uses a biomimetic approach to improve impact strength. Based on the structure of the spicules of a deep-sea glass sponge, in which hard layers of bioglass alternate with soft layers of proteins, the toughness of kenaf/epoxy composites was significantly improved by a multilayer structure of kenaf and cellulose acetate (CA) foils as impact modifiers. Due to the alternating structure, cracks are deflected, and toughness is improved. One to five CA foils were stacked with kenaf layers and processed to composite plates with bio-based epoxy resin by compression moulding. Results have shown a significant improvement in toughness using CA foils due to increased crack propagation. The unnotched Charpy impact strength increased from 9.0 kJ/m2 of the pure kenaf/epoxy composite to 36.3 kJ/m2 for the sample containing five CA foils. The tensile and flexural strength ranged from 74 to 81 MPa and 112 to 125 MPa, respectively. The tensile modulus reached values between 9100 and 10,600 MPa, and the flexural modulus ranged between 7200 and 8100 MPa. The results demonstrate the successful implementation of an abstract transfer of biological role models to improve the toughness of brittle bast fibre-reinforced plastics. Full article
(This article belongs to the Special Issue Advances in Biomaterials, Biocomposites and Biopolymers)
Show Figures

Figure 1

Figure 1
<p>Production of kenaf fibre-reinforced composites with CA foils as an impact modifier. Optionally, the possibility for further processing the multilayer webs into needle felts as semi-finished products is shown in light grey.</p>
Full article ">Figure 2
<p>Tensile strength (<b>A</b>), flexural strength (<b>B</b>), Young’s modulus (<b>C</b>) and flexural modulus (<b>D</b>) of the kenaf composite and hybrid materials with different numbers of CA foil layers. The bars represent the median values, and the error bars represent the mean arithmetic deviation (MAD). An asterisk * indicates results that are not normally distributed; different letters represent significant differences.</p>
Full article ">Figure 3
<p>Stress-strain curves from the tensile tests (<b>A</b>) and the bending tests (<b>B</b>) of different materials.</p>
Full article ">Figure 4
<p>Unnotched Charpy impact strength of the kenaf composite and hybrid materials with different numbers of CA foil layers. The bars represent the median values, and the error bars represent the mean arithmetic deviation (MAD). All results are normally distributed; other letters represent significant differences.</p>
Full article ">Figure 5
<p>SEM micrographs of the fracture surfaces from the tensile test of a kenaf-reinforced composite. (<b>A</b>) overview, (<b>B</b>,<b>C</b>) detailed images.</p>
Full article ">Figure 6
<p>SEM images of the fracture surfaces from the tensile test of a kenaf/CA foil hybrid composite with four layers of CA foils. (<b>A</b>,<b>B</b>) overview, (<b>C</b>,<b>D</b>) detail images of the connection between the punched CA foils and the kenaf multilayer webs.</p>
Full article ">Figure 7
<p>SEM micrographs of the fracture surface of a deep-sea glass sponge (Hexactinellida) anchor spicule, (<b>A</b>) overview, (<b>B</b>) detailed micrograph from the marked region in (<b>A</b>).</p>
Full article ">Figure 8
<p>Unnotched Charpy impact strength (mean ± standard deviation) for different bast fibre-reinforced composites and hybrid materials. The dotted line represents the values of the pure matrix. (<b>A</b>) kenaf, lyocell and hybrid materials made of kenaf and lyocell produced in compression moulding and injection moulding process with a PLA matrix, (<b>B</b>) unidirectionally reinforced flax, PLA, aramid and hybrid materials with a bio-based epoxy matrix (Greenpoxy) produced using pultrusion process and (<b>C</b>) flax, carbon and hybrid materials made of fabrics in an epoxy matrix produced using compression moulding process.</p>
Full article ">
43 pages, 7157 KiB  
Review
Deep-Sea Sponges and Corals off the Western Coast of Florida—Intracellular Mechanisms of Action of Bioactive Compounds and Technological Advances Supporting the Drug Discovery Pipeline
by Mina Iskandar, Kira M. Ruiz-Houston, Steven D. Bracco, Sami R. Sharkasi, Cecilia L. Calabi Villarroel, Meghna N. Desai, Alexandra G. Gerges, Natalia A. Ortiz Lopez, Miguel Xiao Barbero, Amelia A. German, Vinoothna S. Moluguri, Selina M. Walker, Juliana Silva Higashi, Justin M. Palma, Daena Z. Medina, Miit Patel, Prachi Patel, Michaela Valentin, Angelica C. Diaz, Jonathan P. Karthaka, Atzin D. Santiago, Riley B. Skiles, Luis A. Romero Umana, Maxwell D. Ungrey, Anya Wojtkowiak, Domenica V. Howard, Remy Nurge, Katharine G. Woods and Meera Nanjundanadd Show full author list remove Hide full author list
Mar. Drugs 2023, 21(12), 615; https://doi.org/10.3390/md21120615 - 28 Nov 2023
Viewed by 6299
Abstract
The majority of natural products utilized to treat a diverse array of human conditions and diseases are derived from terrestrial sources. In recent years, marine ecosystems have proven to be a valuable resource of diverse natural products that are generated to defend and [...] Read more.
The majority of natural products utilized to treat a diverse array of human conditions and diseases are derived from terrestrial sources. In recent years, marine ecosystems have proven to be a valuable resource of diverse natural products that are generated to defend and support their growth. Such marine sources offer a large opportunity for the identification of novel compounds that may guide the future development of new drugs and therapies. Using the National Oceanic and Atmospheric Administration (NOAA) portal, we explore deep-sea coral and sponge species inhabiting a segment of the U.S. Exclusive Economic Zone, specifically off the western coast of Florida. This area spans ~100,000 km2, containing coral and sponge species at sea depths up to 3000 m. Utilizing PubMed, we uncovered current knowledge on and gaps across a subset of these sessile organisms with regards to their natural products and mechanisms of altering cytoskeleton, protein trafficking, and signaling pathways. Since the exploitation of such marine organisms could disrupt the marine ecosystem leading to supply issues that would limit the quantities of bioactive compounds, we surveyed methods and technological advances that are necessary for sustaining the drug discovery pipeline including in vitro aquaculture systems and preserving our natural ecological community in the future. Collectively, our efforts establish the foundation for supporting future research on the identification of marine-based natural products and their mechanism of action to develop novel drugs and therapies for improving treatment regimens of human conditions and diseases. Full article
(This article belongs to the Special Issue Marine Natural Products and Signaling Pathways)
Show Figures

Figure 1

Figure 1
<p>NOAA standard data mining. (<b>A</b>) Sponges and corals identified within a region (spanning a 420 km by 210 km area) off the western coast of Florida obtained using data download method from the NOAA portal. The captured map extent displays the Global Ocean and Land Terrain (GEBCO) grid contours at the indicated depths. Scale bar: 20 miles. (<b>B</b>) Number of NOAA entries representing both corals and sponges from Panel A at various depths (in meters). (<b>C</b>) Spread of NOAA entries for both corals and sponges across various depths (in meters) and ridges/escarpments.</p>
Full article ">Figure 2
<p>Venn diagram representation of NOAA standard data mining results. The categories are represented as follows: category A (&lt;50 m); category B (≥50 to &lt;200 m); category C (≥200 to &lt;1000 m); and category D (≥1000 to 3000 m). Categories E to G represent corals and sponges collected at two different sea depths whereas categories I to L represent corals and sponges collected at three different sea depths. Category M represents sponges collected from all four sea depths whereas marine sponges and corals identified in categories O and N were uniquely found at those specific sea depths.</p>
Full article ">Figure 3
<p>NOAA manual data mining. (<b>A</b>) Sponges and corals identified within a region (spanning a 240 km by 420 km area) off the western coast of Florida obtained manually from the NOAA portal. The captured map extent displays sponges and corals identified within four quadrants (A, B, C, and D). Scale bar: &gt;30 miles. (<b>B</b>) Bar graph representation of the total number of NOAA entries of corals and sponges identified at varying depths within each of the four quadrants.</p>
Full article ">Figure 4
<p>Taxonomic classification of identified corals and sponges off the western coast of Florida. (<b>A</b>) In reference to <a href="#marinedrugs-21-00615-t001" class="html-table">Table 1</a>/<a href="#marinedrugs-21-00615-f002" class="html-fig">Figure 2</a>, classification of corals according to phylum, class, subclass, and order. Refer to <a href="#app1-marinedrugs-21-00615" class="html-app">Supplementary File S11</a> (“Corals” worksheet). (<b>B</b>) In reference to <a href="#marinedrugs-21-00615-t001" class="html-table">Table 1</a>/<a href="#marinedrugs-21-00615-f002" class="html-fig">Figure 2</a>, classification of sponges according to phylum, class, subclass, and order. Refer to <a href="#app1-marinedrugs-21-00615" class="html-app">Supplementary File S11</a> (“Sponges” worksheet).</p>
Full article ">Figure 5
<p>Schematic representation of the effects of coral-derived metabolites on cytoskeletal dynamics.</p>
Full article ">Figure 6
<p>Schematic representation of the effects of coral-derived metabolites on protein trafficking.</p>
Full article ">Figure 7
<p>Schematic representation of the effects of coral-derived metabolites on signaling cascades.</p>
Full article ">Figure 8
<p>Schematic representation of the effects of sponge-derived metabolites on cytoskeletal dynamics.</p>
Full article ">Figure 9
<p>Schematic representation of the effects of sponge-derived metabolites on protein trafficking.</p>
Full article ">Figure 10
<p>Schematic representation of the effects of sponge-derived metabolites on the PI3K/AKT and MAPK signaling networks.</p>
Full article ">Figure 11
<p>Schematic representation of the effects of sponge-derived metabolites on the JAK/STAT and cell surface receptor signaling networks.</p>
Full article ">Figure 12
<p>Venn diagram model presentation of similarities and differences between the pathways altered by bioactive compounds derived from corals and sponges. Commonalities between these two sessile invertebrates included deregulation of microtubules and actin filaments, induction of ER stress, activation of PI3K/AKT/MAPK, and JAK/STAT, and EGFR. Bioactive compounds from sponges also elicited bioactivities towards protein transport, intermediate filaments (such as vimentin), and other cell surface receptors including GPCR.</p>
Full article ">Figure 13
<p>Schematic of technological advancements to overcome the “supply issue” and sustain the drug discovery pipeline. Methods involving aquaculture and in vitro cultivation have been developed and applied to specific species of sponges and corals. Future advancements are needed in the fields of metagenomics and genetic engineering/recombinant DNA technology, which are presently limited.</p>
Full article ">
23 pages, 3017 KiB  
Article
Sponge Community Patterns in Mesophotic and Deep-Sea Habitats in the Aegean and Ionian Seas
by Caterina Stamouli, Vasilis Gerovasileiou and Eleni Voultsiadou
J. Mar. Sci. Eng. 2023, 11(11), 2204; https://doi.org/10.3390/jmse11112204 - 20 Nov 2023
Cited by 2 | Viewed by 3031
Abstract
Sponge assemblages play a significant role in the functioning of the Mediterranean benthic ecosystem. The main goal of this study was to investigate the diversity and distribution of poorly known sponge communities in the mesophotic and deep-sea substrates of the eastern Mediterranean Sea. [...] Read more.
Sponge assemblages play a significant role in the functioning of the Mediterranean benthic ecosystem. The main goal of this study was to investigate the diversity and distribution of poorly known sponge communities in the mesophotic and deep-sea substrates of the eastern Mediterranean Sea. More than 1500 sponge specimens belonging to 87 taxa were collected from 156 stations during experimental and commercial bottom trawling in the Aegean Sea and the eastern part of the Ionian ecoregion, at depths of between 10 and 800 m. A total of 79 sponge species were found in the Aegean and 40 species in the Ionian Sea. Eight of these species are included in lists of endangered and threatened species, two were newly recorded in the Aegean and six were first recorded in the east Ionian Sea. Both community structure and diversity differed between the two ecoregions. Species richness, biomass, abundance and diversity decreased with increasing depth, while different species dominated, in terms of biomass, abundance and frequency of appearance, in the two ecoregions and the separate depth zones. In contrast with previous investigations, which mostly examined shallow-water sponges, no clear resemblance patterns were observed among the north and south Aegean subareas, probably due to the homogeneity of the deep-sea habitats under investigation. This study, using sampling material from fish stock monitoring programs for the first time, contributed to our knowledge of the largely unknown eastern Mediterranean mesophotic and deep-sea sponge populations, which are subjected to intensive trawling activities. Full article
Show Figures

Figure 1

Figure 1
<p>Map of the sampling stations in both studied ecoregions. Bottom left corner: the location of the study area (Aegean and Ionian Sea ecoregions) in the Mediterranean Sea. AgA—North Aegean; AgB—Central-West Aegean; AgC—Central-East Aegean; AgD—South-West Aegean; Kythira—the marine area around Kythira Isl.; IoN—North Ionian; IoC—Central Ionian; IoS—South Ionian; and Korinthiakos—the Korinthiakos Gulf.</p>
Full article ">Figure 2
<p>Examples of typical sponge samples caught in trawl hauls after the onboard sorting procedures.</p>
Full article ">Figure 3
<p>Sponge distribution in the sampling stations of both studied ecoregions in terms of (<b>a</b>) biomass (kg/km<sup>2</sup>) and (<b>b</b>) abundance (specimens/km<sup>2</sup>). Bottom left corner—the location of the study area (Aegean and Ionian Sea ecoregions) in the Mediterranean Sea. AgA—North Aegean; AgB—Central-West Aegean; AgC—Central-East Aegean; AgD—South-West Aegean; Kythira—the marine area around Kythira Isl.; IoN—North Ionian; IoC—Central Ionian; IoS—South Ionian; and Korinthiakos—the Korinthiakos Gulf.</p>
Full article ">Figure 4
<p>Box plots of Shannon diversity index by (<b>a</b>) depth zone and (<b>b</b>) subareas: AgA—North Aegean; AgB—Central-West Aegean; AgC—Central-East Aegean; AgD—South-West Aegean; Kythira—the marine area around Kythira Isl.; IoN—North Ionian; IoC—Central Ionian; IoS—South Ionian; and Korinthiakos—the Korinthiakos Gulf.</p>
Full article ">Figure 5
<p>Box plots by depth zone for: (<b>a</b>) Aegean sponge species richness, (<b>b</b>) Ionian sponge species richness, (<b>c</b>) Aegean sponge biomass, (<b>d</b>) Ionian sponge biomass, (<b>e</b>) Aegean sponge abundance, and (<b>f</b>) Ionian sponge abundance.</p>
Full article ">Figure 6
<p>MDS showing two-dimensional resemblance of sponge community structures between the Aegean and Ionian Sea stations, based on: (<b>a</b>) biomass and (<b>b</b>) abundance; and between the different depth zones, based on: (<b>c</b>) biomass and (<b>d</b>) abundance.</p>
Full article ">Figure 7
<p>Species cumulative curves per depth zone for (<b>a</b>) Aegean Sea and (<b>b</b>) Ionian Sea.</p>
Full article ">Figure A1
<p>Photographs of the most common and some rare species found on the mesophotic and deep-sea soft bottoms of the Aegean and east Ionian Sea.</p>
Full article ">
82 pages, 7151 KiB  
Review
Peptides from Marine-Derived Fungi: Chemistry and Biological Activities
by Salar Hafez Ghoran, Fatemeh Taktaz, Emília Sousa, Carla Fernandes and Anake Kijjoa
Mar. Drugs 2023, 21(10), 510; https://doi.org/10.3390/md21100510 - 26 Sep 2023
Cited by 16 | Viewed by 3554
Abstract
Marine natural products are well-recognized as potential resources to fill the pipeline of drug leads to enter the pharmaceutical industry. In this circumstance, marine-derived fungi are one of the unique sources of bioactive secondary metabolites due to their capacity to produce diverse polyketides [...] Read more.
Marine natural products are well-recognized as potential resources to fill the pipeline of drug leads to enter the pharmaceutical industry. In this circumstance, marine-derived fungi are one of the unique sources of bioactive secondary metabolites due to their capacity to produce diverse polyketides and peptides with unique structures and diverse biological activities. The present review covers the peptides from marine-derived fungi reported from the literature published from January 1991 to June 2023, and various scientific databases, including Elsevier, ACS publications, Taylor and Francis, Wiley Online Library, MDPI, Springer, Thieme, Bentham, ProQuest, and the Marine Pharmacology website, are used for a literature search. This review focuses on chemical characteristics, sources, and biological and pharmacological activities of 366 marine fungal peptides belonging to various classes, such as linear, cyclic, and depsipeptides. Among 30 marine-derived fungal genera, isolated from marine macro-organisms such as marine algae, sponges, coral, and mangrove plants, as well as deep sea sediments, species of Aspergillus were found to produce the highest number of peptides (174 peptides), followed by Penicillium (23 peptides), Acremonium (22 peptides), Eurotium (18 peptides), Trichoderma (18 peptides), Simplicillium (17 peptides), and Beauveria (12 peptides). The cytotoxic activity against a broad spectrum of human cancer cell lines was the predominant biological activity of the reported marine peptides (32%), whereas antibacterial, antifungal, antiviral, anti-inflammatory, and various enzyme inhibition activities ranged from 7% to 20%. In the first part of this review, the chemistry of marine peptides is discussed and followed by their biological activity. Full article
(This article belongs to the Section Structural Studies on Marine Natural Products)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>FDA-approved marine peptide-based drugs [<a href="#B24-marinedrugs-21-00510" class="html-bibr">24</a>].</p>
Full article ">Figure 2
<p>Marine-derived peptides and their derivatives which are in clinical trials.</p>
Full article ">Figure 3
<p>Structures of <b>1</b>–<b>5</b>.</p>
Full article ">Figure 4
<p>Structures of <b>6</b>–<b>28</b>.</p>
Full article ">Figure 5
<p>Structures of <b>29</b>–<b>38</b>.</p>
Full article ">Figure 6
<p>Structures of <b>39</b>–<b>42</b>.</p>
Full article ">Figure 7
<p>Structures of <b>43</b>–<b>48</b>.</p>
Full article ">Figure 8
<p>Structures of <b>49</b>–<b>54</b>.</p>
Full article ">Figure 9
<p>Structures of <b>55</b> and <b>56</b>.</p>
Full article ">Figure 10
<p>Structures of <b>57</b>–<b>66</b>.</p>
Full article ">Figure 11
<p>Structures of <b>67</b>–<b>73</b>.</p>
Full article ">Figure 12
<p>Structures of <b>74</b>–<b>78</b>.</p>
Full article ">Figure 13
<p>Structures of <b>79</b>–<b>85</b>.</p>
Full article ">Figure 14
<p>Structures of <b>86</b>–<b>140</b>.</p>
Full article ">Figure 15
<p>Structures of <b>141</b>–<b>170</b>.</p>
Full article ">Figure 16
<p>Structures of <b>171</b>–<b>191</b>.</p>
Full article ">Figure 17
<p>Structures of <b>192</b>–<b>210</b>.</p>
Full article ">Figure 18
<p>Structures of <b>211</b>–<b>224</b>.</p>
Full article ">Figure 19
<p>Structures of <b>225</b>–<b>244</b>.</p>
Full article ">Figure 20
<p>Structures of <b>245</b>–<b>268</b>.</p>
Full article ">Figure 21
<p>Structures of <b>269</b>–<b>294</b>.</p>
Full article ">Figure 22
<p>Structures of <b>295</b> and <b>296</b>.</p>
Full article ">Figure 23
<p>Structures of <b>297</b>–<b>300</b>.</p>
Full article ">Figure 24
<p>Structure of <b>301</b>.</p>
Full article ">Figure 25
<p>Structures of <b>302</b>–<b>337</b>.</p>
Full article ">Figure 26
<p>Structures of <b>338</b>–<b>366</b>.</p>
Full article ">Figure 27
<p>The number of isolated peptides from marine-derived fungal resources.</p>
Full article ">Figure 28
<p>Mode of biological activities of peptides isolated from marine-derived fungal resources and their percentages.</p>
Full article ">
26 pages, 28034 KiB  
Article
Honeycomb Biosilica in Sponges: From Understanding Principles of Unique Hierarchical Organization to Assessing Biomimetic Potential
by Alona Voronkina, Eliza Romanczuk-Ruszuk, Robert E. Przekop, Pawel Lipowicz, Ewa Gabriel, Korbinian Heimler, Anika Rogoll, Carla Vogt, Milosz Frydrych, Pawel Wienclaw, Allison L. Stelling, Konstantin Tabachnick, Dmitry Tsurkan and Hermann Ehrlich
Biomimetics 2023, 8(2), 234; https://doi.org/10.3390/biomimetics8020234 - 3 Jun 2023
Cited by 9 | Viewed by 3306
Abstract
Structural bioinspiration in modern material science and biomimetics represents an actual trend that was originally based on the bioarchitectural diversity of invertebrate skeletons, specifically, honeycomb constructs of natural origin, which have been in humanities focus since ancient times. We conducted a study on [...] Read more.
Structural bioinspiration in modern material science and biomimetics represents an actual trend that was originally based on the bioarchitectural diversity of invertebrate skeletons, specifically, honeycomb constructs of natural origin, which have been in humanities focus since ancient times. We conducted a study on the principles of bioarchitecture regarding the unique biosilica-based honeycomb-like skeleton of the deep-sea glass sponge Aphrocallistes beatrix. Experimental data show, with compelling evidence, the location of actin filaments within honeycomb-formed hierarchical siliceous walls. Principles of the unique hierarchical organization of such formations are discussed. Inspired by poriferan honeycomb biosilica, we designed diverse models, including 3D printing, using PLA-, resin-, and synthetic-glass-prepared corresponding microtomography-based 3D reconstruction. Full article
(This article belongs to the Special Issue Bio-Inspired Design for Structure Applications)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Natural biopolymer-based honeycomb structures produced by bees (<b>a</b>,<b>b</b>), or produced by wasps (<b>c</b>,<b>d</b>) have been reproduced numerous times through artificial honeycomb-like structures in larger dimensions, such asthis plastic network with 5 cmlarge segments in diameter (<b>e</b>).</p>
Full article ">Figure 2
<p>Photograph of honeycomb-like siliceous skeleton of the glass sponge <span class="html-italic">Aphrocallistesbeatrix</span>: (<b>a</b>) view from above: sieve-plate covers the atrial cavity; (<b>b</b>) view from the side: in the lower part several holdfast-like points (arrows) of attachment to rocky substrate are seen.</p>
Full article ">Figure 3
<p>Combination of both hexangular and triangular geometries in hierarchical honeycomb-like structures of the <span class="html-italic">A. beatrix</span> glassy skeleton is well visible on digital light microscopy images represented here. This glass sponge constructs unique scaffolds that consist of regularly arrayed hexagonal cylinders.</p>
Full article ">Figure 4
<p>Digital microscopy of the honeycomb structures in the <span class="html-italic">A. beatrix</span> skeleton observed under diverse angles and magnification (<b>a</b>–<b>d</b>). Additionally, the honeycomb-like and triangular structural motifs are to be found in the glassy walls of the sponge skeleton (<b>d</b>–<b>f</b>). Corresponding measurements are represented in the <a href="#app1-biomimetics-08-00234" class="html-app">Supplementary Materials (see Figures S1 and S2</a>).</p>
Full article ">Figure 5
<p>SEM images of the <span class="html-italic">A. beatrix</span> sponge skeleton sample: (<b>a</b>,<b>b</b>) hexangular honeycomb-like macroporous structure; (<b>c</b>,<b>d</b>) the glassy wall of the pore with the triangular structure; (<b>e</b>) axial filament (arrow) in the axial channel inside the siliceous skeleton. (See also <a href="#biomimetics-08-00234-f006" class="html-fig">Figure 6</a>).</p>
Full article ">Figure 6
<p>SEM imagery of the partially demineralized axial filament from a selected sample of <span class="html-italic">A. beatrix</span> glass sponge. This organic structure is made of a network of twisted nanofibers connectedwith each other through nano-bridges (arrows), afeature that between structural proteins is characteristic only for actin filaments.</p>
Full article ">Figure 7
<p>CMXRF of <span class="html-italic">A. beatrix</span> honeycomb skeleton: mosaic image of the analyzed sample with the marked analysis area of a hexangular porein green (<b>a</b>); 3D distribution images of silicon (blue) (<b>b</b>) and potassium (red) (<b>c</b>) within a volume of 2.0 × 2.0 × 1.5 mm; transparent silicon distribution image (scalar opacity unit distance of 0.5) with visualized x-z contours (white to dark red) of the hexagonal cell wall (<b>d</b>) and three x-y contours at different z positions (magenta, cyan, orange) with (<b>e</b>) and without (<b>f</b>) silicon visualization.</p>
Full article ">Figure 8
<p>Digital light microscopy images of the organic-free skeleton fragment of <span class="html-italic">A. beatrix</span> sponge (<b>a</b>) and the same after demineralization with 10% HF (<b>b</b>). The organic scaffold (<b>b</b>) resembles the size and shape of the siliceous exoskeleton of this sponge (<b>a</b>) very well. See also <a href="#biomimetics-08-00234-f009" class="html-fig">Figure 9</a> and <a href="#biomimetics-08-00234-f010" class="html-fig">Figure 10</a>.</p>
Full article ">Figure 9
<p>Principal schematic view of isolation of an organic matrix from glass sponge skeleton. Digital (<b>left</b>) and fluorescence (<b>right</b>) microscopy of the honeycomb-like structures of the <span class="html-italic">A. beatrix</span> skeleton before and after demineralization as well as direct visualization of actin filaments with phalloidin staining.</p>
Full article ">Figure 10
<p>Fluorescence microscopy of axial filaments in with HF-desilicified <span class="html-italic">A. beatrix</span> skeletons: (<b>a</b>,<b>b</b>) stained with 594-Phalloidin; (<b>c</b>,<b>d</b>) stained with 488-Phalloidin. Arrows show the individual actin filaments agglomerated in bundles after desilicification.</p>
Full article ">Figure 11
<p>The schematic view of the 3D-printing process of <span class="html-italic">A. beatrix</span> skeleton 3D models.</p>
Full article ">Figure 12
<p>Microtomography-based3D reconstruction of <span class="html-italic">A. beatrix</span> glass sponge skeleton hierarchical structure.</p>
Full article ">Figure 13
<p>The schematic view of the 3D model of <span class="html-italic">A. beatrix</span>: (<b>a</b>) cylindrical flat honeycomb-like structure; (<b>b</b>,<b>c</b>) hierarchical tubular honeycombs with triangular holes in the walls.</p>
Full article ">Figure 14
<p>A. 3D CAD model of <span class="html-italic">A. beatrix</span> glass sponge skeleton with equal shapes; technical drawing with dimensions of the model, honeycomb, triangles in cross-section.</p>
Full article ">Figure 15
<p>(<b>a</b>) The process of 3D printing of the model of <span class="html-italic">A. beatrix</span> skeleton from PLA using FDM technology; (<b>b</b>,<b>c</b>) general view of the resulting CAD-based simplest cylindrical model; and (<b>d</b>,<b>e</b>) digital microscopy of a designed model wall.</p>
Full article ">Figure 16
<p>(<b>a</b>) The process of 3D printing in the reconstruction of <span class="html-italic">A. beatrix</span> skeleton; the resulting models made from: (<b>b</b>) PLA and (<b>c</b>) resin.</p>
Full article ">Figure 17
<p>Digital microscopy image of wall fragment with the dictyonal skeleton of <span class="html-italic">Lefroyella ceramensis</span> from the Emperors Mountain Chain. Lo—lareal oscula; r—ridges; db—dichotomous branching of ridges.</p>
Full article ">Figure 18
<p>(<b>a</b>) Scheme of the wall fragment of <span class="html-italic">Lefroyella</span>. A: horizontal section; w—wall; r—ridges. B: longitudinal section; r—ridges; ir—intercalary ridges; b—bridgees; lo—lateral oscula; db—dichotomous branching of ridges. (<b>b</b>) Scheme of the wall fragment of <span class="html-italic">Aphrocallistes</span>. A: horizontal section. B: longitudinal section; dc—honey-comb unit. C: longitudinal section, arrows show the suggested direction of the growth of the dictyonal skeleton; c—carina (line of fusion; db—dichotomous branching). It is of note here that dichotomous branching of actin filaments remains to be characteristic for this structural protein [<a href="#B67-biomimetics-08-00234" class="html-bibr">67</a>].</p>
Full article ">Figure 19
<p>The 3D model of <span class="html-italic">A. beatrix</span> made from synthetic glass: (<b>a</b>) flat gyroid (dimensions 20 mm × 20 mm × 6 mm), (<b>b</b>) cubic gyroid (30 mm × 30 mm × 30 mm) and (<b>c</b>) flat structure relatively similar to the sponge skeleton samples.</p>
Full article ">Figure 20
<p>Digital microscopic images of the 3D-reconstruction-based model (<b>a</b>) of <span class="html-italic">A. beatrix</span> glass sponge skeleton made of PLA covered with the diatomite layer (arrows). Diatomite remains to be strongly attached to the surface of the PLA-based construct even after sonication (<b>b</b>,<b>c</b>).</p>
Full article ">
2 pages, 184 KiB  
Correction
Correction: Back et al. A New Micromonospora Strain with Antibiotic Activity Isolated from the Microbiome of a Mid-Atlantic Deep-Sea Sponge. Mar. Drugs 2021, 19, 105
by Catherine R. Back, Henry L. Stennett, Sam E. Williams, Luoyi Wang, Jorge Ojeda Gomez, Omar M. Abdulle, Thomas Duffy, Christopher Neal, Judith Mantell, Mark A. Jepson, Katharine R. Hendry, David Powell, James E. M. Stach, Angela E. Essex-Lopresti, Christine L. Willis, Paul Curnow and Paul R. Race
Mar. Drugs 2023, 21(4), 214; https://doi.org/10.3390/md21040214 - 28 Mar 2023
Viewed by 1224
Abstract
After publication of this article [...] Full article
22 pages, 3642 KiB  
Article
Deep-Sea Epibenthic Megafaunal Assemblages of the Falkland Islands, Southwest Atlantic
by T. R. R. Pearman, Paul E. Brewin, Alastair M. M. Baylis and Paul Brickle
Diversity 2022, 14(8), 637; https://doi.org/10.3390/d14080637 - 10 Aug 2022
Cited by 4 | Viewed by 3202
Abstract
Deep-sea environments face increasing pressure from anthropogenic exploitation and climate change, but remain poorly studied. Hence, there is an urgent need to compile quantitative baseline data on faunal assemblages, and improve our understanding of the processes that drive faunal assemblage composition in deep-sea [...] Read more.
Deep-sea environments face increasing pressure from anthropogenic exploitation and climate change, but remain poorly studied. Hence, there is an urgent need to compile quantitative baseline data on faunal assemblages, and improve our understanding of the processes that drive faunal assemblage composition in deep-sea environments. The Southwest Atlantic deep sea is an undersampled region that hosts unique and globally important faunal assemblages. To date, our knowledge of these assemblages has been predominantly based on ex situ analysis of scientific trawl and fisheries bycatch specimens, limiting our ability to characterise faunal assemblages. Incidental sampling and fisheries bycatch data indicate that the Falkland Islands deep sea hosts a diversity of fauna, including vulnerable marine ecosystem (VME) indicator taxa. To increase our knowledge of Southwest Atlantic deep-sea epibenthic megafauna assemblages, benthic imagery, comprising 696 images collected along the upper slope (1070–1880 m) of the Falkland Islands conservation zones (FCZs) in 2014, was annotated, with epibenthic megafauna and substrata recorded. A suite of terrain derivatives were also calculated from GEBCO bathymetry and oceanographic variables extracted from global models. The environmental conditions coincident with annotated image locations were calculated, and multivariate analysis was undertaken using 288 ‘sample’ images to characterize faunal assemblages and discern their environmental drivers. Three main faunal assemblages representing two different sea pen and cup coral assemblages, and an assemblage characterised by sponges and Stylasteridae, were identified. Subvariants driven by varying dominance of sponges, Stylasteridae, and the stony coral, Bathelia candida, were also observed. The fauna observed are consistent with that recorded for the wider southern Patagonian Slope. Several faunal assemblages had attributes of VMEs. Faunal assemblages appear to be influenced by the interaction between topography and the Falkland Current, which, in turn, likely influences substrata and food availability. Our quantitative analyses provide a baseline for the southern Patagonian shelf/slope environment of the FCZs, against which to compare other assemblages and assess environmental drivers and anthropogenic impacts. Full article
(This article belongs to the Special Issue Deep Atlantic Biodiversity)
Show Figures

Figure 1

Figure 1
<p>Location map of the Falkland Islands and (<b>A</b>–<b>C</b>) zoomed insets of drop-down camera locations (indicated as black circles). The Falklands conservation zones encompass the Falklands Interim Conservation and Management Zone (FICZ) and the Falklands Outer Conservation Zone (FOCZ). (See <a href="#app1-diversity-14-00637" class="html-app">supplementary Table S1</a> for full station list and locations).</p>
Full article ">Figure 2
<p>Example images of fauna and substrata. (<b>A</b>) <span class="html-italic">Bathelia candida</span> reef observed at 1280 m water depth. (<b>B</b>) Mixed cold-water corals, including <span class="html-italic">Stylaster densicaulis</span> and Primnoidae, Stylasteridae, and Porifera morphospecies observed at 1280 m water depth. (<b>C</b>) Stylasteridae and Massive Ball Porifera morphospecies were observed from mixed substratum at 1595 m water depth. (<b>D</b>) Solitary cup corals, including <span class="html-italic">Flabellum</span> sp. and Scleractinia sp. 5, were observed from coarse substratum at 1406 m water depth. (<b>E</b>) <span class="html-italic">Anthoptilum grandiflorum</span> and <span class="html-italic">Flabellum</span> sp. observed from sandy mud at 1225 m water depth. (<b>F</b>) Pennatulacea sp. and <span class="html-italic">Flabellum</span> sp. observed from sandy mud at 1330 m water depth. Scale bar = 30 cm.</p>
Full article ">Figure 3
<p>nMDS plot of multivariate Hellinger distance matrix of transformed morphospecies density data. Samples are coloured to represent the five clusters identified by hierarchal clustering analysis.</p>
Full article ">Figure 4
<p>Spatial distribution of clusters identified from hierarchal clustering of Hellinger distance matrix of transformed morphospecies density data.</p>
Full article ">Figure 5
<p>Canonical redundancy analysis of a Hellinger distance matrix of transformed morphospecies density data and selected environmental variables. For clarity, the triplot is displayed in three separate plots with varying axis limits. (<b>A</b>) Environmental variables and sites colour coded by substratum type. (<b>B</b>) Morphospecies data, only fauna with the strongest effect are labelled. (<b>C</b>) Sites colour coded to represent hierarchal clustering. The vector arrowheads represent high, the origin averages, and the tail (when extended through the origin) low values of the selected environmental variables. Sites are represented by circles. Sites close to one another tend to have similar faunal composition that those further apart.</p>
Full article ">
69 pages, 21855 KiB  
Review
Anthraquinones and Their Analogues from Marine-Derived Fungi: Chemistry and Biological Activities
by Salar Hafez Ghoran, Fatemeh Taktaz, Seyed Abdulmajid Ayatollahi and Anake Kijjoa
Mar. Drugs 2022, 20(8), 474; https://doi.org/10.3390/md20080474 - 25 Jul 2022
Cited by 32 | Viewed by 5484
Abstract
Anthraquinones are an interesting chemical class of polyketides since they not only exhibit a myriad of biological activities but also contribute to managing ecological roles. In this review article, we provide a current knowledge on the anthraquinoids reported from marine-derived fungi, isolated from [...] Read more.
Anthraquinones are an interesting chemical class of polyketides since they not only exhibit a myriad of biological activities but also contribute to managing ecological roles. In this review article, we provide a current knowledge on the anthraquinoids reported from marine-derived fungi, isolated from various resources in both shallow waters such as mangrove plants and sediments of the mangrove habitat, coral reef, algae, sponges, and deep sea. This review also tentatively categorizes anthraquinone metabolites from the simplest to the most complicated scaffolds such as conjugated xanthone–anthraquinone derivatives and bianthraquinones, which have been isolated from marine-derived fungi, especially from the genera Apergillus, Penicillium, Eurotium, Altenaria, Fusarium, Stemphylium, Trichoderma, Acremonium, and other fungal strains. The present review, covering a range from 2000 to 2021, was elaborated through a comprehensive literature search using the following databases: ACS publications, Elsevier, Taylor and Francis, Wiley Online Library, MDPI, Springer, and Thieme. Thereupon, we have summarized and categorized 296 anthraquinones and their derivatives, some of which showed a variety of biological properties such as enzyme inhibition, antibacterial, antifungal, antiviral, antitubercular (against Mycobacterium tuberculosis), cytotoxic, anti-inflammatory, antifouling, and antioxidant activities. In addition, proposed biogenetic pathways of some anthraquinone derivatives are also discussed. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Plausible biosynthetic pathways of fungal anthraquinones.</p>
Full article ">Figure 2
<p>Anthraquinone scaffolds reported from marine-derived fungi.</p>
Full article ">Figure 3
<p>Structures of <b>1</b>–<b>19</b>.</p>
Full article ">Figure 4
<p>Structures of <b>20</b>–<b>43</b>.</p>
Full article ">Figure 5
<p>Structures of <b>44</b>–<b>59</b>.</p>
Full article ">Figure 6
<p>Structures of <b>60</b>–<b>78</b>.</p>
Full article ">Figure 7
<p>Structures of <b>79</b>–<b>95</b>.</p>
Full article ">Figure 8
<p>Structures of <b>96</b> and <b>97</b>.</p>
Full article ">Figure 9
<p>Structures of <b>98</b>–<b>121</b>.</p>
Full article ">Figure 10
<p>Structures of <b>122</b>–<b>134</b> and a plausible biosynthesis of <b>133</b> and <b>134</b>.</p>
Full article ">Figure 11
<p>Structures of <b>135</b>–<b>152</b>.</p>
Full article ">Figure 12
<p>Structures of <b>153</b>–<b>155</b>.</p>
Full article ">Figure 13
<p>Structures of <b>156</b>–<b>159</b>.</p>
Full article ">Figure 14
<p>Structures of <b>160</b>–<b>167</b>.</p>
Full article ">Figure 15
<p>Structures of <b>168</b>–<b>180</b>.</p>
Full article ">Figure 16
<p>Structures of <b>181</b>–<b>189</b>.</p>
Full article ">Figure 17
<p>Structures of <b>190</b>–<b>193</b> and plausible biosynthetic pathways of <b>190</b> and <b>192</b>.</p>
Full article ">Figure 18
<p>Structures of <b>194</b>–<b>196</b> and plausible biosynthetic pathways of <b>194</b>–<b>196</b>.</p>
Full article ">Figure 19
<p>Structures of <b>197</b>–<b>227</b>.</p>
Full article ">Figure 20
<p>Structures of <b>228</b>–<b>232</b>.</p>
Full article ">Figure 21
<p>Plausible biosynthetic pathway for <b>233</b>.</p>
Full article ">Figure 22
<p>Structures of <b>234</b>–<b>236</b>.</p>
Full article ">Figure 23
<p>Structures of <b>237</b> and <b>238</b>.</p>
Full article ">Figure 24
<p>Structures of <b>239</b>–<b>252</b>.</p>
Full article ">Figure 25
<p>Structures of <b>253</b>–<b>269</b>.</p>
Full article ">Figure 26
<p>Structures of <b>270</b>–<b>278</b> and a plausible biosynthetic pathway of <b>278</b>.</p>
Full article ">Figure 27
<p>Structures of <b>279</b>–<b>296</b>.</p>
Full article ">Figure 28
<p>The number of isolated anthraquinone metabolites and their derivatives from the marine-derived fungal resources.</p>
Full article ">
16 pages, 2386 KiB  
Article
A Novel Iridovirus Discovered in Deep-Sea Carnivorous Sponges
by Marta Canuti, Gabrielle Large, Joost T. P. Verhoeven and Suzanne C. Dufour
Viruses 2022, 14(8), 1595; https://doi.org/10.3390/v14081595 - 22 Jul 2022
Cited by 6 | Viewed by 2972
Abstract
Carnivorous sponges (family Cladorhizidae) use small invertebrates as their main source of nutrients. We discovered a novel iridovirus (carnivorous sponge-associated iridovirus, CaSpA-IV) in Chondrocladia grandis and Cladorhiza oxeata specimens collected in the Arctic and Atlantic oceans at depths of 537–852 m. The sequenced [...] Read more.
Carnivorous sponges (family Cladorhizidae) use small invertebrates as their main source of nutrients. We discovered a novel iridovirus (carnivorous sponge-associated iridovirus, CaSpA-IV) in Chondrocladia grandis and Cladorhiza oxeata specimens collected in the Arctic and Atlantic oceans at depths of 537–852 m. The sequenced viral genome (~190,000 bp) comprised 185 predicted ORFs, including those encoding 26 iridoviral core proteins, and phylogenetic analyses showed that CaSpA-IV is a close relative to members of the genus Decapodiridovirus and highly identical to a partially sequenced virus pathogenic to decapod shrimps. CaSpA-IV was found in various anatomical regions of six C. grandis (sphere, stem, root) from the Gulf of Maine and Baffin Bay and of two C. oxeata (sphere, secondary axis) from Baffin Bay. Partial MCP sequencing revealed a divergent virus (CaSpA-IV-2) in one C. oxeata. The analysis of a 10 nt long tandem repeat showed a number of repeats consistent across sub-sections of the same sponges but different between animals, suggesting the presence of different strains. As the genetic material of crustaceans, particularly from the zooplanktonic copepod order Calanoida, was identified in the investigated samples, further studies are required to elucidate whether CaSpA-IV infects the carnivorous sponges, their crustacean prey, or both. Full article
(This article belongs to the Special Issue Iridoviruses)
Show Figures

Figure 1

Figure 1
<p>Morphology of sponges within the species <span class="html-italic">Chondrocladia grandis</span> (<b>A</b>) and <span class="html-italic">Cladorhiza oxeata</span> (<b>B</b>). Arrowheads indicate different anatomical regions. Ax: axis; Sa: secondary axis; Ba: basal axis; St: stem; Rt: root; Sp: sphere. Figure reproduced with permission from Verhoeven and Dufour 2017 [<a href="#B7-viruses-14-01595" class="html-bibr">7</a>].</p>
Full article ">Figure 2
<p>Genome organization of CaSpA-IV. Circular map of the 190,288-bp genome (<b>A</b>). The outer scale is numbered clockwise in kbp. The outer circle depicts predicted ORFs on the forward (right-pointing arrows) and reverse (left-pointing arrows) strands, while the inner circle shows the G/C (blue) and A/T (green) content throughout the genome sequence. Identified core genes are indicated in pink, while other genes are labeled depending on whether predicted proteins show homology to those of other iridoviruses, as shown by the legend. Genetic map showing predicted protein function (<b>B</b>). Arrows represent viral ORFs with their size, position, and orientation (right or left arrowheads) indicated. ORFs coding for proteins with a predicted known function are colored based on the functional category as indicated in the legend at the top. Encoded putative proteins are labeled with the sequential position in the genome followed by the ORF orientation (R: right; L: left) and predicted function. PCNA: proliferating cell nuclear antigen, Rad50: DNA double-strand break repair rad50 ATPase-like protein, HMG: high mobility group protein homolog, TFIIS: transcription elongation factor SII, NF-H1: neurofilament triplet H1-like protein, MMP: myristylated membrane protein, RNRα: ribonucleoside-diphosphate reductase large subunit, DdRP: DNA-dependent RNA polymerase. Genomic positions are expressed in Kbp.</p>
Full article ">Figure 3
<p>Phylogenetic analysis of CaSpA-IV and other iridoviruses. The maximum likelihood phylogenetic tree was built using IQ-Tree and a partition model with a concatenation of 20 alignments of predicted protein sequences encoded by the core genes common to all iridoviruses (<a href="#viruses-14-01595-t001" class="html-table">Table 1</a>). Trees built with the individual alignments and the respective distance models used for phylogenetic inference are available in <a href="#app1-viruses-14-01595" class="html-app">Supplementary Figure S1</a>. The outcomes of the SH-aLRT and bootstrap test (1000 replicates) are shown for the main nodes. Subfamily and genus designations are indicated on the right, and the virus identified in this study is indicated in red. Virus full names and accession numbers are available in <a href="#app1-viruses-14-01595" class="html-app">Supplementary Table S2</a>.</p>
Full article ">Figure 4
<p>Phylogenetic analysis of the MCP of CaSpA-IV. The maximum likelihood phylogenetic tree was built with a 341 aa alignment of the MCP protein using IQ-Tree with the LG + I + G4 model. The outcomes of the SH-aLRT and bootstrap test (1000 replicates) are shown for the main nodes. Branches corresponding to genera of the <span class="html-italic">Alphairidovirinae</span> are collapsed and are indicated by white triangles. Genus designations, when available, are indicated on the right, and the virus identified in this study is indicated in red. Viruses of crustaceans are labeled with a colored circle as indicated in the legend, and the country of identification and the environment where the crustacean hosts were sampled are indicated in the strain name. Virus full names and accession numbers are available in <a href="#app1-viruses-14-01595" class="html-app">Supplementary Table S3</a>.</p>
Full article ">
13 pages, 1469 KiB  
Article
High-Throughput Screening of a Marine Compound Library Identifies Anti-Cryptosporidium Activity of Leiodolide A
by Rachel M. Bone Relat, Priscilla L. Winder, Gregory D. Bowden, Esther A. Guzmán, Tara A. Peterson, Shirley A. Pomponi, Jill C. Roberts, Amy E. Wright and Roberta M. O’Connor
Mar. Drugs 2022, 20(4), 240; https://doi.org/10.3390/md20040240 - 30 Mar 2022
Cited by 7 | Viewed by 3817
Abstract
Cryptosporidium sp. are apicomplexan parasites that cause significant morbidity and possible mortality in humans and valuable livestock. There are no drugs on the market that are effective in the population most severely affected by this parasite. This study is the first high-throughput screen [...] Read more.
Cryptosporidium sp. are apicomplexan parasites that cause significant morbidity and possible mortality in humans and valuable livestock. There are no drugs on the market that are effective in the population most severely affected by this parasite. This study is the first high-throughput screen for potent anti-Cryptosporidium natural products sourced from a unique marine compound library. The Harbor Branch Oceanographic Institute at Florida Atlantic University has a collection of diverse marine organisms some of which have been subjected to medium pressure liquid chromatography to create an enriched fraction library. Numerous active compounds have been discovered from this library, but it has not been tested against Cryptosporidium parvum. A high-throughput in vitro growth inhibition assay was used to test 3764 fractions in the library, leading to the identification of 23 fractions that potently inhibited the growth of Cryptosporidium parvum. Bioassay guided fractionation of active fractions from a deep-sea sponge, Leiodermatium sp., resulted in the purification of leiodolide A, the major active compound in the organism. Leiodolide A displayed specific anti-Cryptosporidium activity at a half maximal effective concentration of 103.5 nM with selectivity indexes (SI) of 45.1, 11.9, 19.6 and 14.3 for human ileocecal colorectal adenocarcinoma cells (HCT-8), human hepatocellular carcinoma cells (Hep G2), human neuroblastoma cells (SH-SY5Y) and green monkey kidney cells (Vero), respectively. The unique structure of leiodolide A provides a valuable drug scaffold on which to develop new anti-Cryptosporidium compounds and supports the importance of screening natural product libraries for new chemical scaffolds. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Scatter plot of the anti-<span class="html-italic">Cryptosporidium</span> activity of 3764 fractions from the HBOI library. All enriched fractions were screened at 10 ug/mL. A cut-off of 80% inhibition of <span class="html-italic">C. parvum</span> proliferation (dotted red line) in at least one of the two tested wells was applied to yield 222 hits for a hit rate of 5.8%. All points are normalized to a 0.1% DMSO vehicle control and log<sub>10</sub> transformed. Red and purple asterisks mark the location of the fraction from which leiodolide A was isolated.</p>
Full article ">Figure 2
<p>Testing strategy and results of high-throughput phenotypic screen of HBOI enriched fraction library for non-cytotoxic, anti-<span class="html-italic">Cryptosporidium</span> fractions. All fractions were tested at 10 μg/mL.</p>
Full article ">Figure 3
<p>Published structure of leiodolide A [<a href="#B23-marinedrugs-20-00240" class="html-bibr">23</a>].</p>
Full article ">Figure 4
<p>Leiodolide A inhibits <span class="html-italic">C. parvum</span> growth at nanomolar concentrations without affecting host cells. (<b>A</b>) EC<sub>50</sub> of leiodolide A for intracellular <span class="html-italic">C. parvum</span> is 103.5 (95% CI 82.9–131.7) nM. HCT-8s were infected with <span class="html-italic">C. parvum</span> for 24 h and then infected cells were treated with the compound for 48 h. Data shown are the results of three independent experiments performed in triplicate. (<b>B</b>) IC<sub>50</sub> of leiodolide A for confluent HCT-8 cells treated for 48 h is 4670 (95% CI 3640–6070) nM giving a selectivity index of 45.1. Data were normalized to a 0.1% DMSO control. The results of four independent experiments performed in triplicate are shown.</p>
Full article ">
11 pages, 1482 KiB  
Article
Unveiling the Chemical Diversity of the Deep-Sea Sponge Characella pachastrelloides
by Sam Afoullouss, Anthony R. Sanchez, Laurence K. Jennings, Younghoon Kee, A. Louise Allcock and Olivier P. Thomas
Mar. Drugs 2022, 20(1), 52; https://doi.org/10.3390/md20010052 - 5 Jan 2022
Cited by 2 | Viewed by 5852
Abstract
Sponges are at the forefront of marine natural product research. In the deep sea, extreme conditions have driven secondary metabolite pathway evolution such that we might expect deep-sea sponges to yield a broad range of unique natural products. Here, we investigate the chemodiversity [...] Read more.
Sponges are at the forefront of marine natural product research. In the deep sea, extreme conditions have driven secondary metabolite pathway evolution such that we might expect deep-sea sponges to yield a broad range of unique natural products. Here, we investigate the chemodiversity of a deep-sea tetractinellid sponge, Characella pachastrelloides, collected from ~800 m depth in Irish waters. First, we analyzed the MS/MS data obtained from fractions of this sponge on the GNPS public online platform to guide our exploration of its chemodiversity. Novel glycolipopeptides named characellides were previously isolated from the sponge and herein cyanocobalamin, a manufactured form of vitamin B12, not previously found in nature, was isolated in a large amount. We also identified several poecillastrins from the molecular network, a class of polyketide known to exhibit cytotoxicity. Light sensitivity prevented the isolation and characterization of these polyketides, but their presence was confirmed by characteristic NMR and MS signals. Finally, we isolated the new betaine 6-methylhercynine, which contains a unique methylation at C-2 of the imidazole ring. This compound showed potent cytotoxicity towards against HeLa (cervical cancer) cells. Full article
(This article belongs to the Special Issue Discovering Marine Bioactive Compounds by Molecular Networking)
Show Figures

Figure 1

Figure 1
<p>Molecular network of <span class="html-italic">Characella pachastrelloides</span> fractions, with annotated metabolites. Pie charts indicate metabolites distribution in fractions (sum precursor ion intensity). Size of node is relative to precursor ion intensity. Edge width increases with higher cosine score.</p>
Full article ">Figure 2
<p>Structure of cyanocobalamin (<b>7</b>).</p>
Full article ">Figure 3
<p>Structure of the new histidine derivative 6-methylhercynine (<b>8</b>).</p>
Full article ">Figure 4
<p>Comparison between the experimental and calculated ECD spectra for two enantiomers of compound <b>8</b>.</p>
Full article ">Scheme 1
<p>Proposed metabolic pathway for <b>8</b>.</p>
Full article ">
Back to TopTop