[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (9)

Search Parameters:
Keywords = renieramycin derivatives

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
29 pages, 11368 KiB  
Article
A New Renieramycin T Right-Half Analog as a Small Molecule Degrader of STAT3
by Preeyaphan Phookphan, Satapat Racha, Masashi Yokoya, Zin Zin Ei, Daiki Hotta, Hongbin Zou and Pithi Chanvorachote
Mar. Drugs 2024, 22(8), 370; https://doi.org/10.3390/md22080370 - 14 Aug 2024
Viewed by 1830
Abstract
Constitutive activation of STAT3 contributes to tumor development and metastasis, making it a promising target for cancer therapy. (1R,4R,5S)-10-hydroxy-9-methoxy-8,11-dimethyl-3-(naphthalen-2-ylmethyl)-1,2,3,4,5,6-hexahydro-1,5-epiminobenzo[d]azocine-4-carbonitrile, DH_31, a new derivative of the marine natural product Renieramycin T, showed potent activity against H292 and H460 cells, with IC50 values of [...] Read more.
Constitutive activation of STAT3 contributes to tumor development and metastasis, making it a promising target for cancer therapy. (1R,4R,5S)-10-hydroxy-9-methoxy-8,11-dimethyl-3-(naphthalen-2-ylmethyl)-1,2,3,4,5,6-hexahydro-1,5-epiminobenzo[d]azocine-4-carbonitrile, DH_31, a new derivative of the marine natural product Renieramycin T, showed potent activity against H292 and H460 cells, with IC50 values of 5.54 ± 1.04 µM and 2.9 ± 0.58 µM, respectively. Structure–activity relationship (SAR) analysis suggests that adding a naphthalene ring with methyl linkers to ring C and a hydroxyl group to ring E enhances the cytotoxic effect of DH_31. At 1–2.5 µM, DH_31 significantly inhibited EMT phenotypes such as migration, and sensitized cells to anoikis. Consistent with the upregulation of ZO1 and the downregulation of Snail, Slug, N-cadherin, and Vimentin at both mRNA and protein levels, in silico prediction identified STAT3 as a target, validated by protein analysis showing that DH_31 significantly decreases STAT3 levels through ubiquitin-proteasomal degradation. Immunofluorescence and Western blot analysis confirmed that DH_31 significantly decreased STAT3 and EMT markers. Additionally, molecular docking suggests a covalent interaction between the cyano group of DH_31 and Cys-468 in the DNA-binding domain of STAT3 (binding affinity = −7.630 kcal/mol), leading to destabilization thereafter. In conclusion, DH_31, a novel RT derivative, demonstrates potential as a STAT3-targeting drug that significantly contribute to understanding of the development of new targeted therapy. Full article
Show Figures

Figure 1

Figure 1
<p>Derivatives of the RT right-half analogs—DH_17, DH_20, DH_23, DH_26, DH_28, DH_30, and DH_31. (<b>A</b>) The structure of Renieramycin T, TM-(−)-18, and the core structure of the RT right-half analog with R. R represents the position of the pyridyl, thiazolyl, or naphthalenyl group in ring C of the RT right-half analog, respectively. (<b>B</b>) Structures of the present RT right-half analogs: DH_17, DH_20, DH_23, DH_26, DH_28, DH_30, and DH_31. (<b>C</b>) Step-by-step synthesis for derivatives of RT right-half analogs.</p>
Full article ">Figure 2
<p>The effect of RT right-half analogs on cytotoxicity in NSCLC and human normal lung epithelial (BEAS-2B) cell lines and apoptotic cell death in NSCLC cells. (<b>A</b>) NSCLC H292 and H460 cells were treated with derivatives of RT right-half analogs for 24 h and analyzed using MTT assay to assess cytotoxicity. (<b>B</b>) IC<sub>50</sub> values for H292 and H460 cell lines were calculated. (<b>C</b>) BEAS-2B cells were treated with DH_28, DH_30, DH_31, and TM-(−)-18 for 24 h. The cytotoxic effects were evaluated using an MTT assay, and the IC<sub>50</sub> values for BEAS-2B cells were determined. (<b>D</b>) H292 and H460 cells were seeded and treated with 0–10 μM of DH_28, DH_30, and DH_31 for 24 h. Hoechst 33342 and PI were used to stain the cell nuclei. Images were obtained under a fluorescence microscope. (<b>E</b>) The percentages of cell death were calculated based on the stained images in H292 and H460 cells. Data represent the mean ± SD (<span class="html-italic">n</span> = 3). *, **, and *** indicate a statistically significant difference between the treated and the untreated cells at <span class="html-italic">p</span> &lt; 0.05, <span class="html-italic">p</span> &lt; 0.01, and <span class="html-italic">p</span> &lt; 0.001, respectively.</p>
Full article ">Figure 3
<p>Putative analysis of NSCLC against DH_31 and the effect of DH_31 on EMT-association proteins. (<b>A</b>) Venn diagram of NSCLC and DH_31 targets and GO enrichment analysis of putative targets was performed to clarify the relevant biologic processes (<span class="html-italic">p</span> &lt; 0.01). The y-axis represents GO terms, and the x-axis indicates the number of genes enriched in that term. The color from blue to red indicates the value of <span class="html-italic">p</span>. The adjust (FDR) is becoming smaller with greater credibility and importance. (<b>B</b>) The expression levels of ZO1, ZEB1, Slug, Snail, N-cadherin, and Vimentin were visualized by fluorescence microscopy. Scale bar, 20 µm. Bar graphs show the relative levels of ZO1, ZEB1, Slug, Snail, N-cadherin, and Vimentin. (<b>C</b>) The protein expression levels of ZO1, Slug, Snail, N-cadherin, Vimentin and β–actin were evaluated by Western blot analysis. The relative protein levels were calculated by densitometry. Data represent the mean ± SD (<span class="html-italic">n</span> = 3). *, **, and *** indicate a statistically significant difference between the treated and untreated cells at <span class="html-italic">p</span> &lt; 0.05, <span class="html-italic">p</span> &lt; 0.01, and <span class="html-italic">p</span> &lt; 0.001, respectively.</p>
Full article ">Figure 4
<p>The effects of DH_31 on migration and anoikis resistance on NSCLC H460. (<b>A</b>) DH_31 decreased the migration of H460 cells. (<b>B</b>) The relative migration levels of the treated and untreated cells were determined at 24, 48, and 72 h. (<b>C</b>) DH_31 increased the sensitivity to anoikis in H460 cells. (<b>D</b>) The relative viability of cells was determined after culture under detachment conditions for 6, 12, and 24 h. Scale bar, 20 µm. Data represent the mean ± SD (<span class="html-italic">n</span> = 3). ** and *** indicate a statistically significant difference between the treated and the untreated cells at <span class="html-italic">p</span> &lt; 0.01 and <span class="html-italic">p</span> &lt; 0.001, respectively.</p>
Full article ">Figure 5
<p>STAT3 identified as a potential target of DH_31. (<b>A</b>) The top 10 targets among the 64 targets were ranked based on the number of degrees, visualized by the CytoHubba plugin. The degree values of the top 10 targets in the PPI network were ranked, with STAT3 having the highest degree. The intensity of the colors corresponded to the degree values, with purple indicating large values, pink indicating moderate values, and yellow indicating small values. (<b>B</b>) H460 cells treated with DH_31 (0–2.5 μM) for 24 h were stained with anti-STAT3 antibody (red) and examined using confocal laser scanning microscopy. Cell nuclei were stained with Hoechst 33342 (blue). Scale bar, 10 µm. Arrows denote localized STAT3 proteins. (<b>C</b>) The relative levels of STAT3 of H460 were determined by immunofluorescence analysis. (<b>D</b>) The protein expression levels of STAT3 and β–actin was evaluated by Western blot analysis. (<b>E</b>) The relative protein levels were calculated by densitometry. Data represent the mean ± SD (<span class="html-italic">n</span> = 3). * and *** indicate a statistically significant difference between the treated and untreated cells at <span class="html-italic">p</span> &lt; 0.05 and <span class="html-italic">p</span> &lt; 0.001, respectively.</p>
Full article ">Figure 6
<p>The effect of DH_31 on enhanced ubiquitin-mediated STAT3 proteasomal degradation in NSCLC H460. H460 cells were treated with DH_31 (0–2.5 μM) for 8 h. (<b>A</b>) The expression levels of STAT3 mRNA were determined by Real-time qPCR. (<b>B</b>) The ubiquitin–proteasome inhibitor MG132 reversed the inhibitory effect of DH_31 on the expression of the STAT3 protein. After treatment with or without MG132 (10 µM) for 1 h, cells were treated with DH_31 (0–2.5 µM) for 6 h. The STAT3 levels were measured using Western blot analysis and calculated by densitometry. (<b>C</b>) DH_31 induced the ubiquitin–proteasomal degradation of STAT3. After treatment with or without MG132 (10 µM) for 1 h, cells were treated with DH_31 (0 and 2.5 µM) for 6 h. The protein lysates were collected subsequent to STAT3 immunoprecipitation, and the ubiquitinated protein levels were measure by Western blot analysis. Ub-STAT3 levels were calculated by densitometry. Data represent the mean ± SD (<span class="html-italic">n</span> = 3). *, and ** indicate a statistically significant difference between the treated and untreated cells at <span class="html-italic">p</span> &lt; 0.05 and <span class="html-italic">p</span> &lt; 0.01, respectively. # and ## indicate a statistically significant difference from the cells without MG132 at <span class="html-italic">p</span> &lt; 0.05 and <span class="html-italic">p</span> &lt; 0.01, respectively.</p>
Full article ">Figure 7
<p>Domain structure of STAT3 and structure of DH_31 with in silico predicted binding configurations. (<b>A</b>) Schematic of the domain structure of STAT3 and the structure of the dimer interface of STAT3 (PDB: 1BG1) illustrating the surface locations of the DNA-binding domain (residues 321–494) (red) and the SH2 domain (residues 584–688) (green), (<b>B</b>) the binding interaction of DH_31 to the SH2 domain of STAT3, (<b>C</b>) the binding interaction of DH_31 to the DNA-binding domain, and (<b>D</b>) the binding interaction of TM-(−)-18 to the DNA-binding domain of STAT3. (<b>E</b>) The binding energy of DH_31 and TM-(−)-18 at the SH2 domain and the DNA-binding domain.</p>
Full article ">Figure 8
<p>The effect of DH_31 on the mRNA expression of EMT markers in NSCLC H460. (<b>A</b>) Schematic representation of the of STAT3 transcription factor binding sites in target genes. (<b>B</b>) The mRNA expression of <span class="html-italic">ZO1</span>, <span class="html-italic">Slug</span>, <span class="html-italic">Snail</span>, <span class="html-italic">N-cadherin</span>, and <span class="html-italic">Vimentin</span> in H460 cells treated with DH_31 (0–2.5 µM).</p>
Full article ">Scheme 1
<p>Synthesis of <b>2e</b>.</p>
Full article ">Scheme 2
<p>Synthesis of <b>2f</b>: DH_30.</p>
Full article ">Scheme 3
<p>Synthesis of <b>3a</b>: DH_17.</p>
Full article ">Scheme 4
<p>Synthesis of <b>3b</b>: DH_20.</p>
Full article ">Scheme 5
<p>Synthesis of <b>3c</b>: DH_23.</p>
Full article ">Scheme 6
<p>Synthesis of <b>3d</b>: DH_26.</p>
Full article ">Scheme 7
<p>Synthesis of <b>3e</b>: DH_28.</p>
Full article ">Scheme 8
<p>Synthesis of <b>3f</b>: DH_31.</p>
Full article ">
32 pages, 49487 KiB  
Article
Simplified Synthesis of Renieramycin T Derivatives to Target Cancer Stem Cells via β-Catenin Proteasomal Degradation in Human Lung Cancer
by Zin Zin Ei, Satapat Racha, Masashi Yokoya, Daiki Hotta, Hongbin Zou and Pithi Chanvorachote
Mar. Drugs 2023, 21(12), 627; https://doi.org/10.3390/md21120627 - 30 Nov 2023
Cited by 1 | Viewed by 2033
Abstract
Cancer stem cells (CSCs) found within cancer tissue play a pivotal role in its resistance to therapy and its potential to metastasize, contributing to elevated mortality rates among patients. Significant strides in understanding the molecular foundations of CSCs have led to preclinical investigations [...] Read more.
Cancer stem cells (CSCs) found within cancer tissue play a pivotal role in its resistance to therapy and its potential to metastasize, contributing to elevated mortality rates among patients. Significant strides in understanding the molecular foundations of CSCs have led to preclinical investigations and clinical trials focused on CSC regulator β-catenin signaling targeted interventions in malignancies. As part of the ongoing advancements in marine-organism-derived compound development, it was observed that among the six analogs of Renieramycin T (RT), a potential lead alkaloid from the blue sponge Xestospongia sp., the compound DH_32, displayed the most robust anti-cancer activity in lung cancer A549, H23, and H292 cells. In various lung cancer cell lines, DH_32 exhibited the highest efficacy, with IC50 values of 4.06 ± 0.24 μM, 2.07 ± 0.11 μM, and 1.46 ± 0.06 μM in A549, H23, and H292 cells, respectively. In contrast, parental RT compounds had IC50 values of 5.76 ± 0.23 μM, 2.93 ± 0.07 μM, and 1.52 ± 0.05 μM in the same order. Furthermore, at a dosage of 25 nM, DH_32 showed a stronger ability to inhibit colony formation compared to the lead compound, RT. DH_32 was capable of inducing apoptosis in lung cancer cells, as demonstrated by increased PARP cleavage and reduced levels of the proapoptotic protein Bcl2. Our discovery confirms that DH_32 treatment of lung cancer cells led to a reduced level of CD133, which is associated with the suppression of stem-cell-related transcription factors like OCT4. Moreover, DH_32 significantly suppressed the ability of tumor spheroids to form compared to the original RT compound. Additionally, DH_32 inhibited CSCs by promoting the degradation of β-catenin through ubiquitin–proteasomal pathways. In computational molecular docking, a high-affinity interaction was observed between DH_32 (grid score = −35.559 kcal/mol) and β-catenin, indicating a stronger binding interaction compared to the reference compound R9Q (grid score = −29.044 kcal/mol). In summary, DH_32, a newly developed derivative of the right-half analog of RT, effectively inhibited the initiation of lung cancer spheroids and the self-renewal of lung cancer cells through the upstream process of β-catenin ubiquitin–proteasomal degradation. Full article
(This article belongs to the Section Synthesis and Medicinal Chemistry of Marine Natural Products)
Show Figures

Figure 1

Figure 1
<p>Derivatives of right-half RT—DH_18, DH_21, DH_32, DH_35, DH_38, and DH_39. (<b>A</b>) Step-by-step synthesis for right-half RT derivatives (DH_18, DH_21, DH_32, DH_35, DH_38, and DH_39). (<b>B</b>) Structures of DH_18, DH_21, DH_32, DH_35, DH_38, and DH_39.</p>
Full article ">Figure 2
<p>Screening for right-half RT derivatives—DH_18, DH_21, DH_32, DH_35, DH_38, and DH_39—on cell viability and apoptosis in lung cancer cells (A549, H23, and H292). (<b>A</b>) Human lung cancer cells were seeded and treated with right-half RT derivatives (0–100 μM) for 24 h. An MTT assay was performed to evaluate IC<sub>50</sub> values for right-half RT derivatives used to treat human lung cancer cells. The IC<sub>50</sub> values were calculated in comparison to the RT parent compound, which acted as a positive control. (<b>B</b>) NSCLC cells were seeded and treated with right-half RT derivatives (0.5 μM) for 24 h. The apoptosis and dead cells were evaluated by co-staining with Hoechst 33342 and PI. The images were captured with a fluorescence microscope, and the percentages for apoptosis and dead cells were calculated. RT served as a positive control. (<b>C</b>) Human dermal papilla cells (DP) and a non-tumorigenic epithelial cell line from human bronchial epithelium cells (BEAS2B) were seeded and treated with DH_32 (0–100 μM) for 24 h. An MTT assay was performed to evaluate IC<sub>50</sub> values for DH_32-treated human dermal papilla cells (DP) and non-tumorigenic epithelial cell line from human bronchial epithelium cells (BEAS2B) cells. The IC<sub>50</sub> values were calculated in comparison to the RT parent compound, which acted as a positive control. Data are presented as mean ± SD (<span class="html-italic">n</span> = 3). Significance is shown as *** <span class="html-italic">p</span> &lt; 0.001 versus untreated control cells.</p>
Full article ">Figure 3
<p>DH_32 demonstrated antiproliferative effects, inhibiting colony formation and inducing apoptosis. RT served as the positive control. (<b>A</b>) The proliferative effect of DH_32 on lung cancer cells was evaluated by an MTT assay for 24, 48, and 72 h, analyzed as the relative value to the control group at 0 h. (<b>B</b>) NSCLC cells were seeded and treated with DH_32 (0–100 nM) for 24 h. The colonies were allowed to grow for 7 days and stained with crystal violet to count the number of colony formations. (<b>C</b>) Human lung cancer cells were treated with DH_32 (0–100 nM) for 24 h. The levels of apoptotic-related proteins, including cleaved PARP, Bcl-2, and Bax, were determined by Western blot analysis. The blot was reprobed with β-actin to confirm the equal loading of proteins. Data are presented as mean ± SD (<span class="html-italic">n</span> = 3). Significant is shown as ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 versus untreated control cells and <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 versus RT-treated NSCLCs.</p>
Full article ">Figure 4
<p>The suppressive effect of DH_32 on the CSC-like phenotype in human lung cancer A549, H23, and H292 cells. Cells were seeded and cultured to form primary spheroids for 7 days. The primary spheroids were trypsinized to form secondary spheroids for 10 days in an ultra-low attachment plate to obtain CSC-rich spheroids. After that, the spheroids were treated with DH_32 (50 nM) for 24 h. The DH_32-treated lung cancer cells were analyzed for the level of stem cell markers CD133, CD44, and transcription factor OCT4. The nucleus of cells was stained with Hoechst33342. (<b>A</b>) The expression level of CD133; (<b>B</b>) the expression level of CD44 and OCT4. The parent compound RT (50 nM) and cisplatin (50 nM) acted as the positive controls. Scale bar = 100 μm.</p>
Full article ">Figure 5
<p>Effects of DH_32 on the mRNA and protein levels of stem cell markers (CD133, CD44, and ALDH1A1) and stem cell transcription factors (OCT4, NANOG, and SOX2). (<b>A</b>) Cells were treated with DH_32 (0–100 nM) for 6 h, and the mRNA expression levels of stem cell transcription factors OCT4, NANOG, and SOX2 were determined. The mRNA levels were normalized by housekeeping GAPDH. The relative mRNA expression was calculated by using comparative Ct cycles. (<b>B</b>) The heat map shows the fluorescence intensity of stem cell markers (CD133, CD44, and ALDH1A1) and stem cell transcription factors (OCT4, NANOG, and SOX2) captured by a fluorescence microscope; the fluorescence intensity was determined by Image J software version 1.52a. (<b>C</b>) The protein expression levels of stem cell markers (CD133, CD44, and ALDH1A1) and stem cell transcription factors (OCT4) were determined by Western blot analysis. The blot was reprobed with β-actin to confirm the equal loading of proteins. Data are presented as mean ± SD (<span class="html-italic">n</span> = 3). Significance is shown as * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001 versus untreated control cells.</p>
Full article ">Figure 6
<p>DH_32 suppresses CSC properties via β-catenin proteasomal degradation in human lung cancer cells. The cancer cells were treated with various concentrations of DH_32 (0–100 nM) for 24 h. (<b>A</b>) The β-catenin level was measured by immunofluorescence analysis, and the fluorescence intensity was measured by Image J software. (<b>B</b>) The protein expression level of β-catenin was evaluated by Western blot analysis. The blot was reprobed with β-actin to confirm the equal loading of proteins. The blots were quantified by densitometry by Image J software. (<b>C</b>) The effect of DH_32 on the ubiquitin–proteasomal degradation of β-catenin in human lung cancer cells was measured by immunoprecipitation analysis. The human lung cancer cells were treated with DH_32 (50 nM). The cell lysates were prepared and immunoprecipitated with anti-β-catenin. After that, ubiquitinated protein levels were measured by Western blotting by using an anti-ubiquitin antibody. The ubiquitin–β-catenin level was analyzed by densitometry. The parent compound, RT, acted as a positive control. Data are presented as mean ± SD (<span class="html-italic">n</span> = 3). Significant is shown as * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001 versus untreated control cells and <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 and <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 versus RT-treated NSCLCs.</p>
Full article ">Figure 7
<p>Binding interaction of DH_32 with GSK-3β. (<b>A</b>) The native ligand (PF-04802367) superposition represented by crystal (tan) and flexible conformations (blue) (<b>B</b>) X-ray co-crystal structure of PF-04802367 (reference compound) bound in the ATP-binding site of GSK-3β. The green dashed lines denote hydrogen-bonding interactions. (<b>C</b>) X-ray co-crystal structure of DH_32 bound in the ATP-binding site of GSK-3β. (<b>D</b>) Footprint analysis for DH_32 (red lines) compared to PF-04802367 (reference compound) (blue lines) into the ATP-binding site of GSK-3β.</p>
Full article ">Figure 8
<p>Binding interaction of DH_32 with β-catenin (<b>A</b>) The native ligand (R9Q) superposition represented by crystal (tan) and flexible conformations (blue) (<b>B</b>) X-ray co-crystal structure of R9Q (reference compound) bound in the binding site of β-catenin. The green dashed lines denote hydrogen-bonding interactions. (<b>C</b>) X-ray co-crystal structure of DH_32 bound in the binding site of β-catenin. (<b>D</b>) Footprint analysis for DH_32 (red lines) compared to the R9Q (reference compound) (blue lines) into the binding site of β-catenin.</p>
Full article ">Figure 9
<p>Summarized figure for the effect of direction interactions of the right-half RT analog, DH_32, with targeted β-catenin in lung cancer cells. This study revealed that DH_32 has the ability to reduce stem cell markers (CD133, CD44, and ALDH1A1) and stem cell transcription factors (OCT4, NANOG, and SOX2) in lung cancer cells. Moreover, the inhibitory impacts on the CSC characteristics and stem cell transcription factors achieved by DH_32 are mediated through the proteasomal degradation of β-catenin.</p>
Full article ">Scheme 1
<p>Synthesis of (1<span class="html-italic">R</span>,4<span class="html-italic">R</span>,5<span class="html-italic">S</span>)-9-methoxy-8,11-dimethyl-7,10-dioxo-3-(pyridin-2-ylmethyl)-1,2,3,4,5,6,7,10-octahydro-1,5-epiminobenzo[d]azocine-4-carbonitrile (<b>4a</b>: DH_18).</p>
Full article ">Scheme 2
<p>Synthesis of (1<span class="html-italic">R</span>,4<span class="html-italic">R</span>,5<span class="html-italic">S</span>)-9-methoxy-8,11-dimethyl-7,10-dioxo-3-(pyridin-3-ylmethyl)-1,2,3,4,5,6,7,10-octahydro-1,5-epiminobenzo[d]azocine-4-carbonitrile (<b>4b</b>: DH_21).</p>
Full article ">Scheme 3
<p>Synthesis of (1<span class="html-italic">R</span>,4<span class="html-italic">R</span>,5<span class="html-italic">S</span>)-9-methoxy-8,11-dimethyl-3-(naphthalen-2-ylmethyl)-7,10-dioxo-1,2,3,4,5,6,7,10-octahydro-1,5-epiminobenzo[d]azocine-4-carbonitrile (<b>4c</b>: DH_32).</p>
Full article ">Scheme 4
<p>Synthesis of (1<span class="html-italic">R</span>,4<span class="html-italic">R</span>,5S)-10-(benzyloxy)-9-methoxy-8,11-dimethyl-3-(prop-2-yn-1-yl)-1,2,3,4,5,6-hexahydro-1,5-epiminobenzo[d]azocine-4-carbonitrile (<b>2d</b>).</p>
Full article ">Scheme 5
<p>Synthesis of (1R,4R,5S)-10-hydroxy-9-methoxy-8,11-dimethyl-3-(prop-2-yn-1-yl)-1,2,3,4,5,6-hexahydro-1,5-epiminobenzo[d]azocine-4-carbonitrile (<b>3d</b>).</p>
Full article ">Scheme 6
<p>Synthesis of (1<span class="html-italic">R</span>,4<span class="html-italic">R</span>,5<span class="html-italic">S</span>)-9-methoxy-8,11-dimethyl-7,10-dioxo-3-(prop-2-yn-1-yl)-1,2,3,4,5,6,7,10-octahydro-1,5-epiminobenzo[d]azocine-4-carbonitrile (<b>4d</b>: DH_35).</p>
Full article ">Scheme 7
<p>Synthesis of (1<span class="html-italic">R</span>,4<span class="html-italic">R</span>,5<span class="html-italic">S</span>)-10-(benzyloxy)-3-(2-hydroxyethyl)-9-methoxy-8,11-dimethyl-1,2,3,4,5,6-hexahydro-1,5-epiminobenzo[d]azocine-4-carbonitrile (<b>2e</b>).</p>
Full article ">Scheme 8
<p>Synthesis of (1<span class="html-italic">R</span>,4<span class="html-italic">R</span>,5<span class="html-italic">S</span>)-10-hydroxy-3-(2-hydroxyethyl)-9-methoxy-8,11-dimethyl-1,2,3,4,5,6-hexahydro-1,5-epiminobenzo[d]azocine-4-carbonitrile (<b>3e</b>).</p>
Full article ">Scheme 9
<p>Synthesis of (1<span class="html-italic">R</span>,4<span class="html-italic">R</span>,5<span class="html-italic">S</span>)-3-(2-hydroxyethyl)-9-methoxy-8,11-dimethyl-7,10-dioxo-1,2,3,4,5,6,7,10-octahydro-1,5-epiminobenzo[d]azocine-4-carbonitrile (<b>4e</b>: DH_38).</p>
Full article ">Scheme 10
<p>Synthesis of (1<span class="html-italic">R</span>,4<span class="html-italic">R</span>,5<span class="html-italic">S</span>)-3-(2-{1,3-dioxoisoindolin-2-yl} ethyl)-9-methoxy-8,11-dimethyl-7,10-dioxo-1,2,3,4,5,6,7,10-octahydro-1,5-epiminobenzo[d]azocine-4-carbonitrile (<b>4f</b>: DH_39).</p>
Full article ">
17 pages, 3851 KiB  
Article
Light-Mediated Transformation of Renieramycins and Semisynthesis of 4′-Pyridinecarbonyl-Substituted Renieramycin-Type Derivatives as Potential Cytotoxic Agents against Non-Small-Cell Lung Cancer Cells
by Suwimon Sinsook, Koonchira Buaban, Iksen Iksen, Korrakod Petsri, Bhurichaya Innets, Chaisak Chansriniyom, Khanit Suwanborirux, Masashi Yokoya, Naoki Saito, Varisa Pongrakhananon, Pithi Chanvorachote and Supakarn Chamni
Mar. Drugs 2023, 21(7), 400; https://doi.org/10.3390/md21070400 - 13 Jul 2023
Viewed by 2096
Abstract
The semisynthesis of renieramycin-type derivatives was achieved under mild and facile conditions by attaching a 1,3-dioxole-bridged phenolic moiety onto ring A of the renieramycin structure and adding a 4′-pyridinecarbonyl ester substituent at its C-5 or C-22 position. These were accomplished through a light-induced [...] Read more.
The semisynthesis of renieramycin-type derivatives was achieved under mild and facile conditions by attaching a 1,3-dioxole-bridged phenolic moiety onto ring A of the renieramycin structure and adding a 4′-pyridinecarbonyl ester substituent at its C-5 or C-22 position. These were accomplished through a light-induced intramolecular photoredox reaction using blue light (4 W) and Steglich esterification, respectively. Renieramycin M (4), a bis-tetrahydroisoquinolinequinone compound isolated from the Thai blue sponge (Xestospongia sp.), served as the starting material. The cytotoxicity of the 10 natural and semisynthesized renieramycins against non-small-cell lung cancer (NSCLC) cell lines was evaluated. The 5-O-(4′-pyridinecarbonyl) renieramycin T (11) compound exhibited high cytotoxicity with half-maximal inhibitory concentration (IC50) values of 35.27 ± 1.09 and 34.77 ± 2.19 nM against H290 and H460 cells, respectively. Notably, the potency of compound 11 was 2-fold more than that of renieramycin T (7) and equal to those of 4 and doxorubicin. Interestingly, the renieramycin-type derivatives with a hydroxyl group at C-5 and C-22 exhibited weak cytotoxicity. In silico molecular docking and dynamics studies confirmed that the mitogen-activated proteins, kinase 1 and 3 (MAPK1 and MAPK3), are suitable targets for 11. Thus, the structure–cytotoxicity study of renieramycins was extended to facilitate the development of potential anticancer agents for NSCLC cells. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Structures of the bis-tetrahydroisoquinoline alkaloids as anticancer drugs and promising drug leads.</p>
Full article ">Figure 2
<p>Semisynthetic derivatives of the renieramycin-type derivatives containing a 4′-pyridinecarbonyl ester substituent.</p>
Full article ">Figure 3
<p>HMBCs (blue arrows) for the 4′-pyridinecarbonyl-substituted derivatives (<b>11</b> and <b>12</b>).</p>
Full article ">Figure 4
<p>Structure and putative analysis of <b>10</b> and <b>11</b> against NSCLC. (<b>A</b>) Structure of <b>10</b>. (<b>B</b>) Structure of <b>11</b>. (<b>C</b>) Venn diagram analysis showing both <b>10</b> and <b>11</b> associated with 116 targets, whereas the NSCLC cells are associated with 6431 targets. The interception analysis showing 94 overlapping targets between <b>10</b>/<b>11</b> and NSCLC.</p>
Full article ">Figure 5
<p>Protein–protein interaction (PPI). (<b>A</b>) The main cluster of the PPI obtained from the 94 overlapping targets and visualized using Cytoscape 3.9.1. (<b>B</b>) Top 10 targets based on the number of degrees visualized by the CytoHubba plugin. More than 10 degrees are displayed for all the main targets. The intensity of the colors represents the degree values, where red, orange, and yellow correspond to large, medium, and small values, respectively. Red arrow indicates the MAPK family.</p>
Full article ">Figure 6
<p>Molecular docking of <b>10</b> and <b>11</b> with both MAPK1 (ERK2) and MAPK3 (ERK1). (<b>A</b>) Binding energies indicate that the interactions of <b>10</b> and <b>11</b> with both MAPK1 and MAPK3, compared to Ravoxertinib, a known inhibitor. (<b>B</b>) Interactions between the amino acid residues of MAPK1 and both <b>10</b> and <b>11</b>. (<b>C</b>) Interactions between the amino acid residues of MAPK3 and both <b>10</b> and <b>11</b>.</p>
Full article ">Figure 7
<p>Molecular dynamics study of the interactions between <b>11</b> and MAPK1 and MAPK3. Green line shows the interaction with MAPK1, whereas the red line shows the interaction with MAPK3. (<b>A</b>) The ligand conformation and (<b>B</b>) ligand movement interaction of <b>11</b> with MAPK1 and MAPK3.</p>
Full article ">Scheme 1
<p>Photoredox reaction of the natural renieramycins.</p>
Full article ">Scheme 2
<p>Proposed transformation mechanism of <b>5</b> into <b>6</b> and <b>8</b>.</p>
Full article ">Scheme 3
<p>Semisynthesis of the 4′-pyridinecarbonyl-substituted renieramycin-type derivatives.</p>
Full article ">
21 pages, 4009 KiB  
Article
Novel Synthetic Derivative of Renieramycin T Right-Half Analog Induces Apoptosis and Inhibits Cancer Stem Cells via Targeting the Akt Signal in Lung Cancer Cells
by Korrakod Petsri, Masashi Yokoya, Satapat Racha, Sunisa Thongsom, Chorpaka Thepthanee, Bhurichaya Innets, Zin Zin Ei, Daiki Hotta, Hongbin Zou and Pithi Chanvorachote
Int. J. Mol. Sci. 2023, 24(6), 5345; https://doi.org/10.3390/ijms24065345 - 10 Mar 2023
Cited by 4 | Viewed by 1787
Abstract
Akt is a key regulatory protein of cancer stem cells (CSCs) and is responsible for cancer aggressiveness and metastasis. Targeting Akt is beneficial for the development of cancer drugs. renieramycin T (RT) has been reported to have Mcl-1 targeting activity, and the study [...] Read more.
Akt is a key regulatory protein of cancer stem cells (CSCs) and is responsible for cancer aggressiveness and metastasis. Targeting Akt is beneficial for the development of cancer drugs. renieramycin T (RT) has been reported to have Mcl-1 targeting activity, and the study of the structure-activity relationships (SARs) demonstrated that cyanide and the benzene ring are essential for its effects. In this study, novel derivatives of the RT right-half analog with cyanide and the modified ring were synthesized to further investigate the SARs for improving the anticancer effects of RT analogs and evaluate CSC-suppressing activity through Akt inhibition. Among the five derivatives, a compound with a substituted thiazole structure (DH_25) exerts the most potent anticancer activity in lung cancer cells. It has the ability to induce apoptosis, which is accompanied by an increase in PARP cleavage, a decrease in Bcl-2, and a diminishment of Mcl-1, suggesting that residual Mcl-1 inhibitory effects exist even after modifying the benzene ring to thiazole. Furthermore, DH_25 is found to induce CSC death, as well as a decrease in CSC marker CD133, CSC transcription factor Nanog, and CSC-related oncoprotein c-Myc. Notably, an upstream member of these proteins, Akt and p-Akt, are also downregulated, indicating that Akt can be a potential target of action. Computational molecular docking showing a high-affinity interaction between DH_25 and an Akt at the allosteric binding site supports that DH_25 can bind and inhibit Akt. This study has revealed a novel SAR and CSC inhibitory effect of DH_25 via Akt inhibition, which may encourage further development of RT compounds for cancer therapy. Full article
(This article belongs to the Section Molecular Pharmacology)
Show Figures

Figure 1

Figure 1
<p>Derivatives of RT right-half analogs (DH_16, DH_19, DH_22, DH_24, DH_25, and TM-(–)-4a) induce cytotoxicity in NSCLC cells (H460 and H23). (<b>A</b>) Structures of DH_16, DH_19, DH_22, DH_24, and DH_25. (<b>B</b>) NSCLC cells were seeded and treated with 0–100 µM of DH_16, DH_19, DH_22, DH_24, and DH_25 for 24 h. Then, the MTT assay was performed to evaluate cell viability. IC<sub>50</sub> was calculated in comparison with the untreated control. (<b>C</b>) Structures and the cytotoxicity profile of TM-(–)-4a in H460 cells. The IC50 was calculated in comparison with the untreated control. Data are represented as the mean ± SD (<span class="html-italic">n</span> = 3) (* 0.01 ≤ <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, compared with the untreated control).</p>
Full article ">Figure 2
<p>Derivatives of the RT right-half analog (DH_16, DH_19, DH_22, DH_24, and DH_25) induce apoptotic morphological changes in NSCLC cells (H460 and H23). (<b>A</b>,<b>B</b>) NSCLC cells were seeded and treated with 0–50 µM of DH_16, DH_19, DH_22, DH_24, and DH_25 for 24 h. Hoechst 33342 and PI were used to stain the cell nuclei. Images were obtained under a fluorescence microscope, and the percentages of apoptotic cells were calculated. Data are presented as the mean ± SD (<span class="html-italic">n</span> = 3) (* 0.01 ≤ <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, compared with the untreated control).</p>
Full article ">Figure 3
<p>DH-25 inhibits colony formation and induces apoptosis in NSCLC cells (H460 and H23). (<b>A</b>) NSCLC cells were seeded and treated with 0–10 µM of DH_25 for 24 h. The cells were then cultured for seven days before being stained with crystal violet to count the colony number. (<b>B</b>) NSCLC cells (H460 and H23) were seeded and treated with 0–50 µM of DH_25 for 24 h. Flow cytometry was used to detect apoptotic cells using annexin V-FITC/PI staining. The percentages of cells in each stage were calculated. (<b>C</b>) NSCLC cells (H460 and H23) were seeded and treated with 0–10 µM of DH_25 for 24 h. Western blot analysis was performed to detect the protein levels of PARP, cleaved PARP, Mcl-1, Bcl-2, and Bax. The blots were reprobed with β-actin to confirm an equal loading. Densitometry was used to calculate protein expression levels, and results were presented as fold changes relative to the uncleaved form, or β-actin. The uncropped blotting bands are presented in <a href="#app1-ijms-24-05345" class="html-app">Figure S6</a>. Data are presented as the mean ± SD (<span class="html-italic">n</span> = 3) (* 0.01 ≤ <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, compared with the untreated control).</p>
Full article ">Figure 4
<p>DH-25 suppresses CSC spheroid formation and CSC signals in NSCLC cells (H460 and H23). (<b>A</b>) NSCLC cells (H460 and H23) were seeded and allowed to form primary spheroids for 7 days. The primary spheroids were suspended into single cells to form CSC-rich spheroids for 14 days in ultralow-attachment 96-well plates. The CSC-rich spheroids were then treated with 0–10 µM of DH_25 for 3 days. Hoechst 33342 and PI were used to stain the cell nuclei. Images were visualized using a fluorescence microscope. The relative size of CSC spheroids was quantified. (<b>B</b>) NSCLC cells (H460 and H23) were seeded and treated with 0–10 µM of DH_25 for 24 h. Western blot analysis was performed to detect the protein levels of CD133, Akt, p-Akt, c-Myc, and Nanog. The blots were reprobed with β-actin to confirm an equal loading. Densitometry was used to calculate protein expression levels, and results were presented as fold changes relative to the uncleaved form, or β-actin. Uncropped blotting bands are presented in <a href="#app1-ijms-24-05345" class="html-app">Figure S7</a>. Data are presented as the mean ± SD (<span class="html-italic">n</span> = 3) (* 0.01 ≤ <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, compared with the untreated control).</p>
Full article ">Figure 5
<p>Molecular docking of DH_25 and Akt at the ATP-binding site compared with the ATP-competitive inhibitor (capivasertib). (<b>A</b>) Structural superimposition of redocked (blue) and experimental native ligands (green) at the active site of the ATP-binding pocket. (<b>B</b>) Binding energies of the ligand in complex with the ATP-binding site of Akt. (<b>C</b>) The ATP-binding site of Akt complexed with DH_25 or capivasertib, for reference. The yellow dashed lines denote hydrogen-bonding interactions. (<b>D</b>) Footprint analysis for DH_25 (red lines) compared with the capivasertib reference (blue lines) into the ATP-binding site of Akt.</p>
Full article ">Figure 6
<p>Molecular docking of DH_25 and Akt at the allosteric binding site in comparison to miransertib, an Akt inhibitor. (<b>A</b>) Structural superimposition of redocked (blue) and experimental native ligands (green) at the allosteric binding site. (<b>B</b>) Binding energies of the ligand in complex with the allosteric binding site of Akt. (<b>C</b>) The allosteric binding site of Akt in complex with DH_25 or miransertib for reference. The yellow dashed lines denote hydrogen-bonding interactions. (<b>D</b>) Footprint analysis for DH_25 (red lines) compared with the miransertib reference (blue lines) into the allosteric binding site of Akt.</p>
Full article ">Figure 7
<p>Molecular docking of DH_25 and Akt at the PH domain compared with PIP3. (<b>A</b>) Structural superimposition of redocked (blue) and experimental native ligand (green) at the PH domain. (<b>B</b>) Binding energies of the ligand in complex with the PH domain of Akt. (<b>C</b>) The PH domain of Akt in complex with DH_25 or PIP3 for reference. The yellow dashed lines denote hydrogen-bonding interactions. (<b>D</b>) Footprint analysis for DH_25 (red lines) compared with the PIP3 reference (blue lines) into the PH domain of Akt.</p>
Full article ">Figure 8
<p>DH-25 regulates Nanog and CD133 through p-Akt inhibition in NSCLC cells (H460). NSCLC cells (H460) were seeded and treated with 0–10 µM of DH_25 for 24 h. In the conditions that LY294002 (5 µM) was used, it was added to pretreat the cells for 0.5 h before treatment with DH_25. Western blot analysis was performed to detect the protein levels of p-Akt, Nanog, and CD133. The blots were reprobed with β-actin to confirm an equal loading. Densitometry was used to calculate protein expression levels, and results were presented as fold changes relative to the uncleaved form, or β-actin. Uncropped blotting bands are presented in <a href="#app1-ijms-24-05345" class="html-app">Figure S8</a>. Data are presented as the mean ± SD (<span class="html-italic">n</span> = 3) (* 0.01 ≤ <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, compared with the untreated control, and # 0.01 ≤ <span class="html-italic">p</span> &lt; 0.05, compared with the treatment alone).</p>
Full article ">
14 pages, 1249 KiB  
Article
Chemistry of Renieramycins. Part 19: Semi-Syntheses of 22-O-Amino Ester and Hydroquinone 5-O-Amino Ester Derivatives of Renieramycin M and Their Cytotoxicity against Non-Small-Cell Lung Cancer Cell Lines
by Supakarn Chamni, Natchanun Sirimangkalakitti, Pithi Chanvorachote, Khanit Suwanborirux and Naoki Saito
Mar. Drugs 2020, 18(8), 418; https://doi.org/10.3390/md18080418 - 10 Aug 2020
Cited by 13 | Viewed by 3196
Abstract
Two new series of synthetic renieramycins including 22-O-amino ester and hydroquinone 5-O-amino ester derivatives of renieramycin M were semi-synthesized and evaluated for their cytotoxicity against the metastatic non-small-cell lung cancer H292 and H460 cell lines. Interestingly, the series of [...] Read more.
Two new series of synthetic renieramycins including 22-O-amino ester and hydroquinone 5-O-amino ester derivatives of renieramycin M were semi-synthesized and evaluated for their cytotoxicity against the metastatic non-small-cell lung cancer H292 and H460 cell lines. Interestingly, the series of 22-O-amino ester derivatives displayed a potent cytotoxic activity greater than the hydroquinone derivatives. The most cytotoxic derivative of the series was the 22-O-(N-Boc-l-glycine) ester of renieramycin M (5a: IC50 3.56 nM), which showed 7-fold higher potency than renieramycin M (IC50 24.56 nM) and 61-fold more than jorunnamycin A (IC50 217.43 nM) against H292 cells. In addition, 5a exhibited a significantly higher cytotoxic activity than doxorubicin (ca. 100 times). The new semi-synthetic renieramycin derivatives will be further studied and developed as potential cytotoxic agents for non-small-cell lung cancer treatment. Full article
(This article belongs to the Special Issue Chemical Modification of Marine Natural Products)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The Thai blue sponge <span class="html-italic">Xestospongia</span> sp., renieramycin M (<b>1</b>) and its semi-synthetic derivatives (<b>2</b>,<b>3</b>) with highly potent cytotoxicity.</p>
Full article ">Figure 2
<p>Ecteinascidin 743 as a template of the new series of 22-<span class="html-italic">O</span>-amino ester and hydroquinone 5-<span class="html-italic">O</span>-amino ester derivatives of <b>1</b>.</p>
Full article ">Scheme 1
<p>Semi-synthesis of 22-<span class="html-italic">O</span>-amino ester derivatives of <b>1</b> (<b>5a</b>–<b>5e</b>).</p>
Full article ">Scheme 2
<p>Semi-synthesis of hydroquinone 5-<span class="html-italic">O</span>-amino ester derivatives of <b>1</b> (<b>6a</b>–<b>6e</b>).</p>
Full article ">
25 pages, 19676 KiB  
Article
Structure–Activity Relationships and Molecular Docking Analysis of Mcl-1 Targeting Renieramycin T Analogues in Patient-derived Lung Cancer Cells
by Korrakod Petsri, Masashi Yokoya, Sucharat Tungsukruthai, Thanyada Rungrotmongkol, Bodee Nutho, Chanida Vinayanuwattikun, Naoki Saito, Takehiro Matsubara, Ryo Sato and Pithi Chanvorachote
Cancers 2020, 12(4), 875; https://doi.org/10.3390/cancers12040875 - 3 Apr 2020
Cited by 15 | Viewed by 3894
Abstract
Myeloid cell leukemia 1 (Mcl-1) and B-cell lymphoma 2 (Bcl-2) proteins are promising targets for cancer therapy. Here, we investigated the structure–activity relationships (SARs) and performed molecular docking analysis of renieramycin T (RT) and its analogues and identified the critical functional groups of [...] Read more.
Myeloid cell leukemia 1 (Mcl-1) and B-cell lymphoma 2 (Bcl-2) proteins are promising targets for cancer therapy. Here, we investigated the structure–activity relationships (SARs) and performed molecular docking analysis of renieramycin T (RT) and its analogues and identified the critical functional groups of Mcl-1 targeting. RT have a potent anti-cancer activity against several lung cancer cells and drug-resistant primary cancer cells. RT mediated apoptosis through Mcl-1 suppression and it also reduced the level of Bcl-2 in primary cells. For SAR study, five analogues of RT were synthesized and tested for their anti-cancer and Mcl-1- and Bcl-2-targeting effects. Only two of them (TM-(–)-18 and TM-(–)-4a) exerted anti-cancer activities with the loss of Mcl-1 and partly reduced Bcl-2, while the other analogues had no such effects. Specific cyanide and benzene ring parts of RT’s structure were identified to be critical for its Mcl-1-targeting activity. Computational molecular docking indicated that RT, TM-(–)-18, and TM-(–)-4a bound to Mcl-1 with high affinity, whereas TM-(–)-45, a compound with a benzene ring but no cyanide for comparison, showed the lowest binding affinity. As Mcl-1 helps cancer cells evading apoptosis, these data encourage further development of RT compounds as well as the design of novel drugs for treating Mcl-1-driven cancers. Full article
Show Figures

Figure 1

Figure 1
<p>Effects of renieramycin T (RT) on cell viability and apoptotic cell death in non-small cell lung cancer (NSCLC) cell lines (H460, H292, H23, and A549) and patient-derived primary cancer cell lines (ELC12, ELC16, ELC17, and ELC20). (<b>a</b>) The structure of RT. (<b>b</b>) The morphology of NSCLC and patient-derived primary cancer cell lines and their molecular characteristics. (<b>c</b>) H460, ELC12, ELC16, ELC17, and ELC20 cells were seeded and treated with 0–25 μM of RT or chemotherapeutic drugs (cisplatin, etoposide, and doxorubicin) for 24 h. Then, the MTT assay was performed to determine the percentages of cell viability. (<b>d</b>) The IC<sub>50</sub> in all cells was calculated in comparison to the untreated control. (<b>e–f</b>) Cells were seeded and treated with 0–10 μM of RT or chemotherapeutic drugs (cisplatin, etoposide, and doxorubicin) for 24 h before adding Hoechst 33342 and PI to stain the cell nucleuses. Images were detected by using a fluorescence microscope and the percentages of nuclear-fragmented and propidium iodide (PI)-positive cells were calculated. (<b>g</b>) ELC12, ELC16, ELC17, and ELC20 cells were treated with 0–1 μM of RT for 24 h. Western blot analysis was performed to detect the PARP and cleaved PARP protein levels. The blots were reprobed with β-actin to confirm an equal loading of each of the protein samples and densitometry was used to calculate the protein expression levels. Densitometric values of protein levels were presented as the fold changes relative to uncleaved form of the protein. Data represent the mean ± SEM (<span class="html-italic">n</span> = 3) (* 0.01 ≤ <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, compared with the untreated control). Gene symbols: KRAS (Kirsten Rat Sarcoma Viral Oncogene Homolog), PI3KCA (Phosphatidylinositol-4, 5-Bisphosphate 3-Kinase Catalytic Subunit Alpha), CDKN2A (Cyclin Dependent Kinase Inhibitor 2A), EGFR (Epidermal Growth Factor Receptor), ALK (Anaplastic Lymphoma Receptor Tyrosine Kinase). Abbreviations: N/A (not available), del (deletion), WT (wild type), Rx (treatment).</p>
Full article ">Figure 2
<p>Cytotoxic effects of the simplified right-half model of RT compounds on the NSCLC cell line (H460). (<b>a</b>) The structures of RT and the simplified right-half model of RT compounds: TM-(–)-45, TM-(–)-18, TM-(–)-4a, TM-(–)-52, and TM-(–)-55. (<b>b</b>) The H460 cell line was treated with 0–25 μM of the compounds for 24 h. Then, the MTT assay was used to determine the percentages of cell viability. (<b>c</b>) IC<sub>50</sub> values in each cell line were calculated in comparison to the untreated control. Data represent the mean ± SEM (<span class="html-italic">n</span> = 3) (* 0.01 ≤ <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, compared with the untreated control).</p>
Full article ">Figure 3
<p>Effects of RT on the expression of the apoptotic-related proteins: Mcl-1, Bcl-2, and Bax in the NSCLC cell line (H460) and patient-derived primary cancer cell lines (ELC12, ELC16, ELC17, and ELC20). (<b>a</b>) H460, ELC12, and ELC16 cell lines were seeded and treated with 0–1 μM of RT for 24 h. Then, immunofluorescence analysis was performed using an antibody against Mcl-1. The Alexa Fluor 488 conjugated secondary antibody and Hoechst 33342 were added and the images were visualized under a fluorescence microscope. (<b>b</b>) ELC12, ELC16, ELC17, and ELC20 cell lines were seeded and treated with 0–1 μM of RT for 24 h. Western blot analysis was performed to detect the Mcl-1, Bcl-2, and Bax protein levels. The blots were reprobed with β-actin to confirm an equal loading of each of the protein samples and densitometry was used to calculate the relative protein levels. Densitometric values of protein expression levels were presented as the fold changes relative to β-actin. (<b>c</b>) H460 cells were seeded and treated with 0–1 μM of RT, TM-(–)-45, TM-(–)-18, TM-(–)-4a, TM-(–)-52, and TM-(–)-55 for 24 h. After that, Western blot analysis was performed to detect the Mcl-1, Bcl-2, and Bax protein levels. Data represent the mean ± SEM (<span class="html-italic">n</span> = 3) (* 0.01 ≤ <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, compared with the untreated control).</p>
Full article ">Figure 4
<p>Effects of the simplified right-half model of RT compounds: TM-(–)-18 and TM-(–)-4a on cell viability, apoptotic induction, and apoptotic-related protein expressions in patient-derived primary cancer cell lines (ELC12, ELC16, ELC17, and ELC20). (<b>a</b>) ELC12, ELC16, ELC17, and ELC20 cells were seeded and treated with 0–25 μM of TM-(–)-18 and TM-(–)-4a for 24 h. Then, the MTT assay was performed to determine the percentages of cell viability. (<b>b</b>) H460, ELC12, and ELC16 cells were seeded and treated with 0–10 μM of TM-(–)-18 and TM-(–)-4a for 24 h. Hoechst 33342 and PI were added and then the images were visualized using a fluorescence microscope. (<b>c</b>) H460, ELC12, and ELC16 cells were seeded and treated with 0–1 μM of TM-(–)-18 and TM-(–)-4a for 24 h before performing immunofluorescence using the Mcl-1 primary antibody. Alexa Fluor 488 conjugated secondary antibody and Hoechst 33342 were added and the images were visualized under a fluorescence microscope. (<b>d</b>) ELC12, ELC16, ELC17, and ELC20 cell lines were seeded and treated with 0–1 μM of TM-(–)-18 and TM-(–)-4a for 24 h. Western blot analysis was performed to detect the Mcl-1, Bcl-2, and Bax protein levels. The blots were reprobed with β-actin to confirm an equal loading of each of the protein samples and densitometry was used to calculate the relative protein levels. Densitometric values of protein expression levels were presented as the fold changes relative to β-actin. Data represent the mean ± SEM (<span class="html-italic">n</span> = 3) (* 0.01 ≤ <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, compared with the untreated control).</p>
Full article ">Figure 5
<p>(<b>a</b>) Structures of RT, TM-(–)-18, and TM-(–)-4a. Circled part represents a similar structure between these three compounds. (<b>b</b>) Stereochemical quality of the homology model of Mcl-1 created by the I-TASSER server. Ramachandran plot of Mcl-1 generated by PROCHECK. Areas colored by red, yellow, beige, and white indicate the most favored, additionally allowed, generously allowed, and disallowed regions, respectively.</p>
Full article ">Figure 6
<p>Binding mode and docking energy of: (<b>a</b>) renieramycin T, (<b>b</b>) TM-(−)-18, (<b>c</b>) TM-(−)-4a, and (<b>d</b>) TM-(−)-45 bound to the binding site of Mcl-1 (residues 137–143) taken from the AutoDock Vina molecular docking study. (<b>e</b>) Superimposition structure of each compound at the binding site of Mcl-1.</p>
Full article ">Figure 7
<p>Renieramycin T (RT) and the simplified right-half model of RT compounds: TM-(–)-18 and TM-(–)-4a could enhance the absence of the Mcl-1 protein levels through Mcl-1 proteasomal degradation, resulting in the apoptosis of NSCLC cells. Mcl-1 normally functions as an anti-apoptotic protein by forming a complex with Bak. However, when Mcl-1 disappears, Bak is relieved to form an oligomerization that can permeabilize the outer membrane of mitochondria. Cytochrome c is released to initiate the apoptosis mechanism through caspase activation.</p>
Full article ">Figure 8
<p>Simplified right-half model of RT compounds: TM-(–)-45, TM-(–)-18, TM-(–)-4a, TM-(–)-52, and TM-(–)-55 synthetic pathways. Compound <b>1</b> could be easily obtained from L-Tyr [<a href="#B65-cancers-12-00875" class="html-bibr">65</a>]. The phenol of compound <b>1</b> was selectively protected with 1.2 eq. of NaH and BnBr in DMF to give compound <b>2</b>. The alkylation of the lactam nitrogen of compound <b>2</b> with allyl bromide gave compound <b>3</b> in an 81% yield. The lactam carbonyl of compound <b>2</b> was partially reduced with LiAlH<sub>2</sub>(OEt)<sub>2</sub> in THF to generate the aminal, [<a href="#B66-cancers-12-00875" class="html-bibr">66</a>] which was then treated with KCN and H<sub>2</sub>O to give the α-aminonitrile TM-(–)-52 as a single diastereomer. Chemoselective debenzylation was achieved with BCl<sub>3</sub> in the presence of pentamethylbenzene to give the phenol. [<a href="#B67-cancers-12-00875" class="html-bibr">67</a>] Finally, oxidation of this obtained phenol using salcomine with O<sub>2</sub> afforded TM-(–)-55 in a 74% yield; whereas compound <b>2</b> was bisbenzylated with 2 eq. of benzylbromide in the presence of 10 eq. of NaH to give compound <b>5</b> in a 97% yield. The debenzylation of 19 by using BCl<sub>3</sub> gave the phenol TM-(–)-45. The reductive cyanation of compound <b>5</b> generated the aminonitrile compound <b>6</b>, which was chemoselectively debenzylated with BCl<sub>3</sub> to afford TM-(–)-18 in an 89% yield. Finally, the phenol compound <b>5</b> was oxidized with salcomine in an oxygen atmosphere to give TM-(–)-4a.</p>
Full article ">
15 pages, 2442 KiB  
Article
Renieramycin T Induces Lung Cancer Cell Apoptosis by Targeting Mcl-1 Degradation: A New Insight in the Mechanism of Action
by Korrakod Petsri, Supakarn Chamni, Khanit Suwanborirux, Naoki Saito and Pithi Chanvorachote
Mar. Drugs 2019, 17(5), 301; https://doi.org/10.3390/md17050301 - 21 May 2019
Cited by 19 | Viewed by 4787
Abstract
Among malignancies, lung cancer is the major cause of cancer death. Despite the advance in lung cancer therapy, the five-year survival rate is extremely restricted due to therapeutic failure and disease relapse. Targeted therapies selectively inhibiting certain molecules in cancer cells have been [...] Read more.
Among malignancies, lung cancer is the major cause of cancer death. Despite the advance in lung cancer therapy, the five-year survival rate is extremely restricted due to therapeutic failure and disease relapse. Targeted therapies selectively inhibiting certain molecules in cancer cells have been accepted as promising ways to control cancer. In lung cancer, evidence has suggested that the myeloid cell leukemia 1 (Mcl-1) protein, an anti-apoptotic member of the Bcl-2 family, is a target for drug action. Herein, we report the Mcl-1 targeting activity of renieramycin T (RT), a marine-derived tetrahydroisoquinoline alkaloid that was isolated from the Thai blue sponge Xestospongia sp. RT was shown to be dominantly toxic to lung cancer cells compared to the normal cells in the lung. The cytotoxicity of this compound toward lung cancer cells was mainly exerted through apoptosis induction. For the mechanism of action, we found that RT mediated activation of p53 protein and caspase-9 and -3 activations. While others Bcl-2 family proteins (Bcl-2, Bak, and Bax) were minimally changed in response to RT, Mcl-1 protein was dramatically diminished. We further performed the cycloheximide experiment and found that the half-life of Mcl-1 was significantly shortened by RT treatment. When MG132, a potent selective proteasome inhibitor, was utilized, it could restore the Mcl-1 level. Furthermore, immunoprecipitation analysis revealed that RT significantly increased the formation of Mcl-1-ubiquitin complex compared to the non-treated control. In conclusion, we report the potential apoptosis induction of RT with a mechanism of action involving the targeting of Mcl-1 for ubiquitin-proteasomal degradation. As Mcl-1 is critical for cancer cell survival and chemotherapeutic failure, this novel information regarding the Mcl-1-targeted compound would be beneficial for the development of efficient anti-cancer strategies or targeted therapies. Full article
(This article belongs to the Special Issue Bioactive Compounds from Marine Sponges 2020)
Show Figures

Figure 1

Figure 1
<p>The structures of renieramycin T (RT) and ecteinascidin 743 (Et 743).</p>
Full article ">Figure 2
<p>Renieramycin T (RT) reduced cell viability and induced apoptosis in NSCLC and human normal lung epithelial (BEAS-2B) cell lines. (<b>A</b>) NSCLC and BEAS-2B cell lines were treated with various concentrations of RT (0–25 µM) for 24 h. Percentages of cell viability were determined using the MTT assay. (<b>B</b>) The half maximal inhibitory concentrations (IC<sub>50</sub>) in NSCLC and BEAS-2B cell lines were calculated by comparison with the untreated control. (<b>C</b>–<b>G</b>) H460 and BEAS-2B cell lines were treated with RT (0–25 µM) for 24 h. Hoechst 33342 and propidium iodide (PI) were added. Then, Images were detected by using an inverted fluorescence microscope (a–c). A condensed blue fluorescence of Hoechst 33342 reflected fragmented chromatin in apoptotic cells (c) while a red fluorescence of PI reflected late apoptotic or necrotic cells (b) comparing with no staining condition (a). Percentages of nuclear fragmented and PI positive cells were calculated. (<b>H</b>) H460 was treated with RT (0–25 µM) for 24 h. Apoptotic and necrotic cells were determined using annexin V-FITC/PI staining with flow cytometry. (<b>I</b>) Percentages of cells at each stage were calculated. Data represented the mean ± SEM (n = 3) (* 0.01 ≤ <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, compared with the untreated control).</p>
Full article ">Figure 3
<p>Renieramycin T (RT) induced activation of PARP, caspase-3, and caspase-9 in the H460 cell line. Moreover, RT also significantly up-regulated p53 and down-regulated Mcl-1 in the H460 cell line. H460 cells were treated with RT (0–25 µM) for 24 h. (<b>A</b>) and (<b>C</b>) Apoptotis-related proteins were measured with Western blot analysis. The blots were reprobed with β-actin to confirm equal loading of each of the protein samples. (<b>B</b>) and (<b>D</b>) The relative protein levels were calculated by densitometry. Data represented the mean ± SEM (n = 3) (* 0.01 ≤ <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, compared with the untreated control).</p>
Full article ">Figure 4
<p>Renieramycin T (RT) induced ubiquitin-mediated Mcl-1 proteasomal degradation. (<b>A</b>) Cycloheximide (CHX) chasing assay was performed to measure Mcl-1 half-lives. H460 cells were treated with RT (0–5 µM) with or without 50 µg/mL CHX as indicated by the time in h. Western blot analysis was performed for determined Mcl-1 levels. The blots were reprobed with β-actin to confirm equal loading of each of the protein samples. (<b>B</b>) The relative protein levels were calculated by densitometry (** <span class="html-italic">p</span> &lt; 0.01, compared with the untreated control at 0 h, # 0.01 ≤ <span class="html-italic">p</span> &lt; 0.05, ## <span class="html-italic">p</span> &lt; 0.01, compared with the untreated control at the same time). (<b>C</b>) Mcl-1 half-lives were calculated. (<b>D</b>) H460 cell line was treated with RT (0–5 µM) with or without MG132 (0–20 µM) for 4 h. Mcl-1 expression levels were measured using Western blot analysis. The blots were reprobed with β-actin to confirm equal loading of each of the protein samples. (<b>E</b>) The reversal of RT-mediated down-regulation of Mcl-1 levels by MG132 was calculated by densitometry compared to the non-MG132 treated group (* 0.01 ≤ <span class="html-italic">p</span> &lt; 0.05, compared with the non-MG132 treated group). (<b>F</b>) H460 was treated with RT (5 µM) and MG132 (10 µM) for 4 h. Then, protein lysates were collected subsequent to Mcl-1 immunoprecipitation, and the ubiquitinated protein levels were measured by Western blotting. (<b>G</b>) Ub-Mcl-1 levels were quantified using densitometry (* <span class="html-italic">p</span> &lt; 0.01, compared with the untreated control) All data represented the mean ± SEM (n = 3).</p>
Full article ">Figure 5
<p>Renieramycin T (RT) enhances apoptosis induction through Mcl-1 proteasomal degradation. Normally, Mcl-1 forms complex with Bak to inhibit its apoptotic function, but when Mcl-1 is degraded through the ubiquitin-mediated proteasomal degradation by treatment of RT, Bak is relieved. Activated Bak forms oligomerization that can permeabilize the outer membrane of mitochondria and release cytochrome c to initiate the apoptosis mechanism.</p>
Full article ">
17 pages, 1777 KiB  
Article
Asymmetric Synthesis and Cytotoxicity Evaluation of Right-Half Models of Antitumor Renieramycin Marine Natural Products
by Takehiro Matsubara, Masashi Yokoya, Natchanun Sirimangkalakitti and Naoki Saito
Mar. Drugs 2019, 17(1), 3; https://doi.org/10.3390/md17010003 - 20 Dec 2018
Cited by 10 | Viewed by 3578
Abstract
A general protocol for the asymmetric synthesis of 3-N-arylmethylated right-half model compounds of renieramycins was developed, which enabled structure–activity relationship (SAR) study of several 3-N-arylmethyl derivatives. The most active compound (6a) showed significant cytotoxic activity against human [...] Read more.
A general protocol for the asymmetric synthesis of 3-N-arylmethylated right-half model compounds of renieramycins was developed, which enabled structure–activity relationship (SAR) study of several 3-N-arylmethyl derivatives. The most active compound (6a) showed significant cytotoxic activity against human prostate cancer DU145 and colorectal cancer HCT116 cell lines (IC50 = 11.9, and 12.5 nM, respectively). Full article
(This article belongs to the Special Issue Chemical Modification of Marine Natural Products)
Show Figures

Figure 1

Figure 1
<p>Antitumor tetrahydroisoquinoline marine natural products.</p>
Full article ">Figure 2
<p>Structures of right-half model compounds of renieramycins.</p>
Full article ">Figure 3
<p>Structural comparison of right-half model compounds <b>6a</b>, <b>6b</b>, and <b>6c</b> with renieramycin M (<b>1m</b>).</p>
Full article ">Figure 4
<p>Retrosynthetic analysis of chiral CDE-ring model compound <b>6</b>.</p>
Full article ">Scheme 1
<p>Preparation of compound (±)-<b>6a</b> [<a href="#B9-marinedrugs-17-00003" class="html-bibr">9</a>,<a href="#B11-marinedrugs-17-00003" class="html-bibr">11</a>].</p>
Full article ">Scheme 2
<p>Construction of tricyclic lactam <b>18</b>.</p>
Full article ">Scheme 3
<p>Preparation of right-half model compound <b>6a</b>.</p>
Full article ">Scheme 4
<p>Preparation of right-half model compound <b>6d</b>.</p>
Full article ">Scheme 5
<p>Preparation of right-half model compounds <b>6b</b> and <b>6c</b>.</p>
Full article ">
443 KiB  
Article
Chemistry of Renieramycins. Part 14: Total Synthesis of Renieramycin I and Practical Synthesis of Cribrostatin 4 (Renieramycin H)
by Masashi Yokoya, Keiichiro Kobayashi, Mitsuhiro Sato and Naoki Saito
Mar. Drugs 2015, 13(8), 4915-4933; https://doi.org/10.3390/md13084915 - 6 Aug 2015
Cited by 13 | Viewed by 7298
Abstract
The first total synthesis of (±)-renieramycin I, which was isolated from the Indian bright blue sponge Haliclona cribricutis, is described. The key step is the selenium oxide oxidation of pentacyclic bis-p-quinone derivative (3) stereo- and regioselectively. We also [...] Read more.
The first total synthesis of (±)-renieramycin I, which was isolated from the Indian bright blue sponge Haliclona cribricutis, is described. The key step is the selenium oxide oxidation of pentacyclic bis-p-quinone derivative (3) stereo- and regioselectively. We also report a large-scale synthesis of cribrostatin 4 (renieramycin H) via the C3-C4 double bond formation in an early stage based on the Avendaño’s protocol, from readily available 1-acetyl-3-(3-methyl-2,4,5-trimethylphenyl)methyl-piperazine-2,5-dione (8) in 18 steps (8.3% overall yield). The synthesis provides unambiguous evidence supporting the original structure of renieramycin I. Full article
(This article belongs to the Special Issue Marine Secondary Metabolites)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Structures of bis-1,2,3,4-tetrahydroisoquinoline marine natural products.</p>
Full article ">Figure 2
<p>Structure of 1-<span class="html-italic">epi</span>-pentacyclic alcohol.</p>
Full article ">Figure 3
<p>Epimerization at C-3 through regioselective bromination at C-3 position and reduction sequences by Avendaño and co-workers.</p>
Full article ">Figure 4
<p>Strategy for practical synthesis of compound <b>3</b>, which will be converted into <b>1i</b> and <b>2</b>.</p>
Full article ">Figure 5
<p><b>8</b> → <b>9</b>: (1) TMSCl, TEA, CH<sub>2</sub>Cl<sub>2</sub>; (2) (EtO)<sub>2</sub>CHCH<sub>2</sub>OBz, TMSOTf, Ac<sub>2</sub>O, CH<sub>2</sub>Cl<sub>2</sub>; (3) NBS, CCl<sub>4</sub>, 60 °C, 3 h; (4) H<sub>2</sub>, Pd/C, <span class="html-italic"><sup>i</sup></span>PrOH/DMF, 25 °C, 11 h.</p>
Full article ">Figure 6
<p>Preparation of key intermediate <b>11a</b>.</p>
Full article ">Figure 7
<p>Construction of pentacyclic primary alcohol <b>12</b>.</p>
Full article ">Figure 8
<p>Transformation of compound <b>12</b> into renieramycin I (<b>1i</b>) and cribrostatin 4 (<b>2</b>) through compound <b>23</b>.</p>
Full article ">
Back to TopTop