Next Article in Journal
The Pentose Phosphate Pathway in Yeasts–More Than a Poor Cousin of Glycolysis
Next Article in Special Issue
Identification, Synthesis, Conformation and Activity of an Insulin-like Peptide from a Sea Anemone
Previous Article in Journal
Antibiotic Therapy of Plague: A Review
Previous Article in Special Issue
New Triterpene Glycosides from the Far Eastern Starfish Solaster pacificus and Their Biological Activity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Cytotoxic Compounds of Two Demosponges (Aplysina aerophoba and Spongia sp.) from the Aegean Sea

1
Institute of Pharmacy/Pharmacognosy, University of Innsbruck, Innrain 80-82, 6020 Innsbruck, Austria
2
Functional Proteo-Metabolomics, Department of Biochemistry, University of Innsbruck, Innrain 80-82, 6020 Innsbruck, Austria
3
Tyrolean Cancer Research Institute, 6020 Innsbruck, Austria
4
Faculty of Pharmacy, Phenikaa University, Hanoi 12116, Vietnam
5
A&A Green Phoenix Group JSC, Phenikaa Research and Technology Institute (PRATI), No. 167 Hoang Ngan, Trung Hoa, Cau Giay, Hanoi 11313, Vietnam
6
Institute of Pharmaceutical Biology and Phytochemistry, University of Münster, Corrensstraße 48, 48149 Münster, Germany
7
Cancer Cluster Salzburg, IIIrd Medical Department, Paracelsus Medical University Salzburg, Müllner Hauptstraße 48, 5020 Salzburg, Austria
*
Author to whom correspondence should be addressed.
Biomolecules 2021, 11(5), 723; https://doi.org/10.3390/biom11050723
Submission received: 10 March 2021 / Revised: 6 May 2021 / Accepted: 7 May 2021 / Published: 12 May 2021
(This article belongs to the Special Issue Marine Natural Compounds with Biomedical Potential)

Abstract

:
The class of demosponges is the biggest and most diverse of all described sponge species and it is reported to produce a plethora of chemically different metabolites with interesting biological activities. The focus of the present study was to investigate the chemical composition of two Mediterranean demosponges, targeting their brominated compounds and prenylated hydroquinones, compounds with interesting cytotoxic and anti-microbial properties. In order to gain a deeper insight into the chemical diversity of their metabolites and their activities, 20 pure secondary metabolites including new natural products were isolated from two different species (Aplysina aerophoba and Spongia sp.) using various chromatographic techniques. Their structures were confirmed by NMR and HRMS, revealing molecules with various chemical scaffolds, mainly prenylated hydroquinones from Spongia sp. and halogenated compounds from Aplysina aerophoba, including 5 novel natural products. The isolated compounds were investigated for their cytotoxic properties using 9 different cell lines, and especially one compound, 2,6-dibromo-4-hydroxy-4-methoxycarbonylmethylcyclohexa-2,5-dien-1-one showed good activities in all tested models.

1. Introduction

Sponges are a prolific source of bioactive natural compounds with unique structural features unprecedented in the terrestrial environment [1]. Demospongiae is their biggest and most diverse class [2], and they are known to produce a variety of chemically different metabolites, including terpenes, alkaloids, macrolides, peptides, betaines, ceramides, lipids, and halogenated compounds with potential interest regarding industrial and medical applications including antiviral, antitumor, antimicrobial, or generally cytotoxic properties. Therefore, they are of considerable biotechnological interest [3,4]. These compounds often possess multiple ecological functions, primarily the protection against predators, competitors for space, biofoulers, or opportunistic pathogenic microorganisms [5].
The sponge Aplysina aerophoba Schmidt, 1862 (Demospongiae, Verongida, Aplysinidae) is a common Mediterranean, photophilic species growing on stable substrates like rocks or rocky walls [6]. Its bright-yellow color is due to the highly labile pigment uranidine which undergoes polymerization when exposed to air [7]. It is known to harbor large amounts of bacteria and contains high concentrations of brominated alkaloids [8,9], mainly isoxazoline alkaloids, which after wound-induced bioconversion, are enzymatically cleaved to the lower molecular weight antibacterial compounds, showing 3,5 dibromotyrosin structure. These compounds protect the sponge from penetration of pathogenic bacteria [10,11]. The isolated brominated metabolites from A. aerophoba have been tested mainly for their cytostatic and antimicrobial activities, often showing moderate to strong effects [12,13]. Potential interest regarding industrial and medical applications has led to attempts for sponge cultivation and increased production of sponge biomass [14].
The genus Spongia, Linnaeus 1759, comprising of approximately 90 species, belongs to the family Spongidae of the order Dictyoceratida [15]. They have commercial value since they are harvested as natural bath sponges in the Mediterranean and the West Indies [16]. Sponges of the genus Spongia are unarmored, soft to firm, and compressible. They have a skeletal network of primary fibers, and they are a rich source of collagen proteins [16,17]. They mainly produce terpenes (sesquiterpene quinones, diterpenes, furanoterpenes, sesterterpenes), sterols, macrolides, and alkaloids [15]. Additionally, long chain lipid compounds may be considered as part of the chemical fingerprint of Spongia genus [15]. Regarding bioactivity, some of these metabolites, especially those with terpene structure, are anti-viral and antioxidant compounds, they act as immunomodulators, and they are active against different cancer cell lines [15,18].
The aim of the present study was the investigation of the chemical composition of two Mediterranean demosponges, Aplysina aerophoba and Spongia sp., targeting their brominated compounds and prenylated hydroquinones, compounds with interesting cytotoxic properties.
Fifteen metabolites with diverse structures were isolated from the ethanolic extract of A. aerophoba, including two new chemical constituents for this species, five new natural products, and eight known ones. Additionally, five known compounds could be isolated from the ethanolic extract of Spongia sp. (Figure 1). The cytotoxic effects of the pure metabolites were evaluated in nine different cell lines, including a human urinary bladder carcinoma cell line (T24), a human stomach carcinoma cell line (AGS), a human neuroblastoma cell line (SH-SY5Y), three human colon adenocarcinoma cell lines (DLD-1, SW-480, LOVO), and three myeloma cell lines (NCI-H929, OPM-2, U266).

2. Materials and Methods

2.1. Biological Material

Spongia sp. was collected at Heraklion, Crete, Greece in April 2018, and Aplysina aerophoba at the same place in June 2018. Both were morphologically identified by Dimitris Poursanidis (Postdoc fellow, IACM-FORTH, Foundation for Research and Technology—Heraklion, Crete, Greece—CEO of terraSolutions marine environment research). The sponges were cut into small pieces and immediately stored in EtOH 96% until further processing. Voucher samples are deposited at the Institute of Pharmacy, Pharmacognosy, University of Innsbruck, Austria.

2.2. Instrumentation

Optical rotations were measured with a polarimeter P-2000 (JASCO, Tokyo, Japan) using a 10.0 cm tube and CHCl3 as the solvent. IR spectra were obtained on a Platinum ATR FTIR spectrometer (Bruker, MA, USA), and ECD experiments conducted on a J-1500 spectrophotometer (JASCO, Tokyo, Japan). NMR experiments were performed on two spectrometers from Bruker, Bruker Avance II 600 (600 MHz for 1H, 150 MHz for 13C) and Avance III HD (400 MHz for 1H, 100 MHz for 13C). The isolated compounds were dissolved in MeOD or chloroform using tetramethylsilane (TMS) as internal standard. High-resolution mass spectra were measured with aQ-Exactive HF-X Orbitrap mass spectrometer (Thermo, MA, USA) and a micrOTOF-Q II mass spectrometer (Bruker-Daltonics, Bremen, Germany) whereas low-resolution mass spectra were recorded on an Agilent InfinityLab LC/MSD System. It comprised of an Agilent 1260 HPLC, equipped with binary pump, autosampler, column oven, and photodiode array detector (Santa Clara, CA, USA). For the purification of compounds, a Reveleris® X2 iES flash chromatography system (Büchi, Flawil, Switzerland) and a semi-preparative UltiMate 3000 HPLC from Dionex (Thermo, Waltham, MA, USA), comprising a P580 pump, an ASI 100 automated sample injector, an UVD 170 U detector, and a fraction collector, were used. Sephadex LH-20 material was purchased from Sigma-Aldrich (St. Louis, MI, USA). Analytical HPLC experiments were performed on a LC-20AD XR System (Shimadzu, Tokyo, Japan).

2.3. Chemicals and Reagents

All solvents required for extraction and isolation were purchased from VWR International (Vienna, Austria) and petroleum ether (PE), dichloromethane, acetone, and ethyl acetate (EtOAc) were distilled before use. Solvents for analytical experiments had pro analysis (p.a.) quality at least and were obtained from Merck (Darmstadt, Germany). Deuterated solvents were supplied by Euriso-Top (Saint-Aubin, France). Ultrapure water was produced by a Sartorius arium® 611 UV (Göttingen, Germany) purification system. Silica gel 40–63 μm and pre-packed cartridges for flash chromatography were purchased from Merck (Darmstadt, Germany) and Büchi (Flawil, Switzerland), respectively.

2.4. Extraction and Isolation

Aplysina aerophoba (approximately 1 kg) was extracted five times in an ultrasonic bath (Bandelin Sonorex 35 KHz, Berlin, Germany) for 15 min each using EtOH 96%. Afterwards, the ethanolic extract (89 g) was partitioned successively between PE, EtOAc, BuOH, and H2O (3 times each × 500 mL). HPLC analysis of the fractions indicated that the PE (6.2 g), EtOAc (14.0 g), and BuOH (8.7 g) fractions contained brominated compounds, therefore, they were combined and used for further fractionation, while the water fraction (58.0 g) was dismissed.
The combined PE, EtOAc, and BuOH fractions were fractionated on a silica gel column using PE, EtOAc, and methanol (from 10:0:0 to 0:10:0 to 0:0:10) as solvent, resulting in 27 subfractions (Figure 2). Fraction 5 (380 mg) was separated with flash chromatography using hexane-EtOAc (10:0–8:2) to give seven subfractions. Further purification of subfraction 6 (138 mg) on silica gel CC elution with heptane-EtOAc (10:0–8:2) yielded compound 15 (4 mg). Separation of fraction 10 (294 mg) by Sephadex LH-20 with CH2Cl2-acetone (85:15) resulted in 7 subfractions from which subfraction 7 was a pure compound (compound 12, 130 mg). Subfraction 4 (61 mg) was subjected to semi-preparative HPLC using an Aqua C18 125 Å column (250 × 10 mm, 5 μm; Phenomenex, Torrance, CA, USA). The mobile phase comprised of CH3CN and 0.1% formic acid (FA) in water to give compound 13 (7 mg). Fraction 14 (3 g) was separated with Sephadex LH-20 using CH2Cl2-acetone (85:15) and flash chromatography with PE-EtOAc (10:0–0:10) and then purified by semi-preparative HPLC using the above-mentioned Aqua C18 column (CH3CN-0.1% FA in water) to yield compound 7 (58 mg). Fraction 15 (433 mg) was separated into 7 subfractions by flash chromatography with H2O-MeOH (10:0–0:10). Subfraction 7 (127 mg) was purified on Sephadex LH-20 material with CH2Cl2-acetone (85:15) and semi-preparative HPLC (Aqua C18 column, CH3CN-0.1% FA in water) to yield compound 8 (32 mg), whereas subfraction 6 (23 mg) was purified with semi-preparative HPLC (Aqua C18 column CH3CN-0.1% FA in water) to give compounds 3 (2 mg) and 14 (6 mg). Compound 6 (32 mg) was obtained from fraction 17 (224 mg) by Sephadex LH-20 with CH2Cl2-acetone (85:15), while compounds 5 (32 mg) and 11 (32 mg) were obtained from fraction 19 (389 mg) using flash chromatography with EtOAc-MeOH (10:0–0:10). Fraction 22 was purified by flash chromatography with H2O-MeOH (10:0–0:10) and on Sephadex LH-20 material with MeOH to give compound 10 (28 mg). From fraction 25 (1.6 g), compound 9 (260 mg) was obtained by flash chromatography (H2O-MeOH, 10:0–0:10). Finally, fraction 23 (1.4 g) was subjected to flash chromatography with H2O-MeOH (10:0–0:10) as eluent to obtain seven subfractions. Semi-preparative HPLC of subfraction 7 using a Synergi Polar-RP 80Å column (250 × 4.6 mm, 4 μm; Phenomenex, USA) and MeOH and 0.1% FA in water as mobile phase resulted in compounds 1 (5 mg) and 2 (7 mg), while by semi-preparative HPLC of the combined subfractions 5 and 6 using the same Synergi Polar-RP column, compound 4 (2 mg) could be obtained.
The ethanolic extract of Spongia sp. was fractionated using similar approaches as described above, resulting in the isolation of 5 compounds (Figure 3).
Initially, the material was extracted with EtOH 96% by sonication, the ethanolic extract (58 g) partitioned successively with PE, EtOAc, BuOH, and H2O (3 times each × 500 mL) and after HPLC analysis, the PE (5.7 g), EtOAc (2.5 g) and BuOH (4.7 g) soluble fractions combined, as results suggested the presence of prenylated hydroquinones.
The combined PE, EtOAc, and BuOH fractions were fractionated on a silica gel column using PE, EtOAc, and methanol (from 10:0:0 to 0:10:0 to 0:0:10), resulting in 23 subfractions. Fraction 9 (860 mg) was separated with Sephadex LH-20 with CH2Cl2-acetone (85:15) followed by flash chromatography using PE-EtOAc (10:0–0:10) to give 5 subfractions. Further purification of subfraction 4 (158 mg) via flash chromatography with H2O-CH3CN (10:0–0:10) yielded compound 19 (46 mg). Separation of fraction 8 (1.3 g) on Sephadex LH-20 material (CH2Cl2-acetone = 85:15 as eluent) resulted in 9 subfractions. Subfraction 7 was purified by normal phase flash chromatography with PE-EtOAc (10:0–0:10) and reversed phase flash chromatography with H2O-CH3CN (10:0–0:10) to give compounds 16 (71 mg) and 17 (520 mg), while subfraction 7 (410 mg) was subjected to reversed phase flash chromatography with H2O-CH3CN (10:0–0:10) resulting in compound 18 (14 mg). Finally, fraction 5 (300 mg) could be purified using flash chromatography with PE-EtOAc (10:0–0:10) and reversed phase flash chromatography with H2O-CH3CN (10:0–0:10) to yield compound 20 (8 mg).

2.4.1. Compound 1: Yellowish Amorphous Powder

Rf 0.70 (EtOAc:MeOH = 8:2); UV (MeOH) λmax (log ε) 235 (3.7) nm; IR (KBr) νmax 3286, 2933, 1687, 1540, 1456, 1252, 1171, 1033, and 775 cm−1; 1H and 13C NMR see Table 1; HR-ESI-MS m/z 500.9340 [M-H] (calcd. for C14H19Br2N2O6S, 500.93250).

2.4.2. Compound 2: Colourless Amorphous Powder

Rf 0.33 (EtOAc:MeOH = 8:2); UV (MeOH) λmax (log ε) 236 (4.4) nm; IR (KBr) νmax 3236, 3061, 2930, 1657, 1628, 1541, 1471, 1455, 1420, 1396, 1256, 1199, 1021.9, 1001, 866, and 737 cm−1; 1H and 13C NMR see Table 2; HR-ESI-MS m/z 775.7943 [M-H] (calcd. for C21H22Br4N3O7S, 775.7906).

2.4.3. Compound 3: Yellowish Amorphous Powder

Rf 0.82 (EtOAc:MeOH = 8:2); α D 21 −13.2 (c 0.9, CHCl3); UV (acetonitrile) λmax (log ε) 236 (4.0) nm; IR (KBr) νmax 3313, 2926, 2854, 1744, 1662, 1542, 1471, 1422, 1397, 1338, 1258, 1080, 1000, 939, and 739 cm−1; 1H and 13C NMR see Table 2; HR-ESI-MS m/z 753.8063 [M-H] (calcd. for C22H20Br4N3O7, 753.8059).

2.4.4. Compound 4: Yellowish Amorphous Powder

Rf 0.67 (EtOAc:MeOH = 8:2); UV (MeOH) λmax (log ε) 236 (3.8) nm, 284 (3.5) nm; IR (KBr) νmax 3280, 2942, 2834, 1713, 1659, 1532, 1508, 1494, 1366, 1283, 1207, 1016, 930, 801, and 671 cm−1; 1H and 13C NMR see Table 1; HR-ESI-MS m/z 366.9912 [M+Na]+ (calcd. for C12H13BrN2NaO5, 366.9906).

2.4.5. Compound 5: Yellowish Amorphous Powder

Rf 0.78 (EtOAc:MeOH = 8:2); α D 21 −21.8 (c 0.1, CHCl3); UV (MeOH) λmax (log ε) 235 (3.8) nm; IR (KBr) νmax 3338, 2979, 2933, 2252, 1691, 1445, 1404, 1373, 1342, 1237, 1185, 1093, 1066, 764, and 706 cm−1; 1H and 13C NMR see Table 3; HR-ESI-MS m/z 256.1166 [M+H]+ calcd for C12H18NO5, 256.1185, 533.2119 [2M+H]+ calcd. for C24H34N2O10Na, 533.2111.

2.5. Computational Methods

The 3D structures of the molecules were drawn and subjected to conformational analysis in MacroModel v. 9 (Schrödinger LLC, New York, NY, USA) using OPLS-3 forcefield in chloroform. The conformers obtained in an energy window of 10 Kcal/mol were further submitted to geometrical optimization at B3LYP/6-31G(d,p) level for compounds 3 and 14, and B3LYP/6-31++(d,p) for compound 5. Further NMR chemical shift calculation was performed by using mPW1PW91/6-31G+(d,p)/CPCM/methanol for compounds 3 and 14 (data are not shown), and specific rotation calculation of compound 5 was conducted at B3LYP/6-311G++(d,p)/CPCM level in chloroform by considering the sodium D line frequency in the calculation. The obtained specific rotation values were Boltzmann-averaged and utilized for comparison with the experimentally obtained value in chloroform.

2.6. Cell Cultures

T24 cell line (ATCC HTB-4), derived from human urinary bladder carcinoma were obtained from Prof. Dr. Straube, University of Jena, GermanyT24 cells were cultured in Dulbecco’s modified Eagle medium (Merck Millipore, Darmstadt, Germany), supplemented with 10% (v/v) FCS (Merck) and 0.5% penicillin/streptomycin (Merck) at 5% CO2/37 °C. Passaging took place at 80 to 90% confluence of T24 cells.
The AGS cell line (300408, human stomach adenocarcinoma) were obtained from CLS Cell Lines Service GmbH and cells were cultured in RMPI-1640 (Merck), supplemented with 10% (v/v) FCS (Merck) and 1% penicillin/streptomycin (Merck) at 5% CO2/37 °C. Passaging took place at 80 to 90% confluence of AGS cells.
SH-SY5Y human neuroblastoma cell line was kindly provided by Dr. Obexer (Tyrolean Cancer Research Institute); human colon adenocarcinoma cell lines DLD-1, SW-480, and LOVO were purchased from DSMZ (Braunschweig, Germany). Human myeloma cell lines NCI-H929, OPM-2, and U266 were also purchased from DSMZ (Braunschweig, Germany). Cell lines were routinely fingerprinted and tested for mycoplasma negativity. Primary human foreskin fibroblasts were purchased from Promocell, Heidelberg, Germany.
PBMCs (peripheral blood mononuclear cells) from healthy donors were utilized after obtaining written consent at the University Hospital Salzburg (Ethics Committee approval 415-E/1287/6-2011). Cells were subjected to Ficoll separation (Ficoll PaqueTM, VWR, Darmstadt, Germany). All cell lines and primary cells were grown in RPMI-1640 (Life Technologies, Paisley, UK), supplemented with 10% fetal calf serum (PAA, Linz, Austria), L-glutamine 100 µg/mL (Biochrom, Berlin, Germany), and penicillin-streptomycin 100 U/mL (Applichem, Darmstadt, Germany).

2.7. Determination of Cell Viability of T24 and AGS Cell Lines (MTT Assay)

To determine the influence of the isolated substances on cell viability towards T24 bladder and AGS stomach cells, MTT assay was performed (Mosmann, 1983). T24 cells were seeded into 96-well plates with 2.0 × 104 cells per well (100 μL), incubated for 24 h at 37 °C with 5% CO2, and washed with 200 μL/well of PBS. Incubation of the cells with 100 μL of extract or pure substances at different concentrations (100 to 0.1 µM for pure compounds and 500–100 µg/mL for extracts in DMEM without additives) was performed for 24 h at 37 °C/5% CO2. Subsequently, the supernatant was removed and cells were washed twice with PBS (200 μL/well). In addition, 50 μL of MTT reagent were added to each well and after an incubation period of 24 h at 37 °C/5% CO2, the MTT reagent was removed and replaced by 50 µL DMSO per well to dissolve the formed insoluble formazan crystals. After 10 min, the amount of formazan was quantified spectrophotometrically in a plate reader at λ = 595 nm, with λ = 690 nm as a reference wavelength. Medium + 10% FCS served as a positive control, while the respective medium used for the sample preparation served as an untreated control. As a negative control, 10% DMSO was used. For AGS, a cell density of 5 × 104 was used and substances were diluted in RPMI medium. At least three analyses in triplicates were performed for each cell line and each concentration of the compounds tested and a solvent control was always included. Data are shown as mean percentage of viable cells and standard error of the mean (SEM) (error bars).

2.8. Cytotoxicity Assays Using FACS Analysis

The induction of apoptosis was measured in cancer cell lines and in fibroblasts/peripheral blood mononuclear cells of healthy donors using established protocols [19]. Adherent cell lines (neuroblastoma cell line SH-SY5Y, colon carcinoma cell lines DLD-1, SW-480, and LOVO, as well as primary foreskin fibroblasts) were seeded at a concentration of 1 × 105 cells/mL in 96-well plates the day before treatment in order to allow attachment. This medium was then replaced by media containing the respective compounds at different concentrations and cells were incubated for 24 h. Solvent controls were always included. Supernatants were collected and pooled with the trypsinized cells of the respective wells. These samples were then stained with AnnexinV-FITC (MabTag GmbH, Friesoythe, Germany) and propidium iodide (Merck, Darmstadt, Germany) and processed by flow cytometry using FACS Canto II and Diva software (Becton Dickinson, San Jose, CA, USA) Biosciences). Data were further analyzed using GraphPad Prism 5.0 software.
Myeloma cell lines NCI-H929, OPM-2, and U266 as well as peripheral blood mononuclear cells (PBMCs) of healthy donors (5 × 105 cells/mL) were incubated with the same compounds for 24 h and subjected to FACS analysis as described above. Etoposide (Merck, Darmstadt, Germany) was used as positive control for adherent cancer cell lines and Bortezomib (Eubio, Vienna, Austria) was used as positive control for myeloma cell lines. The extent of non-apoptotic cells (AnnexinV/propidium iodide negativity) was calculated as percentage of control (untreated) and mean percentage of viable cells and standard deviation (error bars) are shown.

3. Results

3.1. Structure Elucidation

Five out of fifteen compounds isolated from A. aerophoba were novel natural products and their structure elucidation is described below.

3.1.1. Compound 1

This substance was assigned to the molecular formula C14H20Br2N2O6S as established by a negative ion [M-H] at m/z 500.9340 (calcd. for C14H19Br2N2O6S 500.9325) in HR-ESI mass spectrum. Characteristic NMR chemical shifts (δH 7.49 for H-7, 9 and δC 152.9, 119.0, 134.5, 140.9 for C-5, 6, 7, 8, respectively) and MS patterns indicated the presence of a di-bromo substituted phenolic ring. The COSY spectrum revealed three coupling networks (Figure 4), including the protons of the methylenes H-2/H-3/H-4 and H-11/H-12 and the protons of the ethoxy group H-13/H-14. The substructures were connected through a carbamic acid bond by specific chemical shifts of position 1 (δC 159.3), 2 (δH 3.38, δC 39.2), and 13 (δH 4.07, δC 61.8) and key HMBC correlations of H-2, 13/C-1 (δC 159.2). Additionally, the HMBC correlations of H-4 (δH 4.02) to C-5 (δC 152.9) and H-11 and 12 (δH 2.81 and 3.19) to C-8 (δC 140.8) revealed the substitution of the phenolic ring (Figure 4). Specific chemical shifts of position 12 (δH 3.19, δC 46.2) as well as the suggested by HR-MS molecular formula revealed the existence of the sulfamic acid group. The sulfamic group is relatively rare in nature, however it has been found in sponges, already, for example, in araplysillin N-sulfamate isolated from A. fulva [20]. Compound 1, named aeroplysinin-3, was finally identified as (3,5-dibromo-4-(3-((ethoxycarbonyl)amino)propoxy)phenethyl)sulfamic acid, a new natural product.

3.1.2. Compound 3

Compound 3 was assigned to the molecular formula C22H21Br4N3O7, determined by HR-ESI-MS (753.8063 calcd. for [M-H]-, found 753.8059). The NMR data revealed a new compound bearing the substructure of the right half of fistularin 1 [21], including a 2-oxazolidone ring joined directly to 2,6-dibromophenol, and a N-(2-hydroxypropyl)formamide moiety. Characteristic NMR chemical shifts indicated the presence of an additional 2,6-dibromo-4-methylene-phenol group, which was methylated at position 1, as indicated by the HMBC correlation of the methoxy group (δH 3.83) to C-1 (δC 153.0) (Figure 4). Furthermore, the protons of the methylene group H-7 (δH 3.89) showed an HMBC correlation to the carbon of an oxime group, C-8 (δC 152.6), and a further correlation with the carbon of the amide group of the first substructure, C-9 (δC 163.5). The IR spectrum also showed characteristic vibrational frequencies of N-O and C=N bonds of oxime group at 939 and 1662 cm−1. Compound 3 was finally identified as (Z)-N-(3-(2,6-dibromo-4-(2-oxooxazolidin-5-yl)phenoxy)-2-hydroxypropyl)-3-(3,5-dibromo-4-methoxyphenyl)-2-(hydroxyimino)propanamide, a new natural product with the trivial name aeroplysinin-5. In order to establish the absolute configuration of compounds 3 (and also its known derivative, 14), one needs to decipher the relative stereochemistry prior to the absolute one. Compound 3 has chiral centers at C-11 and C-19. Due to the lack of NOEs between the protons of the respective carbon atoms, arising from being in two distant tails of the molecule, it was not possible to deduce any conclusion regarding their relative configuration. In an attempt to solve this issue, NMR chemical shift calculation along with computing of the DP4+ probability were applied on all the generated conformers of the possible stereoisomers of compound 3. However, the obtained results failed to establish the relative configurations of the chiral centers, which is possibly arising from (i) the lack of interaction or effect between the two chiral centers, (ii) the high flexibility of the molecule which makes it difficult for proper conformational sampling. The process was similarly applied for the other known derivate (compound 14), but no results could be concluded regarding its relative or absolute configuration, too.

3.1.3. Compound 2

Compound 2 was assigned to the molecular formula C21H23Br4N3O7S as established by a negative [M-H] signal at m/z 775.7943 (calcd. for C21H22Br4N3O7S 775.7906) in HR-ESI-MS. Characteristic NMR shifts indicated high similarity with compound 3. However, three differences could be observed; the absence of oxygen at the second carbon of the propanol group, which was indicated by lower chemical shifts of the methylene of position 11 (δH 2.06, δC 30.6), COSY correlations of the methylenes H-10/H-11/H-12 (Figure 4), and the presence of an ethylsulfamic acid group instead of the oxazolidin-2-one ring attached to carbon of position 16. The latter was suggested by COSY correlations of the protons of methylenes H-19/H-20 and key HMBC correlations of H-20/C-16 (δC 140.0) and H-9/C-15, 17 (δC 134.5). Compound 2 named aeroplysinin-4, was finally identified as (Z)-(3,5-dibromo-4-(3-(3-(3,5-dibromo-4-methoxyphenyl)-2-(hydroxyimino)propanamido)propoxy)phenethyl)sulfamic acid, a new natural product.

3.1.4. Compound 4

Compound 4 was assigned the molecular formula C12H13BrN2O5 as established by a positive ion [M+Na]+ at m/z 366.9912 (calcd. for C12H13BrN2NaO5 366.9906). NMR shifts indicated the presence of a 2-bromo-4-methylenyl phenol moiety. The methylene group of the phenol ring was connected to a 2-(hydroxyimino) acetamide group as evidenced by specific chemical shifts of position 7 (δH 3.78, δC 28.8) and HMBC correlations of H-7/C-8, 9 (δC 153.1, 165.8). The presence of an oxime group was further supported by characteristic IR absorption bands at 930 and 1659 cm−1, which were assigned for N-O and C=N bonds. Furthermore, the COSY correlation of two methylenes, H-10 and H-11, as well as the HMBC correlations of H-11 (δH 2.48) to a carboxylic group at δC 176.9 and of H-10 (δH 3.48) to C-9, 12 (δC 165.8 and 176.9), revealed the chain attached to the NH group (Figure 4). Compound 4 was not present in the sponge’s extract (Figure 5), thus it is an artifact produced during the isolation procedure. Collectively, compound 4 named nor-psammaplin M was finally identified as (Z)-3-(3-(3-bromo-4-hydroxyphenyl)-2-(hydroxyimino)propanamido)propanoic acid, a new natural product.

3.1.5. Compound 5

Compound 5 was assigned the molecular formula C12H17NO5 as established by two typical signals: [M+H]+ at m/z 256.1166 (calcd for C12H18NO5, 256.1185) and [2M+Na]+ at m/z 533.2119 (calcd. for C24H34N2O10Na 533.2111). Characteristic NMR shifts indicated high similarity with the known compound subereatensin [22], however, compound 5 showed a different optical rotation value α D 21 −21.8 (c 0.1, CHCl3) compared to the reported compound   α D 21 +25.51 (c 0.002, MeOH) [22]. Additionally, the absence of the NOESY correlation of the methine H-4 to the -OH group of position 3a indicated a different stereochemistry at these positions. Collectively, the relative configuration of compound 5 was deduced as 6aR,3aS,4R. This compound revealed a very week ECD spectrum, which could not be used for the establishment of its absolute stereochemistry. Therefore, optical rotation calculation was implemented to establish its absolute configuration. Briefly, after conformational analysis of compound 5 in chloroform by using MMFF forcefield, 13 conformers were obtained, which were subjected to geometry optimization at B3LYP/6-c31++G(d,p)/CPCM/chloroform and optical rotation at B3LYP/aug-pVTZ/CPCM/chloroform level and by considering the frequency of sodium D line for the calculation. The results demonstrated a Boltzmann-averaged specific rotation value of −59.68 (589.3 nm, sodium D line) for isomer 6aR,3aS,4R. Considering the sign of the calculated specific value, the absolute configuration of compound 5 could be deciphered accordingly. The deviation of the calculated numerical value and the experimental one could be possibly emanated from the conformational analysis and overestimation of the calculation methods. Therefore, compound 5 is a new natural stereoisomer of subereatensin [21]. It was identified as (6aR,3aS,4R)-ethyl 4-ethoxy-3a-hydroxy-2-oxo-1,2,3,3a,4,6a-hexahydrocyclopenta[b]pyrrole-6-carboxylate and given the trivial name iso-subereatensin.
Based on literature and published spectroscopic data, the other 10 isolated compounds were identified as known compounds, and specifically, they are 5-chlorocavernicolin (6) [21], 2-[(1r,4r)-3,5-dibromo-4-ethoxy-1-hydroxy-4-methoxycyclohexa-2,5-dien-1-yl]acetamide (7) [11], 2-[(1s,4s)-3,5-dibromo-4-ethoxy-1-hydroxy-4-methoxycyclohexa-2,5-dien-1-yl]acetamide (8) [11], N,N,N-trimethyl-3,5-dibromotyramine (9) [23], 2-((1s,4s)-3,5-dibromo-4-ethoxy-1-hydroxy-4-methoxycyclohexa-2,5-dien-1-yl)acetic acid (10) [11], 5-[3,5-dibromo-4-(2-oxo-oxazolidin-5-ylmethoxy)-phenyl]-oxazolidin-2-one (11) [24], aeroplysinin-2 (12) [11], 2,6-dibromo-4-hydroxy-4-methoxycarbonylmethylcyclohexa-2,5-dien-1-one (13) [11], fistularin-1 (14) [21], and aplysterol (15) [25]. Compounds 6 and 9 are reported in Aplysina aerophoba for the first time.
A comparison to published data also allowed the identification of the compounds isolated from Spongia sp. They were identified as 2-hexaprenyl-1-4-hydroquinone (16) [26], 2-heptaprenyl-1-4-hydroquinone (17) [27], 2-octaprenyl-1-4-hydroquinone (18) [28], 2-[24 hydroxy] octaprenyl-1-4-hydroquinone (19) [28], and 2-hexaprenylmethyl-2-methylchromen-6-ol (20) [26]. These compounds are already known constituents of Spongia sp.

3.2. Cytotoxic Properties

Cytotoxic properties of almost all compounds isolated from Spongia sp. and Aplysina aerophoba to T24 bladder and AGS stomach tumor cells was determined using the common MTT assay by Mossman et al. Table 4 displays the respective, expressed as IC50 values, as the concentration (µM) of each test compound causing 50% effect in the respective assays. For less active compounds, the % viability at 100 µM concentration is given. Additionally, the two crude extracts were investigated (500 µg/mL). Spongia sp. did not show any toxicity effects whereas the Aplysina aerophoba extract showed a cell viability of only 25.3% after a 24 h treatment. Compounds 13, 16, 17, and 19 showed strong cytotoxic effects towards AGS cells with compound 17 being the most active compound with an IC50 value of 0.99 µM. Compounds 12 and 18 indicated moderate cytotoxic activity towards AGS cells. However only one compound showed significant toxicity towards T24 cells, i.e., for compound 13, an IC50 of 12.42 µM was determined. Compounds 7, 14, 17, and 18 exerted moderate effects. In addition, the same substances (compounds 520) were investigated for a potential apoptotic activity towards the neuroblastoma cell line SH-SY5Y in three concentrations (25 to 100 µM) using a fluorescence activated cell sorting readout (FACS). Significant cytotoxicity was observed for compound 13 in this assay (cell viability was measured below 20% for all concentrations) (Figure 6). Subsequently, the five most promising compounds according to the initial screening in the above-mentioned cell lines (compounds 13, 16, 17, 18, and 19) were chosen for further investigations. They were additionally investigated in three colon cancer cell lines (DLD-1, SW-480 and Lovo) for potential cytotoxic effects. Again, compound 13 revealed activity in all cell lines (Figure 7), and decreased cell viability. Compound 18 and 19 showed moderate activity in the highest concentration (100 µM), with a remaining cell viability of around 60% for SW-480 and Lovo cell lines. Compound 13 was found to be cytotoxic against these colon carcinoma cell lines even at lower concentrations ranging from 1–25 µM (Figure 8A). For SH-SY5Y cells, compound 13 was titrated from 25 to 0.63 µM and the IC50 value was determined with 1.78 [CI 95 1.34–2.38] (Figure 8B). In comparison, etoposide, a standard cytotoxic compound used in clinical practice for neuroblastoma treatment, killed 50% of SH-SY5Y cells at a concentration of 340 nM (data not shown). We also utilized human primary fibroblasts for testing in order to delineate a general cytotoxicity of this compound. These non-cancerous cells were significantly less sensitive to 13 supporting a potential anti-cancer effect of this compound with a certain degree of selectivity (Figure 8C). (FACS Plot figures for 13 are shown in the Supplementary Materials).
Encouraged by the positive results, compound 13 was further investigated in a hematological cancer model, i.e., multiple myeloma. Again, cytotoxicity was found at low concentrations (concentration range 1 to 25 µM), showing a decreased viability of NCI-H929, OPM-2, and U266 cell lines of 4%, 54%, and 34%, respectively, after 24 h incubation time at a concentration of 5 µM. In non-cancerous blood cells, i.e., peripheral blood mononuclear cells (PBMCs) of healthy donors, compound 13 induced significantly lower levels of cell death, again suggesting that cancer cells might be more susceptible (PBMC 1-3, Figure 9).

4. Discussion and Conclusions

Marine sponges are gaining the attention of the scientific community because of their unique secondary metabolites with a diversity of biological activities. Some of them are being tested in clinical trials against various diseases, mainly with the focus on anti-cancer drugs [29,30].
A. aerophoba and Spongia sp. were selected for this study because initial screening experiments already indicated that they contain interesting and possibly new metabolites.
Apart from the commonly studied bromotyrosines from A. aerophoba which are compounds with interesting antibacterial and antitumor activities [13,31], attention has been focused on chitin which has further applications, e.g., as scaffold in tissue engineering, pharmacological applications, and regenerative medicine [13,32]. Furthermore, the farming of Spongia sp. for commercial purposes as bath sponges has been going on for more than a century [33]. These sponges have a high commercial value and they are successfully employed in experimental aquaculture experiences [34]. The enormous filtering capacity of sponges has led to the suggestion that they be farmed for bioremediation purposes, e.g., to reduce the high bacterial loads resulting from sewage discharges and areas subjected to aquaculture activities [33,34]. Additionally, Spongia sp. have recently been studied as an alternative source of collagen which is found in their extracellular skeletal matrix [17].
In our study, twenty metabolites, mainly halogenated compounds and prenylated hydroquinones, were isolated using a purification protocol including liquid–liquid extraction, fractionation on silica gel column and Sephadex LH-20 columns, and a final clean-up step with semipreparative HPLC or flash chromatography. Four brominated metabolites and an isomer of subereatensin isolated from A. aerophoba have been identified as novel natural products and two additional brominated compounds were reported from A. aerophoba for the first time.
It is already known that some of the isolated compounds exert cytotoxic and antimicrobial properties, e.g., aeroplysinin-2 is known to possess moderate antibacterial activity against the Gram-positive bacteria Staphylococcus lentus, Propionibacterium acnes, and Bacillus subtilis [31], the mouse fibroblasts NIH-3T3, the human hepatocellular carcinoma cell line HepG2, and the human colon adenocarcinoma cell line HT-29 [31]. Additionally, aeroplysinin-1 shows cytotoxic effects against human neuroblastoma (SH-SY5Y) [13]. Further, 2-Octaprenyl-1-4-hydroquinone has been reported to inhibit marine bacterial strains (both Gram-positive and -negative bacteria) [35], while 2-[24 hydroxy] octaprenyl-1-4-hydroquinone revealed moderate cytotoxic activity against C98 cells and good antibacterial activity against S. aureus and E. cloacae [36]. In addition, 2,6-Dbromo-4-hydroxy-4-methoxycarbonylmethylcyclohexa-2,5-dien-1-one, the most cytotoxic compound in our study, had only been investigated for its anti-microbial properties showing weak effects [37].
When evaluating the possible cytotoxic effects of the isolated metabolites, only a few of them showed activities in the investigated cell lines. For AGS stomach tumor cells, compound 17 was found to be the most toxic, while compound 13 showed the highest cytotoxic activity in all other cell lines tested. Therefore, this compound seems to be interesting for further studies, especially because our data indicated selectivity towards cancer cells.
Sponges often contain diverse and abundant microbial communities in their tissues and in many cases, the associated bacterial communities account for over 40% of the biomass of their hosts [30,34]. In the past few years, more and more evidence has accumulated in which a part of the isolated metabolites are not produced by the sponges themselves, but instead, they are products of the metabolic activities of bacteria living in the sponge tissue [34]. This also could be the case for the compounds isolated within this study. Our results suggest that A. aerophoba and Spongia sp., among other demosponges which currently attract increased interest in the scientific community, are a rich source of interesting compounds. This refers to unique chemical structures as well as promising bioactivities.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/biom11050723/s1, HRMS, 1D and 2D NMR, IR spectra of all the new compounds 15, HPLC separation of the Spongia spp. extract, conformers of compound 5 and their Boltzmann averaging along with their calculated specific rotation values.

Author Contributions

Investigation, M.O., A.H. (Anja Hartmann), N.M., M.A., K.J., M.K. and H.N.N.; writing—original draft preparation, M.O., A.H. (Anja Hartmann), M.A. and K.J.; writing—review, editing and supervision, A.H. (Anja Hartmann), A.H. (Andreas Hensel), M.G., M.K. and R.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Austrian Science Fund (FWF), project No. P29671. “Open Access Funding by the Austrian Science Fund (FWF)”.

Institutional Review Board Statement

PBMCs from healthy donors were used after obtaining informed consent at the University Hospital Salzburg (ethics committee approval 415-E/1287/6-2011). Approved by the ethics committee of the University Hospital Salzburg, number 415-E/1287/6-2011.

Informed Consent Statement

Informed consent was obtained from all subjects involved in the study.

Data Availability Statement

Not applicable.

Acknowledgments

The authors gratefully thank Dimitris Poursanidis (Postdoc fellow, IACM-FORTH, Foundation for Research and Technology—Heraklion, Crete, Greece—CEO of terraSolutions marine environment research) for the collection and identification of the sponges and Barbara Matuszczak (Institute of Pharmacy, Pharmaceutical Chemistry, University of Innsbruck, Austria) for recording the IR spectra.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. John Faulkner, D. Marine natural products. Nat. Prod. Rep. 2000, 17, 7–55. [Google Scholar] [CrossRef] [PubMed]
  2. Schuster, A.; Vargas, S.; Knapp, I.S.; Pomponi, S.A.; Toonen, R.J.; Erpenbeck, D.; Wörheide, G. Divergence times in demosponges (Porifera): First insights from new mitogenomes and the inclusion of fossils in a birth-death clock model. BMC Evol. Biol. 2018, 18, 114. [Google Scholar] [CrossRef] [PubMed]
  3. Proksch, P.; Ebel, R.; Edrada, R.A.; Wray, V.; Steube, K. Bioactive Natural Products from Marine Invertebrates and Associated Fungi. In Sponges (Porifera); Progress in Molecular and Subcellular Biology; Springer: Berlin/Heidelberg, Germany, 2003; Volume 37, pp. 117–142. [Google Scholar]
  4. Mehbub, M.F.; Perkins, M.V.; Zhang, W.; Franco, C.M.M. New marine natural products from sponges (Porifera) of the order Dictyoceratida (2001 to 2012); a promising source for drug discovery, exploration and future prospects. Biotechnol. Adv. 2016, 34, 473–491. [Google Scholar] [CrossRef] [PubMed]
  5. Sacristán-Soriano, O.; Banaigs, B.; Becerro, M.A. Temporal trends in the secondary metabolite production of the sponge Aplysina aerophoba. Mar. Drugs 2012, 10, 677–693. [Google Scholar] [CrossRef] [Green Version]
  6. Zucht, W.; Sidri, M.; Brümmer, F.; Jaklin, A.; Hamer, B. Ecology and distribution of the sponge Aplysina aerophoba (Porifera, Demospongiae) in the Limski kanal (northern Adriatic Sea, Croatia). Fresenius Environ. Bull. 2008, 17, 890–901. [Google Scholar]
  7. Cimino, G.; De Rosa, S.; De Stefano, S.; Spinella, A.; Sodano, G. The zoochrome of the sponge verongia aerophoba (“Uranidine”). Tetrahedron Lett. 1984, 25, 2925–2928. [Google Scholar] [CrossRef]
  8. Wehrl, M.; Steinert, M.; Hentschel, U. Bacterial Uptake by the Marine Sponge Aplysina aerophoba. Microb. Ecol. 2007, 53, 355–365. [Google Scholar] [CrossRef] [Green Version]
  9. Turon, X.; Becerro, M.A.; Uriz, M.J. Distribution of brominated compounds within the sponge Aplysina aerophoba: Coupling of X-ray microanalysis with cryofixation techniques. Cell Tissue Res. 2000, 301, 311–322. [Google Scholar] [CrossRef]
  10. Ebel, R.; Brenzinger, M.; Kunze, A.; Gross, H.J.; Proksch, P. Wound Activation of Protoxins in Marine Sponge Aplysina aerophoba. J. Chem. Ecol. 1997, 23, 1451–1462. [Google Scholar] [CrossRef]
  11. Santalova, E.A.; Denisenko, V.A.; Glazunov, V.P.; Kalinovskii, A.I.; Anastyuk, S.D.; Stonik, V.A. Dibromotyrosine derivatives from the ethanol extract of the marine sponge Aplysina sp.: Structures, transformations, and origin. Russ. Chem. Bull. 2011, 60, 570. [Google Scholar] [CrossRef]
  12. Koulman, A.; Proksch, P.; Ebel, R.; Beekman, A.C.; van Uden, W.; Konings, A.W.T.; Pedersen, J.A.; Pras, N.; Woerdenbag, H.J. Cytoxicity and Mode of Action of Aeroplysinin-1 and a Related Dienone from the Sponge Aplysina aerophoba. J. Nat. Prod. 1996, 59, 591–594. [Google Scholar] [CrossRef]
  13. Binnewerg, B.; Schubert, M.; Voronkina, A.; Muzychka, L.; Wysokowski, M.; Petrenko, I.; Djurovic, M.; Kovalchuk, V.; Tsurkan, M.; Martinović, R.; et al. Marine biomaterials: Biomimetic and pharmacological potential of cultivated Aplysina aerophoba marine demosponge. Mater. Sci. Eng. C 2019, 109, 110566. [Google Scholar] [CrossRef]
  14. Hausmann, R.; Vitello, M.; Leitermann, F.; Syldatk, C. Advances in the production of sponge biomass Aplysina aerophoba-A model sponge for ex situ sponge biomass production. J. Biotechnol. 2006, 124, 117–127. [Google Scholar] [CrossRef]
  15. Máximo, P.; Ferreira, L.M.; Branco, P.; Lima, P.; Lourenço, A. The Role of Spongia sp. in the Discovery of Marine Lead Compounds. Mar. Drugs 2016, 14, 139. [Google Scholar] [CrossRef] [Green Version]
  16. de C. Cook, S.; Bergquist, P.R. New species of Spongia (Porifera: Demospongiae: Dictyoceratida) from New Zealand, and a proposed subgeneric structure. N. Z. J. Mar. Freshwater Res. 2001, 35, 33–58. [Google Scholar] [CrossRef]
  17. Ghosh, A.; Grosvenor, A.J.; Dyer, J.M. Marine Spongia collagens: Protein characterization and evaluation of hydrogel films. J. Appl. Polym. Sci. 2019, 136, 47996. [Google Scholar] [CrossRef]
  18. Sunassee, S.N.; Davies-Coleman, M.T. Cytotoxic and antioxidant marine prenylated quinones and hydroquinones. Nat. Prod. Rep. 2012, 29, 513–535. [Google Scholar] [CrossRef]
  19. Willer, A.; Jöhrer, K.; Greil, R.; Zidorn, C.; Cicek, S. Cytotoxic Properties of Damiana (Turnera diffusa) Extracts and Constituents and A validated quantitative UHPLC-DAD Assay. Molecules 2019, 24, 855. [Google Scholar] [CrossRef] [Green Version]
  20. Rogers, E.W.; Molinski, T.F. Highly Polar Spiroisoxazolines from the Sponge Aplysina fulva. J. Nat. Prod. 2007, 70, 1191–1194. [Google Scholar] [CrossRef] [Green Version]
  21. Gopichand, Y.; Schmitz, F.J. Marine natural products: Fistularin-1, -2 and -3 from the sponge Aplysina fistularis forma fulva. Tetrahedron Lett. 1979, 20, 3921–3924. [Google Scholar] [CrossRef]
  22. Kijjoa, A.; Watanadilok, R.; Sonchaeng, P.; Sawangwong, P.; Pedro, M.; Nascimento, M.; Silva, A.; Eaton, G.; Herz, W. Further Halotyrosine Derivatives from the Marine Sponge Suberea aff. praetensa. Z. Naturforsch. 2002, 57, 732–738. [Google Scholar] [CrossRef] [Green Version]
  23. Peng, J.; Li, J.; Hamann, M.T. The marine bromotyrosine derivatives. Alkaloids Chem. Biol. 2005, 61, 59–262. [Google Scholar] [CrossRef]
  24. Norte, M.; Rodriguez, M.L.; Fernández, J.J.; Eguren, L.; Estrada, D.M. Aplysinadiene and (r,r) 5 [3,5-dibromo-4-[(2-oxo-5-oxazolidinyl)] methoxyphenyl]-2-oxazolidinone, two novel metabolites from aplysina aerophoba synthes. Tetrahedron 1988, 44, 4973–4980. [Google Scholar] [CrossRef]
  25. Rahelivao, M.; Lübken, T.; Gruner, M.; Kataeva, O.; Andriamanantoanina, H.; Checinski, M.; Bauer, I.; Knölker, H.-J. Isolation and structure elucidation of natural products of three soft corals and a sponge from the coast of Madagascar. Org. Biomol. Chem. 2017, 15, 2593. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Lee, H.; Lee, Y.; Lee, J.; Shin, H.J.; Lee, J.; Kim, K.; Yoon, W.; Heo, S.-J.; Kim, H. Identification of New Polyprenyl Hydroquinone Derivatives from Tropical Marine Sponge Ircinia sp. Heterocycles 2012, 85, 1437. [Google Scholar] [CrossRef]
  27. Gil, B.; Jesús^Sanz, M.; Terencio, M.C.; De Giulio, A.; De Rosa, S.; Alcaraz, M.J.; Paya, M. Effects of marine 2-polyprenyl-1,4-hydroquinones on phospholipase A2 activity and some inflammatory responses. Eur. J. Pharmacol. 1995, 285, 281–288. [Google Scholar] [CrossRef]
  28. Tziveleka, L.-A.; Kourounakis, A.P.; Kourounakis, P.N.; Roussis, V.; Vagias, C. Antioxidant Potential of Natural and Synthesised Polyprenylated Hydroquinones. Bioorg. Med. Chem 2002, 10, 935–939. [Google Scholar] [CrossRef]
  29. Singh, A.; Thakur, N. Significance of investigating allelopathic interactions of marine organisms in the discovery and development of cytotoxic compounds. Chem. Biol. Interact 2015, 243. [Google Scholar] [CrossRef]
  30. Thoms, C.; Schupp, P. Biotechnological Potential of Marine Sponges and their Associated Bacteria as Producers of New Pharmaceuticals (Part II). J. Int. Biotechnol. Law 2005, 2, 257. [Google Scholar] [CrossRef] [Green Version]
  31. Shaker, K.H.; Zinecker, H.; Ghani, M.A.; Imhoff, J.F.; Schneider, B. Bioactive Metabolites from the Sponge Suberea sp. Chem. Biodivers. 2010, 7, 2880–2887. [Google Scholar] [CrossRef]
  32. Martinez Gil, A.; Alonso Gordillo, D.; Dorronsoro Diaz, I.; Garcia Palomero, E.; De Austria De Luque, C.; Usan Egea, P.; Del Monte Millan, M.; Median Padilla, M. Marine Compounds with Calcium Channel Blocking Properties for the Treatment of Cognitive or Neurodegenerative Diseases 2004. U.S. Patent 11/570,753, 17 June 2005. [Google Scholar]
  33. Stabili, L.; Licciano, M.; Giangrande, A.; Longo, C.; Mercurio, M.; Marzano, C.N.; Corriero, G. Filtering Activity of Spongia officinalis Var. adriatica (Schmidt) (Porifera, Demospongiae) on Bacterioplankton: Implications for Bioremediation of Polluted Seawater. Water Res. 2006, 40, 3083–3090. [Google Scholar] [CrossRef]
  34. Taylor, M.W.; Radax, R.; Steger, D.; Wagner, M. Sponge-associated microorganisms: Evolution, ecology, and biotechnological potential. Microbiol. Mol. Biol. Rev. 2007, 71, 295–347. [Google Scholar] [CrossRef] [Green Version]
  35. Tsoukatou, M.; Hellio, C.; Vagias, C.; Harvala, C.; Roussis, V. Chemical Defense and Antifouling Activity of Three Mediterranean Sponges of the Genus Ircinia. Z. Naturforsch. 2002, 57, 161–171. [Google Scholar] [CrossRef] [Green Version]
  36. Mihopoulos, N.; Vagias, C.; Chinou, I.; Roussakis, C.; Scoullos, M.; Harvala, C.; Roussis, V. Antibacterial and Cytotoxic Natural and Synthesized Hydroquinones from Sponge Ircinia spinosula. Z. Naturforsch. 1999, 54, 417–423. [Google Scholar] [CrossRef]
  37. Shestak, O.P.; Novikov, V.L.; Ivanova, E.P.; Gorshkova, N.M. Synthesis and Antimicrobial Activity of [3,5-Dibromo(dichloro)-1-hydroxy-4-oxocyclohexa-2,5-dien-1-yl]acetic Acids and Their Derivatives. Pharm. Chem. J. 2001, 35, 366–369. [Google Scholar] [CrossRef]
Figure 1. The chemical structures of compounds 1 to 20 as elucidated by NMR and MS; compounds 15 are novel natural products isolated from A. aerophoba, compounds 615 were isolated from A. aerophoba and they have been previously described in the literature, compounds 1620 were isolated from Spongia sp. and they are known compounds.
Figure 1. The chemical structures of compounds 1 to 20 as elucidated by NMR and MS; compounds 15 are novel natural products isolated from A. aerophoba, compounds 615 were isolated from A. aerophoba and they have been previously described in the literature, compounds 1620 were isolated from Spongia sp. and they are known compounds.
Biomolecules 11 00723 g001
Figure 2. The isolation scheme for the A. aerophoba extract.
Figure 2. The isolation scheme for the A. aerophoba extract.
Biomolecules 11 00723 g002
Figure 3. The isolation scheme for the Spongia sp. extract.
Figure 3. The isolation scheme for the Spongia sp. extract.
Biomolecules 11 00723 g003
Figure 4. Key HMBC (1H → 13C) and 1H-1H COSY correlations of compounds 14.
Figure 4. Key HMBC (1H → 13C) and 1H-1H COSY correlations of compounds 14.
Biomolecules 11 00723 g004
Figure 5. HPLC-UV-MS separation of the A. aerophoba extract. Peak assignment is according to Figure 1. Stationary phase: Synergi Polar-RP 80A (250 × 4, 6 mm; 4 μm) from Phenomenex (Torrance, CA, USA); mobile phase: 0.1% (v/v) FA in water (A) and CH3CN (B); Gradient: 0 min: 2% B, 10 min: 15% B, 18 min: 50% B, 30–35 min: 98% B, 35.1–45 min: 2% B; λ = 210, 254, 280, 310, and 350 nm; flow rate = 0.7 mL/min; T = 22 °C.
Figure 5. HPLC-UV-MS separation of the A. aerophoba extract. Peak assignment is according to Figure 1. Stationary phase: Synergi Polar-RP 80A (250 × 4, 6 mm; 4 μm) from Phenomenex (Torrance, CA, USA); mobile phase: 0.1% (v/v) FA in water (A) and CH3CN (B); Gradient: 0 min: 2% B, 10 min: 15% B, 18 min: 50% B, 30–35 min: 98% B, 35.1–45 min: 2% B; λ = 210, 254, 280, 310, and 350 nm; flow rate = 0.7 mL/min; T = 22 °C.
Biomolecules 11 00723 g005
Figure 6. Viability of neuroblastoma cell line SH-SY5Y after treatment with different concentrations (100/50/25 µM) of the indicated compounds for 24 h.
Figure 6. Viability of neuroblastoma cell line SH-SY5Y after treatment with different concentrations (100/50/25 µM) of the indicated compounds for 24 h.
Biomolecules 11 00723 g006
Figure 7. Viability of human colon adenocarcinoma cell lines DLD-1, SW-480, and Lovo treated with the respective compounds for 24 h. Viability was measured by flow cytometry (AnnV/PI staining) and % of cells alive compared to control (untreated) are shown +/− SD.
Figure 7. Viability of human colon adenocarcinoma cell lines DLD-1, SW-480, and Lovo treated with the respective compounds for 24 h. Viability was measured by flow cytometry (AnnV/PI staining) and % of cells alive compared to control (untreated) are shown +/− SD.
Biomolecules 11 00723 g007
Figure 8. Influence of 13 at different concentrations on (A) different colon carcinoma cell lines, on (B) SH-SY5Y neuroblastoma cell line, and on (C) primary human fibroblasts as non-cancerous control cells. Incubation time was 24 h followed by flow cytometric analysis (AnnV/PI staining). Percentage of cells alive was calculated compared to control (no treatment).
Figure 8. Influence of 13 at different concentrations on (A) different colon carcinoma cell lines, on (B) SH-SY5Y neuroblastoma cell line, and on (C) primary human fibroblasts as non-cancerous control cells. Incubation time was 24 h followed by flow cytometric analysis (AnnV/PI staining). Percentage of cells alive was calculated compared to control (no treatment).
Biomolecules 11 00723 g008
Figure 9. Activity of compound 13 on (A) hematological cancer cells (multiple myeloma cell lines) and on (B) primary peripheral blood mononuclear cells of three healthy donors (PBMC 1–3). Incubation time was 24 h, followed by flow cytometric analysis.
Figure 9. Activity of compound 13 on (A) hematological cancer cells (multiple myeloma cell lines) and on (B) primary peripheral blood mononuclear cells of three healthy donors (PBMC 1–3). Incubation time was 24 h, followed by flow cytometric analysis.
Biomolecules 11 00723 g009
Table 1. 1H NMR Data (δ in ppm, J in Hz) of compounds 1 (400 MHz) and 4 (600 MHz).
Table 1. 1H NMR Data (δ in ppm, J in Hz) of compounds 1 (400 MHz) and 4 (600 MHz).
No.Compound 1 (400 MHz)Compound 4 (600 MHz)
δH (J in Hz)δCδH (J in Hz)δC
1 159.3 153.9
23.38 t (7.2)39.26.76 d (8.4)117.1
32.02 m31.47.07 dd (1.8/8.4)130.5
44.02 t (6.4)72.2 130.8
5 152.97.37 d (1.8)134.7
6 119.0 110.6
77.49 s134.53.78 s28.8
8 140.8 153.1
97.49 s134.5 165.8
10 119.03.48 t (6.6)36.7
112.80 t (7.6)35.92.48 t (6.6)35.6
123.19 t (7.6)46.2 176.9 *
134.07 q (7.2)61.8
141.23 t (7.2)15.2
Measured in MeOD, * δC established from the HMBC spectrum.
Table 2. 1H NMR Data (δ in ppm, J in Hz) of compounds 2 (400 MHz) and 3 (600 MHz).
Table 2. 1H NMR Data (δ in ppm, J in Hz) of compounds 2 (400 MHz) and 3 (600 MHz).
No.Compound 2 aCompound 3 b
δH (J in Hz)δCδH (J in Hz)δC
1 154.0 153.0
2, 6 118.8–119.2 ** 118.1
3, 57.49 s134.77.48 brs133.7
4 137.5 136.4 *
73.85 s28.93.89 s39.0
8 152.2 152.6 *
9 165.5 163.5
103.55 t (6.8)38.03.54 m
3.76 m
42.3
112.06 m30.64.21 m69.9
124.01 t (6.0)72.44.00 dd (4.8,9.0)
4.09 dd (4.8,9.0)
74.6
13 153.0 152.9
14, 18 118.8, 119.2 * 118.9
15, 177.48 s134.57.53 brs130.2
16 140.0 137.8 *
192.84 t (7.6)35.05.54 t (8.0)75.8
203.25 t (7.6)46.43.50 t (8.0)
4.00 t (8.0)
48.1
213.81 s61.2 158.5
22 3.83 s61.0
NH 7.22 m
a Measured in MeOD, b Measured in chloroform-d * δC established from the HMBC spectrum, ** Overlapping signals, Measured in MeOD.
Table 3. 1H NMR Data (δ in ppm, J in Hz) of compound 5 (600 MHz).
Table 3. 1H NMR Data (δ in ppm, J in Hz) of compound 5 (600 MHz).
No.δH (J in Hz)δCNOESYCOSYHMBC
2 174.3
3′2.70 d (17.4)43.1 H-3′′C-2,C-3a,C-4,C-6a
3′′2.48 d (17.4)43.1H-4H-3′C-2,C-4,C-6a
3a 81.2
44.20 m84.4 H-3′′, H-5, H-9, H-10H-5C-3,5,6,10
56.83 d (1.8)140.1 H-4C-3a,C-4,C-6a,C-7
6 139.5
6a4.48 brs68.6 OH C-2,C-3a,C-5
7 163.7
8 61.6 H-9H-9C-7,C-9
91.31 t (7.2)14.4 H-8, H-4H-8C-8
103.73 m67.5H-5, H-11H-10C-4,C-11
111.28 t (7.2)15.6 H-10aH-11C-10
OH6.06 brs H-6a C-3,3a,6a
Measured in chloroform-d.
Table 4. Summary of all compounds and their IC50 values of cell viability/% viability at 100 µM. (95% confidence intervals in parentheses).
Table 4. Summary of all compounds and their IC50 values of cell viability/% viability at 100 µM. (95% confidence intervals in parentheses).
CompoundAGS T24
IC50 µM (CI 95 ±)IC50 µM (CI 95 ±)
5Not activeNot active
6Not activeNot active
7Not active54.57% at 100 µM
(21.88–107.50)
8Not activeNot active
9Not activeNot active
10Not activeNot active
11*--
1254.57% at 100 µM
(17.87–86.17)
Not active
1310.14 (8.03 to 12.82)12.42 (10.65 to 14.48)
14Not active56.20% at 100 µM
(16.81–70.79)
15*--
165.33 (4.03 to 7.06)Not active
170.994 (0.61 to 1.61)66.63% at 100 µM
(45.10–88.16)
1833.41% at 100 µM
(13.02–53.82)
34.38% at 100 µM
(7.454–76.22)
198.09 (5.68 to 11.52)Not active
20Not activeNot active
Spongia sp. extract 500 µg /mLNot activeNot active
A. aerophoba extract 500 µg/mL25.26% Not active
* not tested because of poor solubility.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Orfanoudaki, M.; Hartmann, A.; Alilou, M.; Mehic, N.; Kwiatkowski, M.; Jöhrer, K.; Nguyen Ngoc, H.; Hensel, A.; Greil, R.; Ganzera, M. Cytotoxic Compounds of Two Demosponges (Aplysina aerophoba and Spongia sp.) from the Aegean Sea. Biomolecules 2021, 11, 723. https://doi.org/10.3390/biom11050723

AMA Style

Orfanoudaki M, Hartmann A, Alilou M, Mehic N, Kwiatkowski M, Jöhrer K, Nguyen Ngoc H, Hensel A, Greil R, Ganzera M. Cytotoxic Compounds of Two Demosponges (Aplysina aerophoba and Spongia sp.) from the Aegean Sea. Biomolecules. 2021; 11(5):723. https://doi.org/10.3390/biom11050723

Chicago/Turabian Style

Orfanoudaki, Maria, Anja Hartmann, Mostafa Alilou, Naida Mehic, Marcel Kwiatkowski, Karin Jöhrer, Hieu Nguyen Ngoc, Andreas Hensel, Richard Greil, and Markus Ganzera. 2021. "Cytotoxic Compounds of Two Demosponges (Aplysina aerophoba and Spongia sp.) from the Aegean Sea" Biomolecules 11, no. 5: 723. https://doi.org/10.3390/biom11050723

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop