Next Article in Journal
Development of Novel Micellar-Enhanced High-Throughput Microwell Spectrofluorimetric Method for Quantification of Lorlatinib: Application to In Vitro Drug Release and Analysis of Urine Samples
Previous Article in Journal
Special Issue “Drug Candidates for the Treatment of Infectious Diseases”
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Nor-24-homoscalaranes, Neutrophilic Inflammatory Mediators from the Marine Sponge Lendenfeldia sp.

1
Graduate Institute of Pharmacognosy, College of Pharmacy, Taipei Medical University, Taipei 110301, Taiwan
2
Doctoral Degree Program in Marine Biotechnology, National Sun Yat-sen University, Kaohsiung 804201, Taiwan
3
National Museum of Marine Biology and Aquarium, Pingtung 944401, Taiwan
4
PhD Program in Clinical Drug Development of Herbal Medicine, College of Pharmacy, Taipei Medical University, Taipei 110301, Taiwan
5
Department of Pharmacognosy, Faculty of Pharmacy, Ain-Shams University, Organization of African Unity Street, Abassia, Cairo 11566, Egypt
6
Research Center for Chinese Herbal Medicine, College of Human Ecology, Chang Gung University of Science and Technology, Taoyuan 333324, Taiwan
7
Graduate Institute of Health Industry Technology, College of Human Ecology, Chang Gung University of Science and Technology, Taoyuan 333324, Taiwan
8
Graduate Institute of Natural Products, College of Medicine, Chang Gung University, Taoyuan 333323, Taiwan
9
Department of Chemical Engineering, Ming Chi University of Technology, New Taipei City 243303, Taiwan
10
Department of Anesthesiology, Chang Gung Memorial Hospital, Taoyuan 333423, Taiwan
11
Department of Marine Biotechnology and Resources, National Sun Yat-sen University, Kaohsiung 804201, Taiwan
12
Center for Reproductive Medicine and Sciences, Taipei Medical University Hospital, Taipei 110301, Taiwan
13
Traditional Herbal Medicine Research Center, Taipei Medical University Hospital, Taipei 110301, Taiwan
14
Graduate Institute of Natural Products, Kaohsiung Medical University, Kaohsiung 807378, Taiwan
15
Chinese Medicine Research and Development Center, China Medical University Hospital, Taichung 404394, Taiwan
16
PhD Program in Pharmaceutical Biotechnology, Fu Jen Catholic University, New Taipei City 242062, Taiwan
*
Authors to whom correspondence should be addressed.
Pharmaceuticals 2023, 16(9), 1258; https://doi.org/10.3390/ph16091258
Submission received: 20 July 2023 / Revised: 28 August 2023 / Accepted: 4 September 2023 / Published: 6 September 2023
(This article belongs to the Section Natural Products)

Abstract

:
The marine sponge Lendenfeldia sp., collected from the Southern waters of Taiwan, was subjected to chemical composition screening, resulting in the isolation of four new 24-homoscalarane compounds, namely lendenfeldaranes R–U (14). The structures and relative stereochemistry of the new metabolites 14 were assigned based on NMR studies. The absolute configurations of compounds 14 were determined by comparing the calculated and experimental values of specific optical rotation. The antioxidant and anti-inflammatory activities of the isolated compounds were assayed using superoxide anion generation and elastase release assays. These assays are used to determine neutrophilic inflammatory responses of respiratory burst and degranulation. Compounds 2 and 4 inhibited superoxide anion generation by human neutrophils in response to formyl-L-methionyl-L-leucyl-L-phenylalanine/cytochalasin B (fMLP/CB) with IC50: 3.98–4.46 μM. Compounds 2 and 4 inhibited fMLP/CB-induced elastase release, with IC50 values ranging from 4.73 to 5.24 μM. These findings suggested that these new 24-homoscalarane compounds possess unique structures and potential anti-inflammatory activity.

Graphical Abstract

1. Introduction

Marine sponges belonging to the genus Lendenfeldia (phylum Porifera, class Demospongia, subclass Keratosa, order Dictyoceratida, family Thorectidae) are widely distributed in the flat reefs of the Asia–Pacific region. These sponges can also be found in aquarium tanks and are considered pests as they can proliferate in aquacultures even under standard conditions [1]. Previous research on Lendenfeldia sponges revealed a rich diversity of secondary metabolites, with sesterterpenoids being particularly prominent [2,3,4,5,6,7,8,9]. These sesterterpenoids exhibit a broad spectrum of biological activities, including cytotoxicity [10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31], anti-inflammatory properties [5,7,18,32,33,34,35,36], anti-HIV effects [3,37], anti-microbial [15,16,22,38,39,40,41] and anti-neurofibroma activity [42]. Anti-neutrophil inflammatory agents refer to substances that aim to reduce or inhibit the inflammatory response mediated by neutrophils. Neutrophils are a type of white blood cell that are essential to the immune response. They play a crucial role in the initial response to infection and tissue injury [43]. While neutrophils are important for combating pathogens and promoting tissue repair, their excessive or prolonged activation can lead to harmful inflammation and tissue damage. Therefore, targeting neutrophil-mediated inflammation is an area of interest in various inflammatory conditions, including autoimmune diseases, acute respiratory distress syndrome (ARDS), and inflammatory bowel disease (IBD) [44]. In our two previous reports, we isolated a series of scalarane-type sesterterpenoids from the marine sponge Lendenfeldia sp. and tested their anti-inflammatory activity against superoxide anion generation and elastase release, which represented the neutrophilic inflammatory responses of respiratory burst and degranulation, respectively. The results indicated that the potent activity of this class of compounds could lead to their further development as anti-neutrophilic agents [5,7]. In our ongoing investigation, aiming to discover new metabolites from Lendenfeldia sp., we identified four novel 24-homoscalarane-type sesterterpenoids named lendenfeldaranes R–U (14) (Figure 1). The determination of the structures of compounds 14 involved a thorough analysis of their Infrared (IR), Specific Optical Rotation (SOR), Mass (MS), and Nuclear Magnetic Resonance (NMR) spectra. Moreover, their NMR data were compared with those of known compounds with structural similarities. To establish the absolute configuration of compounds 14, we compared the experimental optical rotation with the calculated optical rotation spectra. Additionally, we evaluated the anti-inflammatory activity of these compounds by assessing their ability to inhibit superoxide anion generation and elastase release in N-formyl-methionyl-leucyl phenylalanine/cytochalasin B (fMLF/CB)-induced human neutrophils.

2. Results and Discussion

Compound 1 was obtained as a white powder, and its molecular formula was determined as C27H42O5 based on the presence of a sodium adduct at m/z 469.29256 in HRESIMS (calculated for C27H42O5 + Na, 469.29245) and 13C NMR data (Table 1). The IR spectrum of 1 exhibited characteristic peaks indicating the presence of ketone carbonyl (νmax 1703 cm−1), ester carbonyl (νmax 1736 cm−1), and hydroxy (νmax 3429 cm−1) functional groups. The 1H NMR spectroscopic data of 1 (Table 1) revealed six methyl groups at δH 0.85, 0.88, 1.13, 1.26, 2.06, and 2.19 (each 3H, singlet), and one oxymethine proton at δH 3.80 (1H, triple doublet, J = 9.6, 9.6, 3.2 Hz). The presence of an oxymethylene group was indicated by the anisochronous signals of the geminal protons observed at δH 4.67 (1H, doublet, J = 12.4 Hz) and 4.21 (1H, doublet of doublets, J = 12.4, 1.2 Hz). An analysis of the HSQC and 13C spectroscopic data revealed that compound 1 consists of 27 carbon atoms, including six methyl groups, nine sp3 methylene groups (including one oxymethine), five sp3 methine groups (including one oxymethine), four sp3 quaternary carbons, and three sp2 quaternary carbons (including three carbonyls). Based on the 1H and 13C NMR spectroscopic data, compound 1 was found to possess an acetoxy group (δH 2.06, 3H, singlet; δC 170.8, carbonyl; 21.1, methyl) and two ketone carbonyls (δC 212.4, carbonyl; 214.4, carbonyl), accounting for three degrees of unsaturation. The remaining four degrees of unsaturation indicated that compound 1 possesses a tetracyclic structure.
To gain a more comprehensive understanding of the structure of 1, 2D NMR spectra, including 1H−1H correlation spectroscopy (COSY) and heteronuclear multiple-bond coherence (HMBC) spectra were employed. These spectra were used to analyze the correlations between protons (1H) and neighboring atoms, as well as the connectivity of multiple bonds between different nuclei (Figure 2). The 1H−1H COSY cross-peaks revealed four distinct proton spin systems: H-5/H2-6/H2-7, H2-1/H2-2/H2-3, H-9/H2-10, and H-14/H2-15/H-16/H-17/H2-18. The HMBC cross-signals between H3-21 and C-7, C-8, C-9, C-14; H3-20 and C-3, C-4, C-5, C-19; H2-22 and C-1, C-9, the ester carbonyl (δC 170.8); H3-23 and C-12, C-13, C-14, C-18; H3-25 and C-17, C-24; H2-11 and C-12 established connections between the partial structures. A comparison with the previous literature aided in successfully constructing the overall structure of 1. The obtained data showed similarities to a known 24-homescalarane compound, felixin F (5) [45]. However, notable differences were observed between compounds 1 and 5. In compound 5, the hydroxy group at C-22 (δH 3.93, 1H, doublet, J = 11.6 Hz; 4.07, 1H, doublet, J = 11.6 Hz; δC 62.7) was replaced by an acetoxy group in compound 1H 4.67, 1H, doublet, J = 12.4 Hz; 4.21, 1H, doublet of doublets, J = 12.4, 1.2 Hz; δC 64.5). Additionally, 1 lacked the aldehyde group present in compound 5, which was substituted by a methylene carbon.
Through a literature search, it was found that all naturally occurring scalaranes exhibit β-oriented methyl groups (Me-23 and Me-22) at C-13 and C-10, respectively, regardless of their oxidation state (-CH2OH, -CH2OAc, -COOH, and -CHO) [7]. The orientations of these methyl groups remained consistent within the carbon skeleton, with C-13 and C-10 serving as anchor points for analysis. The relative configuration of 1 was established through NOESY experiments (Figure 3). The NOESY interactions between H3-23 and H-17, as well as H3-21, and between H2-22 and H3-20, and H3-21, revealed the β-orientation of H3-23, H2-22, H3-21, H3-20, and H-17. An interaction between H-5 and H-9 indicated their spatial proximity. A further correlation analysis showed a correlation between H-9 and H-14, as well as a correlation between H-14 and H-16, suggesting an α-orientation for H-5, H-9, H-14, and H-16. Based on these findings, the structure of 1 was determined, and the assigned stereogenic centers were identified as (5S*,8R*,9S*,10R*,13R*,14S*, 16S*,17S*).
To determine the absolute stereochemistry of 1, two possible configurations were considered: 1-5S,8R,9S,10R,13R,14S,16S,17S, and 1-5R,8S,9R,10S,13S,14R,16R,17R. These configurations were fed into Gaussian 16 software to calculate the conformation, optimize the structure, and determine the specific optical rotation (SOR) values (Table 2). The calculated SOR value of 1-5S,8R,9S,10R,13R,14S,16S,17S (+31) was consistent with the experimental result of compound 1 (positive). Based on these results, the configurations of the stereogenic centers in 1 were determined to be (5S,8R,9S,10R,13R,14S,16S,17S). Consequently, the structure of 1 was identified as a new sesterterpenoid and named lendenfeldarane R.
Compound 2 was discovered to have the molecular formula C25H38O4, which was determined from a (+)-HRESIMS signal at m/z 425.26645 (calculated for C25H38O4+Na, 425.26623) and 13C data indicating the presence of seven unsaturated degrees. The IR spectrum of 2 showed absorption peaks for carbonyl groups (νmax 1707 cm−1 and 1676 cm−1) and a hydroxy group (νmax 3417 cm−1). Analyzing the 1D NMR data (Table 2), it was found that compound 2 was similar to felixin B (6) [46], with the main difference being the presence of a functional group at C-12. In compound 2, the chemical shift of H-12 (δH 3.98, 1H, br s) was shifted downfield compared to its counterpart in felixin B (6) (δH 4.97, 1H, dd, J = 2.8, 2.8 Hz). The absence of acetoxy signals suggested that felixin B (6) was the 12-acetyl derivative of compound 2. Further confirmation of the planar structure of compound 2 was obtained through the interpretation of 2D NMR spectroscopic data (Figure 2). The correlations observed among the chiral centers in the core rings A-D of 2 were identical to those observed in 1. In the NOESY experiment of 2 (Figure 3), the α-orientation of the hydroxy group at C-12 was determined based on the NOESY correlation between H-12 and H3-23. As a result, the configurations of the stereogenic carbons in 2 were established as (5S*,8R*,9S*,10R*,12S*, 13R*,14S*). The SOR was employed to determine the absolute configuration of 2. The calculated SOR values for 2-5S,8R,9S,10R,12S,13R,14S and 2-5R,8S,9R,10S,12R,13S,14R were positive (78) and negative (−78), respectively (Table 2). The experimental SOR data of 2 (positive) matched with the configuration 2-5S,8R,9S,10R,12S,13R,14S. Based on the aforementioned analyses, the structure of 2 was successfully determined, leading to its identification and designation as lendenfeldarane S.
Compound 3 was acquired in the form of a white powder, and its molecular formula was determined as C27H42O6 based on the HRESIMS signal at m/z 485.28737 [M + Na]+ (calculated for C27H42O6 + Na, 485.28736) and 13C data suggesting the presence of seven degrees of unsaturation. The IR spectrum of 3 exhibited prominent peaks at νmax 3459, 1727, and 1666, indicating the presence of hydroxy, ester, and α,β-unsaturated ketone groups, respectively. Analysis of the 1H and 13C NMR spectroscopic data (Table 3) revealed the presence of an acetoxy group (δH 2.03, 3H, s; δC 170.8, C; 21.3, CH3) and a ketonic carbonyl group (δC 199.3) within compound 3. Moreover, the 13C resonances at δC 133.6 (C) and 155.7 (CH) suggested the existence of a trisubstituted olefin, accounting for three degrees of unsaturation. Compound 3 was identified as an analog of tetracyclic sesterterpenoid. The NMR data resembled those of felixin C (7) [46]. The chemical shift of H-16 in felixin C (7) (δH 4.55, d, J = 4.8 Hz) was shifted downfield in 3H 4.91, dd, J = 4.0, 1.6 Hz), with an additional signal for a hydroxyperoxide group (δH 9.27, 1H, s). The presence of a hydroperoxide group substitution at position 16 was deduced from the HMBC cross-peak (Figure 2) between H-16 and C-14, C-17, C-18, and from H-14 to C-16, as well as from the COSY correlation between H-15, H-14, and H-16. Additionally, comparing the 13C NMR data with those of 1 and 4, the C-16 carbon resonating at δC 77.2 in 3 was more downfield than δC 70.9 in 1 and δC 68.1 in 4, revealing that the hydroxyperoxide group was located at C-16 position.
Further insights into the structure of 3 were obtained through the NOESY experiment (Figure 3). The observed correlations provided valuable information regarding the configurations of the chiral centers in the core rings A-C of 3, which were found to be identical to those observed in 1. Notably, H3-23 showed NOE correlations with H3-21 and H-12, indicating the β-orientation of H-12. Additionally, H-14 displayed NOE correlations with H-9 and H-16, indicating the β-orientation of the hydroperoxide group at C-16. To determine the absolute configuration of 3, a comparison was made between its experimental optical rotation and the corresponding SOR value. The calculated SOR values for 3-5S,8R,9S,10R,12S,13R,14S,16S and 3-5R,8S,9R,10S,12R,13S,14R,16R were positive (15) and negative (−15), respectively (Table 2). The experimental SOR data of 3 (positive) matched with the configuration 3-5S,8R,9S,10R,12S,13R,14S,16S. Based on the aforementioned results, the structure of 3 was determined and assigned the name lendenfeldarane T.
Compound 4 was obtained in the form of an unstructured fine powder. Its molecular formula was determined to be C27H42O5 based on the (+)-HRESIMS pseudo-molecular ion peak at m/z 469.29221 (calculated for C27H42O5+Na, 469.29245) and 13C data, which indicated the presence of seven degrees of unsaturation. The IR spectrum of 4 exhibited absorption bands corresponding to hydroxy groups (maximum absorption at 3486 cm−1), ester carbonyl groups (maximum absorption at 1726 cm−1), and α,β-unsaturated ketone groups (maximum absorption at 1655 cm−1). The structure of 4 was determined through the analysis of 1D and 2D NMR spectroscopic data (refer to Table 3). Based on these findings, it was observed that the overall structure of 4 closely resembled that of felixin B (7) [46]. The 13C and 1H NMR data of 4 were similar to those of compound 7, except for carbon resonances at C-16 and C-17, which appeared at δC 63.3 (CH) and 138.2 (C) in compound 7 and at δC 68.1 (CH) and 134.2 (C) in 4, indicating that 4 was the 16S isomer of compound 7. Analysis of the NOESY spectra (refer to Figure 3) revealed a cross-peak from H-14 to H-16, indicating an α-orientation of H-16 in 4. According to the above data, the structure of 4 was determined, and the compound was named lendenfeldarane U.
N-formyl-methionyl-leucyl-phenylalanine (fMLF) and pathogen-associated molecular patterns (PAMPs) can stimulate neutrophils, leading to the initiation of inflammatory responses such as the generation of O2•− (respiratory burst) and the release of elastase (degranulation) [7]. To evaluate the anti-inflammatory properties of nor-24-homoscalaranes, all the isolated compounds were tested on fMLF-induced human neutrophils, and the findings are summarized in Table 4. As a positive control, LY294002, a phosphoinositide 3-kinase (PI3K) inhibitor, was employed, given the established role of PI3K in regulating neutrophil respiratory burst and/or degranulation [47,48]. Compounds 2 and 4 demonstrated significant activity against both O2•− accumulation (IC50 = 3.98–4.46 μM) and elastase release (IC50 = 4.73–5.24 μM). On the other hand, compounds 1 and 3 were inactive at a concentration of 10 μM. These results emphasized the importance of the conjugated functionality at C-17-18-24 for anti-inflammatory activity, while the substitution of the peroxyl group at C-16 may diminish this effect.

3. Materials and Methods

3.1. General Experimental Procedures

IR spectra were recorded using a Thermo Scientific Nicolet iS5 FT-IR spectrophotometer (Thermo Scientific, Waltham, MA, USA). Optical activities were measured using a JASCO P-1010 polarimeter (JASCO, Tokyo, Japan). NMR spectra were obtained using JEOL ECZ 400S or 600R NMR spectrometers (JEOL, Tokyo, Japan) with CDCl3 (Sigma-Aldrich, St. Louis, MO, USA) as the deuterated solvent. The detected signals in 1H and 13C NMR were corrected at 7.26 ppm (singlet) and 77.0 ppm (triplet), respectively. The coupling constants (J) were converted to Hz. MS data, including ESIMS and HRESIMS, were obtained using a Bruker 7 Tesla solera FTMS system (Bruker, Bremen, Germany). Two types of TLC analyses were performed using aluminum plates coated with Kieselgel 60 F254 (0.25 mm) and RP-18 F254S (0.25 mm) (Merck, Darmstadt, Germany). For chromatographic separation, a glass column was used, which was packed with a stationary phase of silica gel 60 (40–63 μm and 63–200 μm, Merck, Darmstadt, Germany). NP-HPLC was conducted using a system consisting of a solvent delivery system (L-7110, Hitachi, Tokyo, Japan) and a preparative packed silica gel column (YMC-Pack SIL, SIL-06, 250 × 20 mm, D. S-5 μm) (Sigma-Aldrich, St. Louis, MO, USA). RP-HPLC was performed using a system consisting of a solvent delivery system (L-2130, Hitachi, Tokyo, Japan), a PDA detector (L-2455, Hitachi, Tokyo, Japan), and a C18 column (Luna 5 μm, C18(2) 100Å AXIA Packed, 250 × 21.2 mm) (Phenomenex, Torrance, CA, USA).

3.2. Animal Material and Isolation of Compounds

The Lendenfeldia sp. specimen was obtained by scuba diving off the coast of Southern Taiwan in April 2019. At the National Museum of Marine Biology & Aquarium, Taiwan, a specimen with a voucher number (specimen No. 2019-04-SP) was deposited. Lendenfeldia sp. was taxonomically identified by Prof. Yusheng M. Huang from the National Penghu University of Science and Technology, Taiwan. Lendenfeldia sp. was collected (2.9 kg) and freeze-dried. The freeze-dried material (213 g, dry weight) was minced and an extract was prepared using a 1:1 mixture of CH2Cl2:MeOH (1 L × 6). The resulting extract underwent liquid–liquid partitioning between EtOAc and H2O. The obtained EtOAc layer (7.9 g) was further subjected to normal-phase column chromatography. Elution was carried out using a gradient solvent system comprising n-hexane, followed by increasing polarity mixtures of n-hexane and EtOAc, pure acetone, and finally pure methanol as eluting solvents. This process resulted in the production of 14 sub-fractions labeled A–N. Sub-fraction H underwent reversed-phase chromatography on C18 silica gel, eluted with a MeOH:H2O mixture (50% MeOH to pure MeOH), and yielded six additional sub-fractions, H1–H6. Sub-fraction H5 then underwent RP-HPLC using an isocratic solvent system of MeOH:H2O (8:2), resulting in the isolation of compound 2 (1.3 mg). Fraction I was purified using NP-HPLC using a CH2Cl2:acetone mixture (4:1) as an isocratic solvent system with a flow rate of 3.0 mL/min, yielding eight sub-fractions, which were labeled I1–I8. Sub-fraction I5 was further purified using RP-HPLC with an isocratic solvent system of MeOH:H2O (4:1) at a flow rate of 5 mL/min, resulting in the isolation of compounds 3 (0.9 mg) and 4 (2.1 mg). Similarly, sub-fraction I6 underwent RP-HPLC using an isocratic solvent system of MeOH:H2O (4:1) at a flow rate of 5 mL/min, leading to the isolation of compound 1 (0.5 mg).
Lendenfeldarane R (1): Amorphous powder; [ α ] D 25 +21 (c 0.025, CHCl3); IR (ATR) νmax 3429, 1736, 1703 cm−1; ESIMS m/z 469 [M + Na]+; HRESIMS m/z 469.29256 [M + Na]+ (calcd. for C27H42O5 + Na, 469.29245); 1H (400 MHz, CDCl3) and 13C (100 MHz, CDCl3) NMR spectroscopic data.
Lendenfeldarane S (2): Amorphous powder; [ α ] D 25 +61 (c 0.025, CHCl3); IR (ATR) νmax 3417, 1707, 1676 cm−1; ESIMS m/z 425 [M + Na]+; HRESIMS m/z 425.26645 [M + Na]+ (calcd. for C25H38O4 + Na, 425.26623); 1H (600 MHz, CDCl3) and 13C (150 MHz, CDCl3) NMR spectroscopic data.
Lendenfeldarane T (3): Amorphous powder; [ α ] D 25 +40 (c 0.045, CHCl3); IR (ATR) νmax 3459, 1727, 1666 cm−1; ESIMS m/z 485 [M + Na]+; HRESIMS m/z 485.28737 [M + Na]+ (calcd. for C27H42O6 + Na, 485.28736); 1H (400 MHz, CDCl3) and 13C (100 MHz, CDCl3) NMR spectroscopic data.
Lendenfeldarane U (4): Amorphous powder; [ α ] D 25 +76 (c 0.105, CHCl3); IR (ATR) νmax 3486, 1726, 1655 cm−1; ESIMS m/z 469 [M + Na]+; HRESIMS m/z 469.29221 [M + Na]+ (calcd. for C27H42O5 + Na, 469.28736); 1H (600 MHz, CDCl3) and 13C (150 MHz, CDCl3) NMR spectroscopic data.

3.3. In Silico Calculations

To minimize molecular energy, the molecular structures were optimized at the MM2 level, resulting in the generation of a mol file. This mol file was analyzed using the MMFF94 force field in GaussView 6.1 (Gaussian Inc., Wallingford, CT, USA) with the assistance of the GMMX package, which allowed us to explore the conformational search results. The obtained data were imported into Gaussian 16 software (Gaussian Inc., Wallingford, CT, USA). In Gaussian 16, the structures underwent further optimization using the time-dependent density functional theory (TDDFT) methodology. The optimization was conducted in the solvent phase using the PCM/mpw1pw91/6-31 g(d,p) levels, enabling GIAO-DFT calculation. In the final step, the computed NMR results were averaged, taking into consideration the proportion of each conformer [49].

3.4. Preparation of Human Neutrophils

Blood samples were collected via venipuncture from human donors aged between 20 and 30 years. The Institutional Review Board (IRB) of Chang Gung Memorial Hospital approved and oversaw the protocol, which was identified as IRB No. 202002493A3. Neutrophil purification was performed using a previously established technique [7]. The process involved several steps, including hypotonic lysis, dextran sedimentation, and separation of erythrocytes using a Ficoll Hypaque gradient. Once the human neutrophils were isolated, they were placed in a 50 mL centrifuge tube containing an HBSS buffer solution devoid of calcium (Ca2+) and magnesium (Mg2+). The pH of the solution was adjusted to 7.4, and the viability of the neutrophils was assessed using the trypan blue exclusion method to ensure that more than 98% of the cells remained viable. Subsequently, the neutrophils were examined in HBSS with 1 mM CaCl2 at a temperature of 37 °C.

3.5. Measurement of Superoxide Anion (O2•−) Generation

The evaluation of O2•− generation involved the utilization of ferricytochrome c and the inhibitory effect of superoxide dismutase (SOD) on its reduction process [15]. After adding ferricytochrome c (0.6 mg/mL), neutrophils at a concentration of 6 × 105 cells/mL were equilibrated at 37 °C and incubated for 5 min. Subsequently, the neutrophils were treated with either pure compounds or DMSO (0.1% as a control). To enhance the reaction, cytochalasin B (CB) was introduced at a concentration of 1 μg/mL [7]. The activated mixture was then incubated for 3 min after stimulation with 0.1 μM fMLF. The reduction of ferricytochrome c was continuously monitored at 550 nm by measuring absorbance changes using a spectrophotometer (U-3010, Hitachi, Tokyo, Japan).

3.6. Measurement of Elastase Release

To evaluate the degranulation of azurophilic granules, an elastase release assay was conducted using MeO-Suc-Ala-Ala-Pro-Val-p-nitroanilide as the substrate for elastase [7]. Neutrophils (6 × 105 cells/mL) were equilibrated at 37 °C after the addition of MeO-Suc-Ala-Ala-Pro-Val-p-nitroanilide (100 μM). The cells were then incubated for 5 min before treatment with pure compounds. To enhance the reaction, CB (0.5 g/mL) was added, followed by the introduction of fMLF (0.1 μM) to induce cellular activation. Elastase release was assessed by continuously monitoring changes in absorbance at 405 nm.

3.7. Statistics

Statistical calculations were performed using Student’s t-test (GraphPad Software 9.0.2, San Diego, CA, USA). A significance level of p < 0.05 was employed to determine statistical significance.

4. Conclusions

In this study, we identified and characterized four new 24-homoscalarane compounds, lendenfeldaranes R–U (14), isolated from Lendenfeldia sp. through comprehensive chromatographic and spectroscopic analyses. The structures, relative stereochemistry, and absolute configurations of these compounds were determined. The in vitro assays demonstrated that compounds possessing C-17-18-24-conjugated functionality exhibited significant inhibitory effects on key inflammatory responses, including the generation of superoxide anions and the release of elastase in human neutrophils. Our findings showed that Lendenfeldia sp. is an excellent source of anti-inflammatory agents with unique structures that can be further developed into potential drug leads.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ph16091258/s1, Figure S1: IR spectrum of compound 1; Figure S2: ESIMS spectrum of compound 1; Figure S3: HRESIMS spectrum of compound 1; Figure S4: 1H-NMR spectrum (400 MHz) of compound 1 in CDCl3; Figure S5: 13C NMR spectrum (100 MHz) of compound 1 in CDCl3; Figure S6: HSQC spectrum of compound 1 in CDCl3; Figure S7: HMBC spectrum of compound 1 in CDCl3; Figure S8: 1H-1H COSY spectrum of compound 1 in CDCl3; Figure S9: NOESY spectrum of compound 1 in CDCl3; Figure S10: IR spectrum of compound 2; Figure S11: ESIMS spectrum of compound 2; Figure S12: HRESIMS spectrum of compound 2; Figure S13: 1H NMR spectrum (600 MHz) of compound 2 in CDCl3; Figure S14: 13C NMR spectrum (100 MHz) of compound 2 in CDCl3; Figure S15: HSQC spectrum of compound 2 in CDCl3; Figure S16: HMBC spectrum of compound 2 in CDCl3; Figure S17: 1H-1H COSY spectrum of compound 2 in CDCl3; Figure S18: NOESY spectrum of compound 2 in CDCl3; Figure S19: IR spectrum of compound 3; Figure S20: ESIMS spectrum of compound 3; Figure S21: HRESIMS spectrum of compound 3; Figure S22: 1H NMR spectrum (400 MHz) of compound 3 in CDCl3; Figure S23: 13C NMR spectrum (100 MHz) of compound 3 in CDCl3; Figure S24: DEPT spectrum of compound 3 in CDCl3; Figure S25: HSQC spectrum of compound 3 in CDCl3; Figure S26: HMBC spectrum of compound 3 in CDCl3; Figure S27: 1H-1H COSY spectrum of compound 3 in CDCl3; Figure S28: NOESY spectrum of compound 3 in CDCl3; Figure S29: IR spectrum of compound 4; Figure S30: ESIMS spectrum of compound 4; Figure S31: HRESIMS spectrum of compound 4; Figure S32: 1H NMR spectrum (600 MHz) of compound 4 in CDCl3; Figure S33: 13C NMR spectrum (150 MHz) of compound 4 in CDCl3; Figure S34: HSQC spectrum of compound 4 in CDCl3; Figure S35: HMBC spectrum of compound 4 in CDCl3; Figure S36: 1H-1H COSY spectrum of compound 4 in CDCl3; Figure S37: NOESY spectrum of compound 4 in CDCl3.

Author Contributions

The experiments were conceived and designed by B.-R.P., K.-H.L. and P.-J.S. Sample collections, extraction, isolation, and structure determination were performed by B.-R.P., L.-G.Z. and L.-Y.C. T.-L.H. and K.-H.L. conducted the pharmacological experiments. J.-H.S., T.-L.H., K.-H.L. and P.-J.S. contributed reagents and analysis tools. Data interpretation, manuscript writing, and paper revision were carried out by B.-R.P., M.E.-S., M.-H.L., K.-H.L. and P.-J.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received primary support from grants awarded to P.-J.S. from the National Museum of Marine Biology & Aquarium and the National Science and Technology Council (Grant Nos MOST 109-2320-B-291-001-MY3 and 111-2320-B-291-001), Taiwan. Additional support was provided by grants from the National Science and Technology Council (MOST 110-2320-B-038-034, MOST 111-2320-B-038-040-MY3; MOST 111-2321-B-255-001), Ministry of Education (DP2-110-21121-01-N-12-03), and Taipei Medical University (TMU109-AE1-B15) awarded to K.-H. L. The authors express their gratitude for all the funding received.

Institutional Review Board Statement

The study adhered to the principles outlined in the Declaration of Helsinki and received approval from the Institutional Review Board (IRB) at Chang Gung Memorial Hospital. The protocol for studies involving human participants was assigned the code No. 202002493A3.

Informed Consent Statement

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

Data Availability Statement

Data is contained within the article and Supplementary Material.

Acknowledgments

The authors express their gratitude to Hsiao-Ching Yu and Chao-Lien Ho from the High Valued Instrument Center, National Sun Yat-Sen University, for providing the mass (MS 006500) and NMR (NMR 001100) spectra (NSTC 112-2740-M-110-002).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Galitz, A.; Cook, S.d.C.; Ekins, M.; Hooper, J.N.; Naumann, P.T.; De Voogd, N.J.; Wahab, M.A.; Wörheide, G.; Erpenbeck, D. Identification of an aquaculture poriferan “Pest with Potential” and its phylogenetic implications. Peer J 2018, 6, e5586. [Google Scholar] [CrossRef]
  2. Alvi, K.; Crews, P. Homoscalarane sesterterpenes from Lendenfeldia frondosa. J. Nat. Prod. 1992, 55, 859–865. [Google Scholar] [CrossRef] [PubMed]
  3. Chill, L.; Rudi, A.; Aknin, M.; Loya, S.; Hizi, A.; Kashman, Y. New sesterterpenes from Madagascan Lendenfeldia sponges. Tetrahedron 2004, 60, 10619–10626. [Google Scholar] [CrossRef]
  4. Kazlauskas, R.; Murphy, P.; Wells, R. Five new C26 tetracyclic terpenes from a sponge (Lendenfeldia sp.). Aust. J. Chem. 1982, 35, 51–59. [Google Scholar] [CrossRef]
  5. Peng, B.-R.; Lai, K.-H.; Chang, Y.-C.; Chen, Y.-Y.; Su, J.-H.; Huang, Y.M.; Chen, P.-J.; Yu, S.S.-F.; Duh, C.-Y.; Sung, P.-J. Sponge-derived 24-homoscalaranes as potent anti-inflammatory agents. Mar. Drugs 2020, 18, 434. [Google Scholar] [CrossRef] [PubMed]
  6. Peng, B.-R.; Lai, K.-H.; Chen, Y.-Y.; Su, J.-H.; Huang, Y.M.; Chen, Y.-H.; Lu, M.-C.; Yu, S.S.-F.; Duh, C.-Y.; Sung, P.-J. Probing anti-proliferative 24-homoscalaranes from a sponge Lendenfeldia sp. Mar. Drugs 2020, 18, 76. [Google Scholar] [CrossRef]
  7. Peng, B.-R.; Lai, K.-H.; Lee, G.-H.; Yu, S.S.-F.; Duh, C.-Y.; Su, J.-H.; Zheng, L.-G.; Hwang, T.-L.; Sung, P.-J. Scalarane-type sesterterpenoids from the marine sponge Lendenfeldia sp. alleviate inflammation in human neutrophils. Mar. Drugs 2021, 19, 561. [Google Scholar] [CrossRef]
  8. Sera, Y.; Adachi, K.; Shizuri, Y. A new epidioxy sterol as an antifouling substance from a Palauan marine sponge, Lendenfeldia chondrodes. J. Nat. Prod. 1999, 62, 152–154. [Google Scholar] [CrossRef]
  9. Dai, J.; Liu, Y.; Zhou, Y.-D.; Nagle, D.G. Cytotoxic metabolites from an Indonesian sponge Lendenfeldia sp. J. Nat. Prod. 2007, 70, 1824–1826. [Google Scholar] [CrossRef]
  10. Alahdal, A.M.; Asfour, H.Z.; Ahmed, S.A.; Noor, A.O.; Al-Abd, A.M.; Elfaky, M.A.; Elhady, S.S. Anti-helicobacter, antitubercular and cytotoxic activities of scalaranes from the red sea sponge Hyrtios erectus. Molecules 2018, 23, 978. [Google Scholar] [CrossRef]
  11. Cao, F.; Wu, Z.-H.; Shao, C.-L.; Pang, S.; Liang, X.-Y.; de Voogd, N.J.; Wang, C.-Y. Cytotoxic scalarane sesterterpenoids from the South China Sea sponge Carteriospongia foliascens. Org. Biomol. Chem. 2015, 13, 4016–4024. [Google Scholar] [CrossRef] [PubMed]
  12. Elhady, S.S.; Al-Abd, A.M.; El-Halawany, A.M.; Alahdal, A.M.; Hassanean, H.A.; Ahmed, S.A. Antiproliferative scalarane-based metabolites from the Red Sea sponge Hyrtios erectus. Mar. Drugs 2016, 14, 130. [Google Scholar] [CrossRef] [PubMed]
  13. Elhady, S.S.; El-Halawany, A.M.; Alahdal, A.M.; Hassanean, H.A.; Ahmed, S.A. A new bioactive metabolite isolated from the Red Sea marine sponge Hyrtios erectus. Molecules 2016, 21, 82. [Google Scholar] [CrossRef] [PubMed]
  14. Hahn, D.; Won, D.H.; Mun, B.; Kim, H.; Han, C.; Wang, W.; Chun, T.; Park, S.; Yoon, D.; Choi, H. Cytotoxic scalarane sesterterpenes from a Korean marine sponge Psammocinia sp. Bioorg. Med. Chem. Lett. 2013, 23, 2336–2339. [Google Scholar] [CrossRef]
  15. Hassan, M.H.; Rateb, M.E.; Hetta, M.; Abdelaziz, T.A.; Sleim, M.A.; Jaspars, M.; Mohammed, R. Scalarane sesterterpenes from the Egyptian Red Sea sponge Phyllospongia lamellosa. Tetrahedron 2015, 71, 577–583. [Google Scholar] [CrossRef]
  16. Jeon, J.-E.; Bae, J.; Lee, K.J.; Oh, K.-B.; Shin, J. Scalarane sesterterpenes from the sponge Hyatella sp. J. Nat. Prod. 2011, 74, 847–851. [Google Scholar] [CrossRef]
  17. Kaweetripob, W.; Mahidol, C.; Tuntiwachwuttikul, P.; Ruchirawat, S.; Prawat, H. Cytotoxic sesterterpenes from Thai marine sponge Hyrtios erectus. Mar. Drugs 2018, 16, 474. [Google Scholar] [CrossRef]
  18. Kwon, O.-S.; Kim, D.; Kim, C.-K.; Sun, J.; Sim, C.J.; Oh, D.-C.; Lee, S.K.; Oh, K.-B.; Shin, J. Cytotoxic scalarane sesterterpenes from the sponge Hyrtios erectus. Mar. Drugs 2020, 18, 253. [Google Scholar] [CrossRef]
  19. Lai, K.-H.; Liu, Y.-C.; Su, J.-H.; El-Shazly, M.; Wu, C.-F.; Du, Y.-C.; Hsu, Y.-M.; Yang, J.-C.; Weng, M.-K.; Chou, C.-H. Antileukemic scalarane sesterterpenoids and meroditerpenoid from Carteriospongia (Phyllospongia) sp., induce apoptosis via dual inhibitory effects on topoisomerase II and Hsp90. Sci. Rep. 2016, 6, 36170. [Google Scholar] [CrossRef]
  20. Lee, Y.-J.; Kim, S.-H.; Choi, H.; Lee, H.-S.; Lee, J.-S.; Shin, H.-J.; Lee, J. Cytotoxic furan-and pyrrole-containing scalarane sesterterpenoids isolated from the sponge Scalarispongia sp. Molecules 2019, 24, 840. [Google Scholar] [CrossRef]
  21. Lu, D.; Luo, X.-C.; Liu, J.; Wu, G.-L.; Yu, Y.; Xu, Y.-N.; Lin, H.-W.; Yang, F. Phyllofolactones NT, bioactive bishomoscalarane sesterterpenoids from the marine sponge Phyllospongia foliascens. Tetrahedron 2023, 137, 133382. [Google Scholar] [CrossRef]
  22. Shin, A.-Y.; Lee, H.-S.; Lee, J. Isolation of Scalimides A–L: β-alanine-bearing scalarane analogs from the marine sponge Spongia sp. Mar. Drugs 2022, 20, 726. [Google Scholar] [CrossRef]
  23. Shin, A.-Y.; Son, A.; Choi, C.; Lee, J. Isolation of scalarane-type sesterterpenoids from the marine sponge Dysidea sp. and stereochemical reassignment of 12-epi-phyllactone D/E. Mar. Drugs 2021, 19, 627. [Google Scholar] [CrossRef] [PubMed]
  24. Su, M.-Z.; Zhang, Q.; Yao, L.-G.; Wu, B.; Guo, Y.-W. Hyrtiosins F and G, two new scalarane sesterterpenes from the South China sea sponge Hyrtios erecta. J. Asian Nat. Prod. Res. 2022, 25, 741–747. [Google Scholar] [CrossRef]
  25. Tommonaro, G.; De Rosa, S.; Carnuccio, R.; Maiuri, M.C.; De Stefano, D. Marine sponge sesterpenoids as potent apoptosis-inducing factors in human carcinoma cell lines. In Handbook of Anticancer Drugs from Marine Origin; Springer: Cham, Switzerland, 2015; pp. 439–479. [Google Scholar]
  26. Tran, H.N.K.; Kim, M.J.; Lee, Y.-J. Scalarane sesterterpenoids isolated from the marine sponge Hyrtios erectus and their Cytotoxicity. Mar. Drugs 2022, 20, 604. [Google Scholar] [CrossRef] [PubMed]
  27. Yang, F.; Gan, J.-H.; Liu, X.-Y.; Lin, H.-W. Scalarane sesterterpenes from the Paracel islands marine sponge Hyrtios sp. Nat. Prod. Commun. 2014, 9, 763–764. [Google Scholar] [CrossRef]
  28. Yang, I.; Lee, J.; Lee, J.; Hahn, D.; Chin, J.; Won, D.H.; Ko, J.; Choi, H.; Hong, A.; Nam, S.-J. Scalalactams A–D, scalarane sesterterpenes with a γ-lactam moiety from a Korean Spongia sp. Marine sponge. Molecules 2018, 23, 3187. [Google Scholar] [CrossRef]
  29. Yang, I.; Nam, S.-J.; Kang, H. Two new scalaranes from a Korean marine sponge Spongia sp. Nat. Prod. Sci. 2015, 21, 289–292. [Google Scholar] [CrossRef]
  30. Yu, H.-B.; Hu, B.; Wu, G.-F.; Ning, Z.; He, Y.; Jiao, B.-H.; Liu, X.-Y.; Lin, H.-W. Phyllospongianes A–E, dinorscalarane sesterterpenes from the marine sponge Phyllospongia foliascens. J. Nat. Prod. 2023, 86, 1754–1760. [Google Scholar] [CrossRef]
  31. Zhang, H.; Crews, P.; Tenney, K.; Valeriote, F.A. Cytotoxic phyllactone analogs from the marine sponge Phyllospongia papyrecea. Med. Chem. 2017, 13, 295–300. [Google Scholar] [CrossRef]
  32. Chakraborty, K.; Francis, P. Hyrtioscalaranes A and B, two new scalarane-type sesterterpenes from Hyrtios erectus with anti-inflammatory and antioxidant effects. Nat. Prod. Res. 2021, 35, 5559–5570. [Google Scholar] [CrossRef] [PubMed]
  33. Francis, P.; Chakraborty, K. Anti-inflammatory scalarane-type sesterterpenes, erectascalaranes A–B, from the marine sponge Hyrtios erectus attenuate pro-inflammatory cyclooxygenase-2 and 5-lipoxygenase. Med. Chem. Res. 2021, 30, 886–896. [Google Scholar] [CrossRef]
  34. Kikuchi, H.; Tsukitani, Y.; Shimizu, I.; Kobayashi, M.; Kitagawa, I. Foliaspongin, an antiinflammatory bishomosesterterpene from the marine sponge Phyllospongia foliascens (Pallas). Chem. Pharm. Bull. 1981, 29, 1492–1494. [Google Scholar] [CrossRef]
  35. Kikuchi, H.; Tsukitani, Y.; Shimizu, I.; Kobayashi, M.; Kitagawa, I. Marine natural products. XI. An antiinflammatory scalarane-type bishomosesterterpene, foliaspongin, from the Okinawan marine sponge Phyllospongia foliascens (Pallas). Chem. Pharm. Bull. 1983, 31, 552–556. [Google Scholar] [CrossRef]
  36. Lee, S.-M.; Kim, N.-H.; Lee, S.; Kim, Y.-N.; Heo, J.-D.; Jeong, E.-J.; Rho, J.-R. Deacetylphylloketal, a new phylloketal derivative from a marine sponge, genus Phyllospongia, with potent anti-inflammatory activity in in vitro co-culture model of intestine. Mar. Drugs 2019, 17, 634. [Google Scholar] [CrossRef]
  37. Chang, L.C.; Otero-Quintero, S.; Nicholas, G.M.; Bewley, C.A. Phyllolactones A–E: New bishomoscalarane sesterterpenes from the marine sponge Phyllospongia lamellose. Tetrahedron 2001, 57, 5731–5738. [Google Scholar] [CrossRef]
  38. Choi, J.-H.; Lee, H.-S.; Campos, W.L. Scalarane-type Sesterterpenes from the Philippines Sponge Hyrtios sp. Ocean Polar Res. 2020, 42, 15–20. [Google Scholar]
  39. Daletos, G.; Ancheeva, E.; Chaidir, C.; Kalscheuer, R.; Proksch, P. Antimycobacterial metabolites from marine invertebrates. Arch. Pharm. 2016, 349, 763–773. [Google Scholar] [CrossRef]
  40. Hertzer, C.; Kehraus, S.; Böhringer, N.; Kaligis, F.; Bara, R.; Erpenbeck, D.; Wörheide, G.; Schäberle, T.F.; Wägele, H.; König, G.M. Antibacterial scalarane from Doriprismatica stellata nudibranchs (Gastropoda, Nudibranchia), egg ribbons, and their dietary sponge Spongia cf. agaricina (Demospongiae, Dictyoceratida). Beilstein J. Org. Chem. 2020, 16, 1596–1605. [Google Scholar] [CrossRef]
  41. Song, J.; Jeong, W.; Wang, N.; Lee, H.-S.; Sim, C.J.; Oh, K.-B.; Shin, J. Scalarane sesterterpenes from the sponge Smenospongia sp. J. Nat. Prod. 2008, 71, 1866–1871. [Google Scholar] [CrossRef]
  42. Chen, L.Y.; Peng, B.R.; Lai, G.Y.; Weng, H.J.; El-Shazly, M.; Su, C.H.; Su, J.H.; Sung, P.J.; Liao, C.P.; Lai, K.H. Chemometric-guided exploration of marine anti-neurofibroma leads. Front. Mar. Sci. 2022, 9, 930736. [Google Scholar] [CrossRef]
  43. Herrero-Cervera, A.; Soehnlein, O.; Kenne, E. Neutrophils in chronic inflammatory diseases. Cell. Mol. Immunol. 2022, 19, 177–191. [Google Scholar] [CrossRef] [PubMed]
  44. Wéra, O.; Lancellotti, P.; Oury, C. The dual role of neutrophils in inflammatory bowel diseases. J. Clin. Med. 2016, 5, 118. [Google Scholar] [CrossRef]
  45. Lai, Y.-Y.; Chen, L.-C.; Wu, C.-F.; Lu, M.-C.; Wen, Z.-H.; Wu, T.-Y.; Fang, L.-S.; Wang, L.-H.; Wu, Y.-C.; Sung, P.-J. New cytotoxic 24-homoscalarane sesterterpenoids from the sponge Ircinia felix. Int. J. Mol. Sci. 2015, 16, 21950–21958. [Google Scholar] [CrossRef] [PubMed]
  46. Lai, Y.-Y.; Lu, M.-C.; Wang, L.-H.; Chen, J.-J.; Fang, L.-S.; Wu, Y.-C.; Sung, P.-J. New scalarane sesterterpenoids from the Formosan sponge Ircinia felix. Mar. Drugs 2015, 13, 4296–4309. [Google Scholar] [CrossRef]
  47. Hii, C.S.; Marin, L.A.; Halliday, D.; Roberton, D.M.; Murray, A.W.; Ferrante, A. Regulation of human neutrophil-mediated cartilage proteoglycan degradation by phosphatidylinositol-3-kinase. Immunology 2001, 102, 59–66. [Google Scholar] [CrossRef] [PubMed]
  48. Chen, P.J.; Tseng, H.H.; Wang, Y.H.; Fang, S.Y.; Chen, S.H.; Chen, C.H.; Tsai, S.C.; Chang, Y.C.; Tsai, Y.F.; Hwang, T.L. Palbociclib blocks neutrophilic phosphatidylinositol 3-kinase activity to alleviate psoriasiform dermatitis. Br. J. Pharmacol. 2023, 180, 2172–2188. [Google Scholar] [CrossRef]
  49. Grimblat, N.; Zanardi, M.M.; Sarotti, A.M. Beyond DP4: An improved probability for the stereochemical assignment of isomeric compounds using quantum chemical calculations of NMR shifts. J. Org. Chem. 2015, 80, 12526–12534. [Google Scholar] [CrossRef]
Figure 1. The identified 24-homoscalaranes from the marine sponge Lendenfeldia sp.
Figure 1. The identified 24-homoscalaranes from the marine sponge Lendenfeldia sp.
Pharmaceuticals 16 01258 g001
Figure 2. Key 2D correlations of planar structures of 14.
Figure 2. Key 2D correlations of planar structures of 14.
Pharmaceuticals 16 01258 g002
Figure 3. Selective NOESY correlations (Pharmaceuticals 16 01258 i001) of relative structures of 14.
Figure 3. Selective NOESY correlations (Pharmaceuticals 16 01258 i001) of relative structures of 14.
Pharmaceuticals 16 01258 g003
Table 1. The NMR data for compounds 1 and 2.
Table 1. The NMR data for compounds 1 and 2.
12
C/HδH (J in Hz) aδC (mult.) bδH (J in Hz) cδC (mult.) d
12.00 m; 0.77 ddd (12.4, 12.4, 4.4) e34.3, CH2 f2.13 m; 0.73 ddd (12.0, 12.0, 3.0) e34.1, CH2 f
21.61 m; 1.48 m17.9, CH21.56 m; 1.37 m17.7, CH2
31.45 m; 1.17 m41.4, CH21.43 m; 1.19 m41.7, CH2
4 33.0, C 33.0, C
51.00 m56.8, CH1.03 dd (12.6, 2.4)56.9, CH
61.61 m18.2, CH21.49 m18.4, CH2
71.95 m; 1.15 m41.5, CH21.77 ddd (12.6, 3.0, 3.0); 1.08 m41.1, CH2
8 37.2, C 37.3, C
91.32 m60.2, CH1.52 m51.8, CH
10 41.0, C 41.8, C
113.97 dd (14.4, 14.4)
2.50 dd (14.4, 2.8)
37.6, CH22.32 m; 1.86 ddd (16.2, 2.4, 2.4)28.3, CH2
12 214.4, C3.98 br s73.9, CH
13 48.5, C 42.6, C
141.19 m55.9, CH2.14 m47.5, CH
151.98 m; 1.57 m27.6, CH22.53 m; 2.24 m34.9, CH2
163.80 ddd (9.6, 9.6, 3.2)70.9, CH 198.2, C
172.49 m53.7, CH 142.8, C
182.05 m36.8, CH7.55 s165.9, CH
190.88 s33.7, CH30.86 s33.9, CH3
200.85 s21.8, CH30.77 s21.8, CH3
211.13 s16.0, CH31.11 s15.8, CH3
224.67 d (12.4)
4.21 dd (12.4, 1.2)
64.5, CH24.06 d (11.4); 3.09 dd (11.4)62.8, CH2
231.26 s19.3, CH31.11 s18.6, CH3
24 212.4, C 198.3, C
252.19 s28.8, CH32.45 s30.7, CH3
22-OAc 170.8, C
2.06 s21.1, CH3
a Spectra recorded at 400 MHz in CDCl3; b spectra recorded at 100 MHz in CDCl3; c spectra recorded at 600 MHz in CDCl3; d spectra recorded at 150 MHz in CDCl3; e J values (in Hz) in parentheses; f attached protons were deduced by the HSQC experiment.
Table 2. The NMR data for compounds 3 and 4.
Table 2. The NMR data for compounds 3 and 4.
34
C/HδH (J in Hz) aδC (mult.) bδH (J in Hz) cδC (mult.) d
12.08 m; 0.58 ddd (13.2, 13.2, 2.8) e34.1, CH2 f2.07 m; 0.56 ddd (12.0, 12.0, 3.0) e34.2, CH2 f
21.58 m17.8, CH21.56 m17.9, CH2
31.41 m; 1.17 m41.8, CH21.44 m; 1.17 m41.7, CH2
4 33.0, C 33.0, C
51.02 m56.8, CH0.99 dd (12.6, 2.4)57.0, CH
61.56 m; 1.45 m18.3, CH21.58 m; 1.52 m18.4, CH2
71.91 m; 1.18 m41.2, CH21.92 m; 1.03 m41.4, CH2
8 36.8, C 37.2, C
91.41 m53.5, CH1.33 m53.5, CH
10 41.8, C 41.8, C
112.35 m; 1.46 m21.9, CH21.91 m; 2.25 m25.1, CH2
124.97 d (2.8)76.2, CH4.96 dd (2.8, 2.8)76.6, CH
13 41.7, C 41.6, C
141.86 dd (13.2, 2.0)43.6, CH1.50 m47.5, CH
151.92 m; 2.29 m25.2, CH22.14 m; 1.50 m25.7, CH2
164.91 dd (4.0, 1.6)77.2, CH4.61 dd (8.8, 6.6)68.1, CH
17 133.6, C 132.4, C
186.73 s155.7, CH6.59 s152.4, CH
190.85 s33.8, CH30.88 s33.8, CH3
200.77 s21.9, CH30.76 s21.9, CH3
211.09 s16.6, CH31.10 s16.4, CH3
224.05 d (11.6); 3.91 dd (11.6)62.8, CH24.27 d (11.4); 3.90 dd (11.4)62.7, CH2
231.06 s19.4, CH31.20 s20.9, CH3
24 199.3, C 202.1, C
252.25 s25.8, CH32.25 s25.7, CH3
12-OAc 170.8, C 170.6, C
2.03 s21.3, CH32.04 s21.3, CH3
16-OOH9.27 s
a Spectra recorded at 400 MHz in CDCl3; b spectra recorded at 100 MHz in CDCl3; c spectra recorded at 600 MHz in CDCl3; d spectra recorded at 150 MHz in CDCl3; e J values (in Hz) in parentheses; f attached protons were deduced by the HSQC experiment.
Table 3. Experimental and calculated specific optical rotation values of 14.
Table 3. Experimental and calculated specific optical rotation values of 14.
Cald. Value aExp. Value
Exp. 1 b 21
Cald. 1-5S,8R,9S,10R,13R,14S,16S,17S31
Cald. 1-5R,8S,9R,10S,13S,14R,16R,17R−31
Exp. 2 c 61
Cald. 2-5S,8R,9S,10R,12S,13R,14S78
Cald. 2-5R,8S,9R,10S,12R,13S,14R−78
Exp. 3 d 41
Cald. 3-5S,8R,9S,10R,12S,13R,14S,16S15
Cald. 3-5R,8S,9R,10S,12R,13S,14R,16R−15
Exp. 4 e 76
Cald. 4-5S,8R,9S,10R,12S,13R,14S,16S103
Cald. 4-5R,8S,9R,10S,12R,13S,14R,16R−103
a Solvent phase in CHCl3; b   [ α ] D 25 (c 0.025, CHCl3); c   [ α ] D 25 (c 0.065, CHCl3), d   [ α ] D 25 (c 0.045, CHCl3); e   [ α ] D 25 (c 0.105, CHCl3).
Table 4. Effect of compounds 14 on superoxide anion generation and elastase release in fMLF/CB-induced human neutrophils.
Table 4. Effect of compounds 14 on superoxide anion generation and elastase release in fMLF/CB-induced human neutrophils.
CompoundsSuperoxide Anion GenerationElastase Release
IC50 (μM)Inh% IC50 (μM)Inh%
1 9.19 ± 1.06*** 1.59 ± 1.90
23.98 ± 0.6295.02 ± 3.42***5.24 ± 0.2895.47 ± 6.29***
3 12.12 ± 1.22*** 9.70 ± 1.06***
44.46 ± 0.7288.28 ± 5.25***4.73 ± 0.4098.00 ± 4.60***
LY2940022.66 ± 03988.30 ± 5.00***2.53 ± 0.5386.71 ± 6.47***
Percentage of inhibition (Inh.%) at 10 μM. The results are presented as mean ± S.E.M. (n = 3–5). *** p < 0.001 compared with the control (DMSO). LY294002 at 10 μM was used as a positive control.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Peng, B.-R.; Zheng, L.-G.; Chen, L.-Y.; El-Shazly, M.; Hwang, T.-L.; Su, J.-H.; Lee, M.-H.; Lai, K.-H.; Sung, P.-J. Nor-24-homoscalaranes, Neutrophilic Inflammatory Mediators from the Marine Sponge Lendenfeldia sp. Pharmaceuticals 2023, 16, 1258. https://doi.org/10.3390/ph16091258

AMA Style

Peng B-R, Zheng L-G, Chen L-Y, El-Shazly M, Hwang T-L, Su J-H, Lee M-H, Lai K-H, Sung P-J. Nor-24-homoscalaranes, Neutrophilic Inflammatory Mediators from the Marine Sponge Lendenfeldia sp. Pharmaceuticals. 2023; 16(9):1258. https://doi.org/10.3390/ph16091258

Chicago/Turabian Style

Peng, Bo-Rong, Li-Guo Zheng, Lo-Yun Chen, Mohamed El-Shazly, Tsong-Long Hwang, Jui-Hsin Su, Mei-Hsien Lee, Kuei-Hung Lai, and Ping-Jyun Sung. 2023. "Nor-24-homoscalaranes, Neutrophilic Inflammatory Mediators from the Marine Sponge Lendenfeldia sp." Pharmaceuticals 16, no. 9: 1258. https://doi.org/10.3390/ph16091258

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