[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (309)

Search Parameters:
Keywords = xanthones

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
26 pages, 8350 KiB  
Review
Naturally Occurring Xanthones and Their Biological Implications
by Ayodeji O. Oriola and Pallab Kar
Molecules 2024, 29(17), 4241; https://doi.org/10.3390/molecules29174241 - 6 Sep 2024
Viewed by 287
Abstract
Xanthones are chemical substances in higher plants, marine organisms, and lower microorganisms. The most prevalent naturally occurring sources of xanthones are those belonging to the families Caryophyllaceae, Guttiferae, and Gentianaceae. Structurally, xanthones (9H xanthan-9-one) are heterocyclic compounds with oxygen and a γ-pyrone component. [...] Read more.
Xanthones are chemical substances in higher plants, marine organisms, and lower microorganisms. The most prevalent naturally occurring sources of xanthones are those belonging to the families Caryophyllaceae, Guttiferae, and Gentianaceae. Structurally, xanthones (9H xanthan-9-one) are heterocyclic compounds with oxygen and a γ-pyrone component. They are densely packed with a two-benzene ring structure. The carbons in xanthones are numbered from their nucleus and biosynthetic construct. They have mixed shikimate-acetate (higher plants) and acetate-malonate (lower organisms) biosynthetic origins, which influence their classification. Based on the level of oxidation of the C-ring, they are classified into monomers, dimers, and heterodimers. While based on the level of oxygenation or the type of ring residue, they can be categorized into mono-, di-, tri-, tetra-, penta- and hexa-oxygenated xanthones, bis-xanthones, prenylated and related xanthones, xanthonolignoids, and other miscellaneous xanthones. This structural diversity has made xanthones exhibit considerable biological properties as promising antioxidant, antifungal, antimicrobial, and anticancer agents. Structure-activity relationship studies suggest C-1, C-3, C-6, and C-8 as the key positions that influence the biological activity of xanthones. Furthermore, the presence of functional groups, such as prenyl, hydroxyl, glycosyl, furan, and pyran, at the key positions of xanthones, may contribute to their spectrum of biological activity. The unique chemical scaffolds of xanthones, their notable biological activities, and the structure–activity relationships of some lead molecules were discussed to identify lead molecules as possible drug candidates. Full article
(This article belongs to the Special Issue Plant Bioactive Compounds in Pharmaceuticals)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Chemical structure of xanthone, showing its (<b>A</b>) basic nucleus/tricyclic ring system and (<b>B</b>) the different oxidation states of the C-ring. The carbon numbers are indicated in blue color.</p>
Full article ">Figure 2
<p>Reaction scheme to produce xanthone-based compounds [<a href="#B34-molecules-29-04241" class="html-bibr">34</a>].</p>
Full article ">Figure 3
<p>Biosynthetic pathways for xanthones via mixed shikimate-acetate origin (<b>A</b>,<b>B</b>) in higher plants [<a href="#B14-molecules-29-04241" class="html-bibr">14</a>] and acetate origin (<b>C</b>) in lower organisms [<a href="#B3-molecules-29-04241" class="html-bibr">3</a>].</p>
Full article ">Figure 3 Cont.
<p>Biosynthetic pathways for xanthones via mixed shikimate-acetate origin (<b>A</b>,<b>B</b>) in higher plants [<a href="#B14-molecules-29-04241" class="html-bibr">14</a>] and acetate origin (<b>C</b>) in lower organisms [<a href="#B3-molecules-29-04241" class="html-bibr">3</a>].</p>
Full article ">Figure 4
<p>Antifungal activity of some naturally occurring xanthones showing the structure–activity relationships (colored moieties) against (<b>A</b>) <span class="html-italic">Cladosporium cucumerinum</span>; and (<b>B</b>) <span class="html-italic">Candida albicans</span> [<a href="#B61-molecules-29-04241" class="html-bibr">61</a>,<a href="#B62-molecules-29-04241" class="html-bibr">62</a>].</p>
Full article ">Figure 5
<p>Antibacterial activity of some xanthones of <span class="html-italic">Garcinia cowa</span> against <span class="html-italic">Bacillus cereus</span> [<a href="#B64-molecules-29-04241" class="html-bibr">64</a>]. The bioactive moieties are illustrated in red and blue.</p>
Full article ">Figure 6
<p>Some natural anticancer xanthones show the structure–activity relationships in red and blue.</p>
Full article ">Figure 7
<p>Some xanthones from <span class="html-italic">Garcinia mangostana</span> showing their structure–activity relationships based on CDK4 inhibition [<a href="#B77-molecules-29-04241" class="html-bibr">77</a>]. Functional groups in blue and red influence anticancer activity.</p>
Full article ">Figure 8
<p>Xanthones from <span class="html-italic">Polygala japonica</span> with their in vitro ferric-reducing antioxidant power [<a href="#B78-molecules-29-04241" class="html-bibr">78</a>]. Functional groups indicated in blue and red are presented to show the structure–activity relationship.</p>
Full article ">Figure 9
<p>Notable anti-inflammatory xanthones showing their active functional groups in blue and red [<a href="#B80-molecules-29-04241" class="html-bibr">80</a>,<a href="#B82-molecules-29-04241" class="html-bibr">82</a>].</p>
Full article ">Figure 10
<p>Anti-HIV xanthones from <span class="html-italic">Swertia franchetiana</span> [<a href="#B84-molecules-29-04241" class="html-bibr">84</a>] and <span class="html-italic">Maclura tinctoria</span> [<a href="#B85-molecules-29-04241" class="html-bibr">85</a>]. [The functional groups that may be involved in the anti-HIV activity are in multicolor, including the green notation to highlight the catechol group].</p>
Full article ">Figure 11
<p>Notable antidiabetic xanthones from <span class="html-italic">Garcinia mangostana</span>, <span class="html-italic">G. hanburyi</span> and <span class="html-italic">Mangifera indica</span> [<a href="#B87-molecules-29-04241" class="html-bibr">87</a>,<a href="#B88-molecules-29-04241" class="html-bibr">88</a>,<a href="#B91-molecules-29-04241" class="html-bibr">91</a>]. [Functional groups in blue and red may be involved in the bioactivity].</p>
Full article ">Figure 12
<p>Chemical structures of some xanthones showing those with insecticidal activities (<b>45</b>–<b>49</b>) [<a href="#B93-molecules-29-04241" class="html-bibr">93</a>,<a href="#B94-molecules-29-04241" class="html-bibr">94</a>,<a href="#B95-molecules-29-04241" class="html-bibr">95</a>,<a href="#B96-molecules-29-04241" class="html-bibr">96</a>,<a href="#B97-molecules-29-04241" class="html-bibr">97</a>] and a proposed structure for possible biological evaluation (<b>50</b>). [The functional groups that may be involved in the bioactivity are in multicolor, including the green notation to highlight the catechol group].</p>
Full article ">
19 pages, 2977 KiB  
Article
Evaluation of the Antidiabetic Potential of Xanthone-Rich Extracts from Gentiana dinarica and Gentiana utriculosa
by Jelena Arambašić Jovanović, Dijana Krstić-Milošević, Branka Vinterhalter, Svetlana Dinić, Nevena Grdović, Aleksandra Uskoković, Jovana Rajić, Marija Đorđević, Ana Sarić, Melita Vidaković and Mirjana Mihailović
Int. J. Mol. Sci. 2024, 25(16), 9066; https://doi.org/10.3390/ijms25169066 - 21 Aug 2024
Viewed by 372
Abstract
Despite the existence of various therapeutic approaches, diabetes mellitus and its complications have been an increasing burden of mortality and disability globally. Hence, it is necessary to evaluate the efficacy and safety of medicinal plants to support existing drugs in treating diabetes. Xanthones, [...] Read more.
Despite the existence of various therapeutic approaches, diabetes mellitus and its complications have been an increasing burden of mortality and disability globally. Hence, it is necessary to evaluate the efficacy and safety of medicinal plants to support existing drugs in treating diabetes. Xanthones, the main secondary metabolites found in Gentiana dinarica and Gentiana utriculosa, display various biological activities. In in vitro cultured and particularly in genetically transformed G. dinarica and G. utriculosa roots, there is a higher content of xanthones. The aim of this study was to investigate and compare antidiabetic properties of secondary metabolites (extracts) prepared from these two Gentiana species, cultured in vitro and genetically transformed with those collected from nature. We compare HPLC secondary metabolite profiles and the content of the main extract compounds of G. dinarica and G. utriculosa methanol extracts with their ability to scavenge DPPH free radicals and inhibit intestinal α-glucosidase in vitro. Anti-hyperglycemic activity of selected extracts was tested further in vivo on glucose-loaded Wistar rats. Our findings reveal that the most prominent radical scavenging potential and potential to control the rise in glucose level, detected in xanthone-rich extracts, were in direct correlation with an accumulation of xanthones norswertianin and norswertianin-1-O-primeveroside in G. dinarica and decussatin and decussatin-1-O-primeveroside in G. utriculosa. Full article
Show Figures

Figure 1

Figure 1
<p>Comparative chromatograms of <span class="html-italic">G. dinarica</span> (Gd) (<b>a</b>,<b>b</b>) and <span class="html-italic">G. utriculosa</span> (Gu) (<b>c</b>) methanol extracts. (<b>a</b>) Gd1—aerial parts of wild growing plants; Gd2—roots of wild growing plants; Gd3—shoot culture. Peaks: 1—swertiamarin; 2—isoorinetin-4′-O-glucoside; 3—gentiopicrin; 4—sweroside; 5—isoorientin; 6—norswertianin-1-O-primeveroside; 7—norswertianin-8-O-primeveroside; 8—gentioside; 9—amarogentin; 10—norswertianin; (<b>b</b>) Gd4—vegetative roots; Gd5—genetically transformed roots, clone B; Gd6—genetically transformed roots, clone 3. Peaks: 6—norswertianin-1-O-primeveroside; 7—norswertianin-8-O-primeveroside; 8—gentioside; 10—norswertianin; (<b>c</b>) Gu1—aerial parts of wild growing plants; Gu2—shoot culture; Gu3—genetically transformed shoots; Gu4—genetically transformed roots. Peaks: 11—mangiferin; 12—gentiakochianin-1-O-primeveroside; 13—decussatin-1-O-primeveroside; 14—1,8-dihydroxy-3-methoxy-7-O-primeveroside; 15—gentiakochianin; 16—decussatin.</p>
Full article ">Figure 2
<p>Chemical structures of compounds identified in methanol extracts of <span class="html-italic">G. dinarica</span> (Gd) and <span class="html-italic">G. utriculosa</span> (Gu). The order of compounds is in accordance with their increasing retention times in the chromatograms. (<b>a</b>) Bitter glycosides identified in <span class="html-italic">G. dinarica</span> methanol extracts: 1—swertiamarin; 3—gentiopicrin; 4—sweroside; 9—amarogentin; (<b>b</b>) C-glucoflavones and xanthone-C-glucoside identified in methanol extracts of <span class="html-italic">G. dinarica</span> and <span class="html-italic">G. utriculosa</span>: 2—isoorinetin-4′-O-glucoside; 5—isoorientin; 11—mangiferin; (<b>c</b>) xanthones identified in methanol extracts of <span class="html-italic">G. dinarica</span> and <span class="html-italic">G. utriculosa</span>: 6—norswertianin-1-O-primeveroside; 7—norswertianin-8-O-primeveroside; 8—gentioside; 10—norswertianin; 12—gentiakochianin-1-O-primeveroside; 13—decussatin-1-O-primeveroside; 14—1,8-dihydroxy-3-methoxy-7-O-primeveroside; 15—gentiakochianin; 16—decussatin.</p>
Full article ">Figure 3
<p>Comparative analysis of DPPH radical scavenging activity at different concentrations of methanol extracts isolated from <span class="html-italic">G. dinarica</span> and <span class="html-italic">G. utriculosa</span>. Gd1—aerial parts of wild growing plants; Gd2—roots of wild growing plants; Gd3—shoot culture; Gd4—vegetative roots; Gd5—genetically transformed roots, clone B; Gd6—genetically transformed roots, clone 3. Gu1—aerial parts of wild growing plants; Gu2—shoot culture; Gu3—genetically transformed shoots; Gu4—genetically transformed roots. The results of the assays are presented as the means ± S.E.M. from three separate measurements (n = 3). For all variables with the same superscript letter, the difference between the means (at the same concentration) is not statistically significant. If two variables have different letters, they are significantly different at <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 4
<p>Chromatograms of <span class="html-italic">G. dinarica</span> extracts derived from transgenic roots clone B (Gd5) before ((<b>top</b>) chromatogram) and after reaction with DPPH radicals ((<b>bottom</b>) chromatogram). Peaks: 6—norswertianin-1-O-primeveroside; 7—norswertianin-8-O-primeveroside; 8—gentioside; 10—norswertianin. * The peaks of compounds that were involved in free radical scavenging activity.</p>
Full article ">Figure 5
<p>Chromatograms of <span class="html-italic">G. utriculosa</span> extracts derived from transgenic shoots (Gu3) before ((<b>top</b>) chromatogram) and after reaction with DPPH radicals ((<b>bottom</b>) chromatogram). Peaks: 11—mangiferin; 12—gentiakochianin-1-O-primeveroside; 13—decussatin-1-O-primeveroside; 14—1,8-dihydroxy-3-methoxy-7-O-primeveroside; 15—gentiakochianin; 16—decussatin. * The peaks of compounds that were involved in free radical scavenging activity.</p>
Full article ">Figure 6
<p>α-Glucosidase inhibitory activity at two different concentrations of methanol extracts obtained from <span class="html-italic">G. dinarica</span> and <span class="html-italic">G. utriculosa</span>. Acarbose was used as standard. Gd1—aerial parts of wild growing plants; Gd2—roots of wild growing plants; Gd3—shoot culture; Gd4—vegetative roots; Gd5—genetically transformed roots, clone B; Gd6—genetically transformed roots, clone 3. Gu1—aerial parts of wild growing plants; Gu2—shoot culture; Gu3—genetically transformed shoots; Gu4—genetically transformed roots. The results of the assays are presented as the means ± S.E.M. from three separate measurements (n = 3). For all variables with the same superscript letter, the difference between the means (at the same concentration) is not statistically significant. If two variables have different letters, they are significantly different at <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 7
<p>In vivo antihyperglycemic effect of selected methanol extracts of <span class="html-italic">G. dinarica</span> and <span class="html-italic">G. utriculosa</span> in oral glucose-loaded rats. Glibenclamide (Glc) was used as a control drug. Gd4—vegetative roots; Gd5—genetically transformed roots, clone B; Gu2—shoot culture; Gu3—genetically transformed shoots. Data are expressed as mean ± S.E.M.; no. of animals (N) = 5. For all variables with the same superscript letter, the difference between the means (same post-glucose overload time) is not statistically significant. If two variables have different letters, they are significantly different at <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 8
<p>Area under the curve (AUC) for the first 60 min of oral glucose overload responses of Wistar rats treated with vehicle or selected methanol extracts of <span class="html-italic">G. dinarica</span> and <span class="html-italic">G. utriculosa</span> compared with standard drug. Glibenclamide (Glc)—standard drug. Gd4—vegetative roots; Gd5—genetically transformed roots, clone B; Gu2—shoot culture; Gu3—genetically transformed shoots. Data are expressed as mean ± S.E.M.; no. of animals (N) = 5. For all variables with the same superscript letter, the difference between the means (same post-glucose overload time) is not statistically significant. If two variables have different letters, they are significantly different at <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">
24 pages, 7360 KiB  
Article
Phytochemical Analysis, Biological Activities, and Docking of Phenolics from Shoot Cultures of Hypericum perforatum L. Transformed by Agrobacterium rhizogenes
by Oliver Tusevski, Marija Todorovska, Jasmina Petreska Stanoeva and Sonja Gadzovska Simic
Molecules 2024, 29(16), 3893; https://doi.org/10.3390/molecules29163893 - 17 Aug 2024
Viewed by 710
Abstract
Hypericum perforatum transformed shoot lines (TSL) regenerated from corresponding hairy roots and non-transformed shoots (NTS) were comparatively evaluated for their phenolic compound contents and in vitro inhibitory capacity against target enzymes (monoamine oxidase-A, cholinesterases, tyrosinase, α-amylase, α-glucosidase, lipase, and cholesterol esterase). Molecular docking [...] Read more.
Hypericum perforatum transformed shoot lines (TSL) regenerated from corresponding hairy roots and non-transformed shoots (NTS) were comparatively evaluated for their phenolic compound contents and in vitro inhibitory capacity against target enzymes (monoamine oxidase-A, cholinesterases, tyrosinase, α-amylase, α-glucosidase, lipase, and cholesterol esterase). Molecular docking was conducted to assess the contribution of dominant phenolic compounds to the enzyme-inhibitory properties of TSL samples. The TSL extracts represent a rich source of chlorogenic acid, epicatechin and procyanidins, quercetin aglycone and glycosides, anthocyanins, naphthodianthrones, acyl-phloroglucinols, and xanthones. Concerning in vitro bioactivity assays, TSL displayed significantly higher acetylcholinesterase, tyrosinase, α-amylase, pancreatic lipase, and cholesterol esterase inhibitory properties compared to NTS, implying their neuroprotective, antidiabetic, and antiobesity potential. The docking data revealed that pseudohypericin, hyperforin, cadensin G, epicatechin, and chlorogenic acid are superior inhibitors of selected enzymes, exhibiting the lowest binding energy of ligand–receptor complexes. Present data indicate that H. perforatum transformed shoots might be recognized as an excellent biotechnological system for producing phenolic compounds with multiple health benefits. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Inhibitory activity (%) of <span class="html-italic">Hypericum perforatum</span> transformed shoot extracts against (<b>a</b>) momoamine oxidase-A (MAO-A), (<b>b</b>) acetylcholinesterase (AChE), (<b>c</b>) butyrylcholinesterase (BChE), (<b>d</b>) tyrosinase (TYR), (<b>e</b>) α-amylase (α-AM), (<b>f</b>) α-glucosidase (α-GL), (<b>g</b>) pancreatic lipase (PL) and (<b>h</b>) cholesterol esterase (CHE). NTS: non-transformed shoots, TSL B, TSL F and TSL H: transformed shoot lines, DCP: 2,4-dichlorophenol.</p>
Full article ">Figure 2
<p>The best-ranked docking pose (<b>a</b>) and key interactions (<b>b</b>) of epicatechin in the active site of monoamine oxidase-A.</p>
Full article ">Figure 3
<p>The best-ranked docking pose (<b>a</b>) and key interactions (<b>b</b>) of pseudohypericin in the active site of acetylcholinesterase. The best-ranked docking pose (<b>c</b>) and key interactions (<b>d</b>) of pseudohypericin in the active site of butyrylcholinesterase.</p>
Full article ">Figure 4
<p>The best-ranked docking pose (<b>a</b>) and key interactions (<b>b</b>) of chlorogenic acid in the active site of tyrosinase.</p>
Full article ">Figure 5
<p>The best-ranked docking pose (<b>a</b>) and key interactions (<b>b</b>) of pseudohypericin in the active site of α-amylase. The best-ranked docking pose (<b>c</b>) and key interactions (<b>d</b>) of pseudohypericin in the active site of α-glucosidase.</p>
Full article ">Figure 6
<p>The best-ranked docking pose (<b>a</b>) and key interactions (<b>b</b>) of pseudohypericin in the active site of lipase. The best-ranked docking pose (<b>c</b>) and key interactions (<b>d</b>) of pseudohypericin in the active site of cholesterol esterase.</p>
Full article ">
21 pages, 20166 KiB  
Article
Hyperthermia Intensifies α-Mangostin and Synthetic Xanthones’ Antimalignancy Properties
by Jakub Rech, Dorota Żelaszczyk, Henryk Marona, Agnieszka Gunia-Krzyżak, Paweł Żmudzki and Ilona Anna Bednarek
Int. J. Mol. Sci. 2024, 25(16), 8874; https://doi.org/10.3390/ijms25168874 - 15 Aug 2024
Viewed by 421
Abstract
In order to improve naturally occurring xanthones’ anticancer properties, chemical synthesis is proposed. In this study, from eight novel xanthone derivatives coupled to morpholine or aminoalkyl morpholine, only the two most active ones were chosen. For additional enhancement of the anticancer activity of [...] Read more.
In order to improve naturally occurring xanthones’ anticancer properties, chemical synthesis is proposed. In this study, from eight novel xanthone derivatives coupled to morpholine or aminoalkyl morpholine, only the two most active ones were chosen. For additional enhancement of the anticancer activity of our tested compounds, we combined chemotherapy with hyperthermia in the range of 39–41 °C, from which the mild conditions of 39 °C were the most influencing. This approach had a profound impact on the anticancer properties of the tested compounds. TOV-21G and SC-OV-3 ovarian cell line motility and metastasis behavior were tested in native and hyperthermia conditions, indicating decreased wound healing properties and clonogenic activity. Similarly, the expression of genes involved in metastasis was hampered. The expression of heat shock proteins involved in cancer progression (Hsc70, HSP90A, and HSP90B) was significantly influenced by xanthone derivatives. Chemotherapy in mild hyperthermia conditions had also an impact on decreasing mitochondria potential, visualized with JC-1. Synthetic xanthone ring modifications may increase the anticancer activity of the obtained substances. Additional improvement of their activity can be achieved by applying mild hyperthermia conditions. Further development of a combined anticancer therapy approach may result in increasing currently known chemotherapeutics, resulting in a greater recovery rate and diminishment of the cytotoxicity of drugs. Full article
(This article belongs to the Special Issue Natural Compounds in Cancer Therapy and Prevention, 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Structure of previously synthesized 6-chloro-2-(((3-morpholino propyl)amino)methyl)-9<span class="html-italic">H</span>-xanthen-9-one dihydrochloride [<a href="#B14-ijms-25-08874" class="html-bibr">14</a>].</p>
Full article ">Figure 2
<p>Proposed novel xanthone derivatives modifications.</p>
Full article ">Figure 3
<p>Effect of different hyperthermia temperatures on growth of SK-OV-3 and TOV-21G cell lines compared to growth in 37 °C set as 100%. ** <span class="html-italic">p</span> &lt; 0.01 vs. untreated 37 °C negative control.</p>
Full article ">Figure 4
<p>Hyperthermia and xanthone derivative treatment influence on the number of colonies formed of (<b>a</b>) TOV-21G and (<b>b</b>) SK-OV-3 cell line, (<b>c</b>) representative images of clonogenic assay. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01 vs. untreated 37 °C negative control.</p>
Full article ">Figure 5
<p>Hyperthermia and xanthone derivative treatment influences apoptosis. Calculated green/orange fluorescence ratio of (<b>a</b>) TOV-21G and (<b>b</b>) SK-OV-3 cell line. Representative images (<b>c</b>) of negative control and after 24 h of compound <b>C8</b> treatment of both cell lines. B—blebbing; CC—chromatin condensation; EA—early apoptosis; LA—late apoptosis. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01 vs. untreated 37 °C negative control.</p>
Full article ">Figure 6
<p>Wound area free from (<b>a</b>) TOV-21G and (<b>b</b>) SK-OV-3 cell growth after 24 and 48 h. Initial wound size was taken as a 100% wound size reference; (<b>c</b>) representative original images of the wound and overlayed wound size prediction by the program (purple/yellow) after <b>C7</b> treatment. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01 vs. initial wound size at T<sub>0</sub>.</p>
Full article ">Figure 7
<p>Average NAO staining green fluorescence corresponding to total mitochondrial mass of (<b>a</b>) TOV-21G and (<b>b</b>) SK-OV-3 cells. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01 vs. untreated 37 °C negative control.</p>
Full article ">Figure 8
<p>The mitochondrial membrane potential of (<b>a</b>) TOV-21G and (<b>b</b>) SK-OV-3, measured by JC-1 staining. Results are presented as orange-to-green fluorescence ratio (O/G ratio). Exemplary images (<b>c</b>) of JC-1 fluorescence after 24 h compound <b>C7</b> treatment. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01 vs. untreated 37 °C negative control.</p>
Full article ">Figure 9
<p>Fold change expression of HSP90A, HSP90B, and Hsc70 genes in (<b>a</b>) TOV-21G and (<b>b</b>) SK-OV-3, under native and hyperthermia conditions. <b>CIS</b> and <b>MAG</b> are used as reference compounds. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01 vs. untreated 37 °C negative control.</p>
Full article ">Scheme 1
<p>Synthesis of the new compounds (<b>a</b>) <b>C1</b>–<b>C4</b>, (<b>b</b>) <b>C5</b>–<b>C6</b>, (<b>c</b>) <b>C7</b>-<b>C8</b>. Reagents and conditions: (i) K<sub>2</sub>CO<sub>3</sub>, water, toluene, rt, 3 h; (ii) K<sub>2</sub>CO<sub>3</sub>, isopropanol, reflux, 3–6 h; (iii) ethanol saturated with HCl, rt, 30 min; and (iv) K<sub>2</sub>CO<sub>3</sub>, toluene, reflux, 4–8 h. The asterisk denote the chiral center.</p>
Full article ">
16 pages, 13502 KiB  
Article
Identification of Penexanthone A as a Novel Chemosensitizer to Induce Ferroptosis by Targeting Nrf2 in Human Colorectal Cancer Cells
by Genshi Zhao, Yanying Liu, Xia Wei, Chunxia Yang, Junfei Lu, Shihuan Yan, Xiaolin Ma, Xue Cheng, Zhengliang You, Yue Ding, Hongwei Guo, Zhiheng Su, Shangping Xing and Dan Zhu
Mar. Drugs 2024, 22(8), 357; https://doi.org/10.3390/md22080357 - 6 Aug 2024
Viewed by 882
Abstract
Ferroptosis has emerged as a potential mechanism for enhancing the efficacy of chemotherapy in cancer treatment. By suppressing nuclear factor erythroid 2-related factor 2 (Nrf2), cancer cells may lose their ability to counteract the oxidative stress induced by chemotherapy, thereby becoming more susceptible [...] Read more.
Ferroptosis has emerged as a potential mechanism for enhancing the efficacy of chemotherapy in cancer treatment. By suppressing nuclear factor erythroid 2-related factor 2 (Nrf2), cancer cells may lose their ability to counteract the oxidative stress induced by chemotherapy, thereby becoming more susceptible to ferroptosis. In this study, we investigate the potential of penexanthone A (PXA), a xanthone dimer component derived from the endophytic fungus Diaporthe goulteri, obtained from mangrove plant Acanthus ilicifolius, to enhance the therapeutic effect of cisplatin (CDDP) on colorectal cancer (CRC) by inhibiting Nrf2. The present study reported that PXA significantly improved the ability of CDDP to inhibit the activity of and induce apoptosis in CRC cells. Moreover, PXA was found to increase the level of oxidative stress and DNA damage caused by CDDP. In addition, the overexpression of Nrf2 reversed the DNA damage and ferroptosis induced by the combination of PXA and CDDP. In vivo experiments using zebrafish xenograft models demonstrated that PXA enhanced the therapeutic effect of CDDP on CRC. These studies suggest that PXA enhanced the sensitivity of CRC to CDDP and induce ferroptosis by targeting Nrf2 inhibition, indicating that PXA might serve as a novel anticancer drug in combination chemotherapy. Full article
(This article belongs to the Special Issue Pharmacological Potential of Marine Natural Products, 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>PXA sensitizes CRC cells to CDDP-induced cytotoxicity and apoptosis. (<b>A</b>) Chemical structural of PXA. (<b>B</b>) CRC cells were co-treated with PXA and CDDP for 48 h, and the percentage of cell viability was determined by CCK-8 assay. (<b>C</b>) 3D visualization of synergy scores between PXA and CDDP obtained using the SynergyFinder tool; these calculated average synergy scores are 20.7 and 11.3 for these two panels of drug combinations (Synergy scores &gt; 10 are considered synergistic). (<b>D</b>) The percentage of apoptotic cells was analyzed and quantified using flow cytometry after Annexin V-FITC/PI staining. (<b>E</b>) The protein levels of Cleaved-PARP, Cleaved-caspase-3, BAX, and Bcl-2 in HCT116 and HT29 cells were detected by Western blot after 24 h treatment; β-acting was used as a loading control. Results are expressed as means ± SD. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 versus the control group, <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05, <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 versus the CDDP-treatment group.</p>
Full article ">Figure 2
<p>PXA increased CDDP-induced ROS production. (<b>A</b>,<b>B</b>) HCT116 and HT29 cells treated with PXA and CDDP were incubated with the DCFH-DA probe for 20 min, and the ROS levels (DCF fluorescence) were observed and analyzed by fluorescence microscopy (<b>A</b>) and flow cytometry (<b>B</b>). Scale bars: 100 μm. Results are expressed as means ± SD. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 versus the control group, <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 versus the CDDP-treatment group.</p>
Full article ">Figure 3
<p>PXA increased CDDP-induced DNA damage and oxidative stress. (<b>A</b>) DNA damage levels in HCT116 and HT29 cells treated with PXA and CDDP for 24 h were assessed using the comet assay. Scale bars: 100 μm. (<b>B</b>) After 24 h of treating HCT116 and HT29 cells with PXA and CDDP, the content of GSH, SOD, and HO-1 was measured using ELISA. Results are expressed as means ± SD. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 versus the control group, <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05, <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 versus the CDDP-treatment group.</p>
Full article ">Figure 4
<p>PXA inhibits Nrf2 protein expression. (<b>A</b>,<b>B</b>) Western blotting analyses of Nrf2 expression in HCT116 and HT29 cells treated with PXA for indicated concentrations (<b>A</b>) and time points (<b>B</b>). (<b>C</b>) CETSA was performed to confirm that PXA targets Nrf2 proteins. Results are expressed as means ± SD. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 versus the control group.</p>
Full article ">Figure 5
<p>Nrf2 overexpression reverses the chemosensitizing activity of PXA. (<b>A</b>) Protein levels of Nrf2 in stably overexpressing empty vector (EV) and Nrf2 HCT116 and HT29 cells were examined by Western blotting. (<b>B</b>) Western blot assay to detect the effect of PXA in combination with CDDP on Nrf2 protein in EV and Nrf2 stable overexpressing HCT116 and HT29 cells. (<b>C</b>) CCK-8 assay was performed on EV and Nrf2 stable overexpressing HCT116 and HT29 cells treated with PXA and CDDP for 48 h. (<b>D</b>) The DNA damage levels of EV and Nrf2 stable overexpressing HCT116 and HT29 cells treated with PXA and CDDP for 24 h were evaluated using the comet assay. Scale bars: 100 μm. Results are expressed as means ± SD. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 versus the EV group.</p>
Full article ">Figure 6
<p>PXA enhances CDDP-induced ferroptosis by inhibiting Nrf2 pathway. (<b>A</b>) HCT116 and HT29 cells were treated with PXA and CDDP with or without ferroptosis inhibitor (Fer-1) for 24h, and cell viability was measured using the CCK-8 assay. (<b>B</b>) Detection of lipid hydroperoxides by fluorescence imaging of Liperfluo in HCT116 and HT29 cells treated with PXA and CDDP with or without Fer-1. Scale bars: 100 μm. (<b>C</b>) EV and Nrf2 stable overexpressing HCT116 and HT29 cells were treated with PXA and CDDP for 24 h and the expression of SLC7A11 and GPX4 were examined by Western blotting. Results are expressed as means ± SD. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 versus the control group, <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 versus the Fer-1 group, <sup>%</sup> <span class="html-italic">p</span> &lt; 0.05, <sup>%%</sup> <span class="html-italic">p</span> &lt; 0.01 versus the EV group.</p>
Full article ">Figure 7
<p>PXA enhances the therapeutic efficacy of CDDP in CRC xenograft zebrafish model. (<b>A</b>) HCT116-CM-Dil cells were injected into zebrafish embryo. At the end of the experiments, phenotypic map of fluorescence of HCT116-CM-Dil cells in zebrafish were photographed. Scale bars: 250 μm (<b>B</b>) The fluorescence area and intensity of HCT116-CM-Dil cells in zebrafish were analyzed by Image J software (Version 1.54j). Results are expressed as means ± SD. ** <span class="html-italic">p</span> &lt; 0.01 versus the control group, <span class="html-italic"><sup>##</sup> p</span> &lt; 0.01 versus the CDDP-treatment group.</p>
Full article ">Figure 8
<p>Schematic representation of PXA enhancing the sensitivity of CRC cells to CDDP by inducing ferroptosis through the inhibition of Nrf2. “↓” indicates promotion; “⊥” indicates inhibition.</p>
Full article ">
13 pages, 3585 KiB  
Article
Anti-Inflammatory Effect of Xanthones from Hypericum beanii on Macrophage RAW 264.7 Cells through Reduced NO Production and TNF-α, IL-1β, IL-6, and COX-2 Expression
by Wei Ma, Fu-Cai Ren, Xue-Ru Wang and Ning Li
Molecules 2024, 29(15), 3705; https://doi.org/10.3390/molecules29153705 - 5 Aug 2024
Viewed by 673
Abstract
Hypericum beanii N. Robson, a perennial upright herb, predominantly inhabits temperate regions. This species has been utilized for the treatment of various inflammation-related diseases. One new xanthone 3,7-dihydroxy-1,6-dimethoxyxanthone (1) and twenty-three known xanthones (224) were isolated from [...] Read more.
Hypericum beanii N. Robson, a perennial upright herb, predominantly inhabits temperate regions. This species has been utilized for the treatment of various inflammation-related diseases. One new xanthone 3,7-dihydroxy-1,6-dimethoxyxanthone (1) and twenty-three known xanthones (224) were isolated from the aerial parts of H. beanii. The structure of the new compound was determined based on high-resolution electrospray ionization mass spectroscopy (HR-ESIMS), nuclear magnetic resonance (NMR), Infrared Spectroscopy (IR), ultraviolet spectrophotometry (UV) spectroscopic data. The anti-inflammatory effects of all the isolates were assessed by measuring the inhibitory effect on nitric oxide (NO) production in LPS-stimulated RAW 264.7 macrophages. Compounds 3,4-dihydroxy-2-methoxyxanthone (15), 1,3,5,6-tetrahydroxyxanthone (19), and 1,3,6,7-tetrahydroxyxanthone (22) exhibited significant anti-inflammatory effects at a concentration of 10 μM with higher potency compared to the positive control quercetin. Furthermore, compounds 15, 19, and 22 reduced inducible NO synthase (iNOS), tumor necrosis factor alpha (TNF-α), interleukin-1β (IL-1β), IL-6, and cyclooxygenase 2 (COX-2) mRNA expression in the LPS-stimulated RAW 264.7 macrophages, suggesting that these compounds may mitigate the synthesis of the aforementioned molecules at the transcriptional level, provisionally confirming their anti-inflammatory efficacy. Full article
Show Figures

Figure 1

Figure 1
<p>Structures of compounds <b>1</b>–<b>24</b>.</p>
Full article ">Figure 2
<p>HMBC (<span class="html-fig-inline" id="molecules-29-03705-i001"><img alt="Molecules 29 03705 i001" src="/molecules/molecules-29-03705/article_deploy/html/images/molecules-29-03705-i001.png"/></span>) of compound <b>1</b>.</p>
Full article ">Figure 3
<p>The cytotoxic effects of compounds (<b>1</b>–<b>24</b>) on the RAW 264.7 macrophage cells. Values are the mean ± SEM, n = 3. (Compared to the control group **<span class="html-italic">* p</span> &lt; 0.001, ** <span class="html-italic">p</span> &lt; 0.01, or * <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>The inhibitory effects of <b>1</b>–<b>24</b> against LPS-induced NO production in RAW264.7 macrophages. Mean ± SEM of three replicates is shown. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 with the LPS group. <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 with the CON (control) group.</p>
Full article ">Figure 5
<p>Concentrations of NO and mRNA expression of iNOS in RAW 264.7 cells treated with compounds <b>15</b>, <b>19</b>, and <b>22</b>. Values are the mean ± SEM, n = 3. (Compared to the control group <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001. Compared to the LPS group, *** <span class="html-italic">p</span> &lt; 0.001, ** <span class="html-italic">p</span> &lt; 0.01, or * <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 6
<p>Concentrations of TNF-<span class="html-italic">α</span>, IL-1<span class="html-italic">β</span>, and IL-6 in RAW 264.7 cells with compounds <b>15</b>, <b>19</b>, and <b>22</b> treatments. Values are the mean ± SEM, n = 3. (Compared to the control group <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001. Compared to the LPS group, *** <span class="html-italic">p</span> &lt; 0.001, ** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 7
<p>mRNA expression of pro-inflammatory cytokines TNF-<span class="html-italic">α</span>, IL-1<span class="html-italic">β</span>, and IL-6 in RAW 264.7 cells treated with compounds <b>15</b>, <b>19</b>, and <b>22</b>. Values are the mean ± SEM, n = 3. (Compared to the control group <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001. Compared to the LPS group, *** <span class="html-italic">p</span> &lt; 0.001, ** <span class="html-italic">p</span> &lt; 0.01, or * <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 8
<p>COX-2 mRNA expression in RAW 264.7 cells treated with compounds <b>15</b>, <b>19</b>, and <b>22</b>. Values are the mean ± SEM, n = 3. (Compared to the control group <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001. Compared to the LPS group, *** <span class="html-italic">p</span> &lt; 0.001 or ** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">
5 pages, 761 KiB  
Communication
Caudiquinol: A Meroterpenoid with an Intact C20 Geranylgeranyl Chain Isolated from Garcinia caudiculata
by Maya Valmiki, Stephen Ping Teo, Pedro Ernesto de Resende, Simon Gibbons and A. Ganesan
Molecules 2024, 29(15), 3613; https://doi.org/10.3390/molecules29153613 - 31 Jul 2024
Viewed by 666
Abstract
The tropical Garcinia genus of flowering plants is a prolific producer of aromatic natural products including polyphenols, flavonoids, and xanthones. In this study, we report the first phytochemical investigation of Garcinia caudiculata Ridl. from the island of Borneo. Fractionation, purification, and structure elucidation [...] Read more.
The tropical Garcinia genus of flowering plants is a prolific producer of aromatic natural products including polyphenols, flavonoids, and xanthones. In this study, we report the first phytochemical investigation of Garcinia caudiculata Ridl. from the island of Borneo. Fractionation, purification, and structure elucidation by MS and NMR resulted in the discovery of two meroterpenoids. One was a benzofuranone lactone previously isolated from Iryanthera grandis and Rhus chinensis, and the second was a new hydroquinone methyl ester that we named caudiquinol. Both natural products are rare examples of plant meroterpenoids with an intact geranylgeranyl chain. Full article
Show Figures

Figure 1

Figure 1
<p>Benzofuranone (<b>1</b>) and caudiquinol (<b>2</b>), geranylgeranyl meroterpenoids isolated from <span class="html-italic">Garcinia caudiculata</span>.</p>
Full article ">Figure 2
<p>Observed COSY (blue; bold) and HMBC (red arrows) correlations in caudiquinol.</p>
Full article ">Figure 3
<p>The geranyl meroterpenoids denudalide (<b>3</b>) and denudaquinol (<b>4</b>).</p>
Full article ">Figure 4
<p>Plant meroterpenoids, other than <b>1</b> and <b>2</b>, with an intact geranylgeranyl sidechain.</p>
Full article ">
12 pages, 12253 KiB  
Article
Photocatalytic N-Formylation of CO2 with Amines Catalyzed by Diethyltriamine Pentaacetic Acid
by Xuexin Yuan, Qiqi Zhou, Yu Chen, Hai-Jian Yang, Qingqing Jiang, Juncheng Hu and Cun-Yue Guo
C 2024, 10(3), 62; https://doi.org/10.3390/c10030062 - 11 Jul 2024
Viewed by 776
Abstract
In the present work, inexpensive and commercially available diethyltriamine pentaacetic acid (DTPA) was used as an initiator to catalyze the N-formylation reaction of CO2 with amines via the construction of C-N bonds in the presence of xanthone as the photosensitizer and PhSiH [...] Read more.
In the present work, inexpensive and commercially available diethyltriamine pentaacetic acid (DTPA) was used as an initiator to catalyze the N-formylation reaction of CO2 with amines via the construction of C-N bonds in the presence of xanthone as the photosensitizer and PhSiH3 as the reducing agent. After a systematic study of various factors, the optimal conditions for the photocatalytic reaction were obtained: 2.5 mmol of amine, 2.5 mmol of PhSiH3, 10 mol% of DTPA, 20 mol% of xanthone, 1 mL of dimethylsulfoxide (DMSO), atmospheric pressure, and 35 W UV lamp irradiation for 48 h. Under the optimal conditions, the catalyst system afforded high performance for the N-formylation of amines (primary and secondary amines) and CO2, and the yields of the N-formylated products of dialkylamines were above 70%. Further studies exhibit that the catalytic system has a wide scope of substrate applications. For various alicyclic secondary amines, heterocyclic secondary amines, aliphatic primary amines, and aromatic primary amines, the corresponding N-formylation products can be obtained efficiently. In addition, the catalyst can be recycled by simple precipitation and filtration. After five cycles of recycling, there was no significant change in the catalytic and structural properties of DTPA. Finally, a possible reaction mechanism is proposed. Full article
(This article belongs to the Section CO2 Utilization and Conversion)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Photocatalyzed CO<sub>2</sub> N-formylation reaction.</p>
Full article ">Figure 2
<p>Effect of solvent on the N-formylation of N-methylaniline with CO<sub>2</sub>. Reaction conditions: N-methylaniline (2.5 mmol), xanthone (20 mol%), DTPA (10 mol%), solvent (1 mL), CO<sub>2</sub> (0.1 MPa), PhSiH<sub>3</sub> (2.5 mmol), rt, and 48 h. The yield was determined by <sup>1</sup>H NMR spectra analysis using 1,3,5-trimethoxybenzene as an internal standard.</p>
Full article ">Figure 3
<p>Recycling experiment of DTPA-catalyzed N-formylation reaction of CO<sub>2</sub> with aniline.</p>
Full article ">Figure 4
<p>Infrared comparison spectra of DTPA before and after 5 times of reuse.</p>
Full article ">Figure 5
<p>Reaction mechanism of CO<sub>2</sub> with aniline N-formylation.</p>
Full article ">
23 pages, 7459 KiB  
Article
Photophysical Characterization and In Vitro Evaluation of α-Mangostin-Loaded HDL Mimetic Nano-Complex in LN-229 Glioblastoma Spheroid Model
by Ammar Kapic, Nirupama Sabnis, Akpedje S. Dossou, Jose Chavez, Luca Ceresa, Zygmunt Gryczynski, Rafal Fudala, Rob Dickerman, Bruce A. Bunnell and Andras G. Lacko
Int. J. Mol. Sci. 2024, 25(13), 7378; https://doi.org/10.3390/ijms25137378 - 5 Jul 2024
Viewed by 858
Abstract
Cytotoxic activity has been reported for the xanthone α-mangostin (AMN) against Glioblastoma multiforme (GBM), an aggressive malignant brain cancer with a poor prognosis. Recognizing that AMN’s high degree of hydrophobicity is likely to limit its systemic administration, we formulated AMN using reconstituted high-density [...] Read more.
Cytotoxic activity has been reported for the xanthone α-mangostin (AMN) against Glioblastoma multiforme (GBM), an aggressive malignant brain cancer with a poor prognosis. Recognizing that AMN’s high degree of hydrophobicity is likely to limit its systemic administration, we formulated AMN using reconstituted high-density lipoprotein (rHDL) nanoparticles. The photophysical characteristics of the formulation, including fluorescence lifetime and steady-state anisotropy, indicated that AMN was successfully incorporated into the rHDL nanoparticles. To our knowledge, this is the first report on the fluorescent characteristics of AMN with an HDL-based drug carrier. Cytotoxicity studies in a 2D culture and 3D spheroid model of LN-229 GBM cells and normal human astrocytes showed an enhanced therapeutic index with the rHDL-AMN formulation compared to the unincorporated AMN and Temozolomide, a standard GBM chemotherapy agent. Furthermore, treatment with the rHDL-AMN facilitated a dose-dependent upregulation of autophagy and reactive oxygen species generation to a greater extent in LN-229 cells compared to astrocytes, indicating the reduced off-target toxicity of this novel formulation. These studies indicate the potential therapeutic benefits to GBM patients via selective targeting using the rHDL-AMN formulation. Full article
(This article belongs to the Special Issue New Insights into Glioblastoma: Cellular and Molecular (2nd Edition))
Show Figures

Figure 1

Figure 1
<p>Fluorescence characterization of α-mangostin (AMN). (<b>A</b>): Excitation and emission spectra of AMN. (<b>B</b>): Anisotropy (blue line) overlaid on top of the emission spectra. The experiment was repeated at least 3 times.</p>
Full article ">Figure 2
<p>Fluorescence lifetime measurements (FLT) of AMN formulations. (<b>A</b>) FLT curve of free AMN; (<b>B</b>) The components of the FLT curve of free AMN; (<b>C</b>) FLT curve of rHDL-AMN NPs in 1X PBS; (<b>D</b>) The components of the FLT curve of rHDL-AMN NPs; (<b>E</b>) FLT curve of the disrupted of rHDL-AMN NPs in DMSO; and (<b>F</b>) The FLT components of the DMSO-disrupted rHDL-AMN NPs. The experiment was repeated at least 3 times.</p>
Full article ">Figure 3
<p>Cytotoxic effect of the free AMN, rHDL-AMN NPs, and TMZ on LN-229 GBM cells and astrocytes. (<b>A</b>,<b>B</b>) Percent survival featuring free AMN and rHDL-AMN NPs in 2D models of LN-229 and astrocytes, respectively. Statistical significance was conducted using a 2-way ANOVA with a post-hoc Tukey test (* <span class="html-italic">p</span> Value = 0.0224, ** <span class="html-italic">p</span> value = &lt;0.0001). (<b>C</b>,<b>D</b>) Percent survival featuring free AMN and rHDL-AMN NPs in 3D models of LN-229 and astrocytes, respectively. Statistical significance was conducted using a 2-way ANOVA with a post-hoc Tukey test (** <span class="html-italic">p</span> value = &lt;0.0001). (<b>E</b>) Representative images of LN-229 spheroids treated with free AMN or rHDL-AMN NPs. (<b>F</b>) Representative images of astrocyte spheroids treated with free AMN or rHDL-AMN. (<b>G</b>,<b>H</b>) Percent survival of TMZ in the 2D and 3D models conducted using LN-229 cells and astrocytes. Statistical significance was conducted using a 2-way ANOVA with a post-hoc Tukey test (# <span class="html-italic">p</span> values = 1.25 × 10<sup>−4</sup>, ** <span class="html-italic">p</span> value = &lt;0.0001). (<b>I</b>) Therapeutic indices of free AMN, rHDL-AMN, and TMZ with LN-229 cells and astrocytes under given conditions. Statistical significance was conducted using 1-way ANOVA (* <span class="html-italic">p</span> values &lt; 0.0001, ** <span class="html-italic">p</span> value = 3.68 × 10<sup>−4</sup>, *** <span class="html-italic">p</span> value = 2.217 × 10<sup>−4</sup>). Measurements are the average of 3 independent experiments.</p>
Full article ">Figure 3 Cont.
<p>Cytotoxic effect of the free AMN, rHDL-AMN NPs, and TMZ on LN-229 GBM cells and astrocytes. (<b>A</b>,<b>B</b>) Percent survival featuring free AMN and rHDL-AMN NPs in 2D models of LN-229 and astrocytes, respectively. Statistical significance was conducted using a 2-way ANOVA with a post-hoc Tukey test (* <span class="html-italic">p</span> Value = 0.0224, ** <span class="html-italic">p</span> value = &lt;0.0001). (<b>C</b>,<b>D</b>) Percent survival featuring free AMN and rHDL-AMN NPs in 3D models of LN-229 and astrocytes, respectively. Statistical significance was conducted using a 2-way ANOVA with a post-hoc Tukey test (** <span class="html-italic">p</span> value = &lt;0.0001). (<b>E</b>) Representative images of LN-229 spheroids treated with free AMN or rHDL-AMN NPs. (<b>F</b>) Representative images of astrocyte spheroids treated with free AMN or rHDL-AMN. (<b>G</b>,<b>H</b>) Percent survival of TMZ in the 2D and 3D models conducted using LN-229 cells and astrocytes. Statistical significance was conducted using a 2-way ANOVA with a post-hoc Tukey test (# <span class="html-italic">p</span> values = 1.25 × 10<sup>−4</sup>, ** <span class="html-italic">p</span> value = &lt;0.0001). (<b>I</b>) Therapeutic indices of free AMN, rHDL-AMN, and TMZ with LN-229 cells and astrocytes under given conditions. Statistical significance was conducted using 1-way ANOVA (* <span class="html-italic">p</span> values &lt; 0.0001, ** <span class="html-italic">p</span> value = 3.68 × 10<sup>−4</sup>, *** <span class="html-italic">p</span> value = 2.217 × 10<sup>−4</sup>). Measurements are the average of 3 independent experiments.</p>
Full article ">Figure 4
<p>Payload uptake of cargo via rHDL using Nile Red (40× magnification). (<b>A</b>) Representative images of the NR uptake in the LN-229 and astrocyte cells incubated with the free NR and rHDL-NR NPs at 1 µM concentration for 1 h at 37 °C; (<b>B</b>) Comparison of average NR MFI of the LN-229 and astrocyte cells incubated with the free NR and rHDL-NR NPs. The images and data are representative of three independent experiments and the data are graphed as the mean ± SD. (one-way ANOVA with a post-hoc Tukey test ** <span class="html-italic">p</span> value ≤ 0.001).</p>
Full article ">Figure 5
<p>Effect of AMN formulations on autophagy in LN-229 cells and astrocytes. (<b>A</b>) LN-229 and (<b>B</b>) astrocyte cells were treated with free AMN and rHDL-AMN for 48 h. Using the fluorescence intensity, tamoxifen was used as a positive control to calculate the percentage autophagy in each cell line. Percentage autophagy was calculated considering the positive control as 100%. Statistical significance was conducted using a 2-way ANOVA with a post-hoc Tukey test (* <span class="html-italic">p</span> values = 0.0487, ** <span class="html-italic">p</span> value = &lt;0.0001). Measurements are the average of 3 independent experiments.</p>
Full article ">Figure 6
<p>Effect of free AMN, TMZ, and rHDL-AMN NPs at 10 µM drug on the migration of LN-229 cells at 0, 12, and 24 h using a scratch wound assay. (<b>A</b>) Representative images (4× magnification) of migration activity after treatment. Scale bars show a length of 1000 micrometers (<b>B</b>) and (<b>C</b>) Mean ratio of the recovered area at 12 and 24 h respectively using Image J 2.15.0 software. Statistical significance was conducted using a 1-way ANOVA with a post-hoc Tukey test (For (<b>B</b>) * <span class="html-italic">p</span> values = 0.003, ** <span class="html-italic">p</span> value = 9.05 × 10<sup>−4</sup>, *** <span class="html-italic">p</span> value 1.15 × 10<sup>−4</sup>, **** <span class="html-italic">p</span> value 3.05 × 10<sup>−4</sup>). Statistical significance was conducted using a 1-way ANOVA with a post-hoc Tukey test (For (<b>C</b>) * <span class="html-italic">p</span> values ≤ 0.0001). Measurements are the average of 3 independent experiments.</p>
Full article ">Figure 7
<p>Effect of AMN formulations on reactive oxygen species (ROS activity) in (<b>A</b>) LN-229 and (<b>B</b>) astrocytes. The ROS activity was measured using the fluorescence of 2′,7′–dichlorofluorescein diacetate (DCFDA) in LN-229 and astrocytes after 24 h treatment with rHDL-AMN and free AMN. Tert-butyl hydroperoxide (TBHP) was used as a positive control. Net fluorescence units were measured at excitation/emission 485 nm/535 nm. Statistical significance was conducted using a 2-way ANOVA with a post-hoc Tukey test (* <span class="html-italic">p</span> value ≤ 0.0001). Measurements are the average of 3 independent experiments.</p>
Full article ">Figure 8
<p>A schematic representation of the rHDL-AMN nanoparticle synthesis protocol and characterization process.</p>
Full article ">
42 pages, 4030 KiB  
Review
A Review of Traditional Applications, Geographic Distribution, Botanical Characterization, Phytochemistry, and Pharmacology of Hypericum ascyron L.
by Meihui Liu, Yongmei Zhou, Xiaoxiao Rui, Zi Ye, Linyu Zheng, Hao Zang and Yuan Zhong
Horticulturae 2024, 10(6), 555; https://doi.org/10.3390/horticulturae10060555 - 25 May 2024
Viewed by 543
Abstract
Hypericum ascyron L. (H. ascyron) is a significant medicinal plant traditionally used for various conditions like hematemesis, hemoptysis, injuries from falls, irregular menses, dysmenorrhea, and liver fire-induced headaches. A comprehensive literature search was conducted using databases like SciFinder and Web of [...] Read more.
Hypericum ascyron L. (H. ascyron) is a significant medicinal plant traditionally used for various conditions like hematemesis, hemoptysis, injuries from falls, irregular menses, dysmenorrhea, and liver fire-induced headaches. A comprehensive literature search was conducted using databases like SciFinder and Web of Science to explore its traditional uses, geographical distribution, botanical description, phytochemistry, and pharmacology. The objective of this review is to lay groundwork and suggest fresh avenues of investigation into the possible uses of the plant. Currently, two hundred and seventy compounds have been isolated and identified from H. ascyron, including phloroglucinols, xanthones, flavonoids, phenolics, steroids and triterpenoids, volatile components, and other compounds. Notably, phloroglucinols, xanthones, and flavonoids have exhibited remarkable pharmacological effects like antioxidant, antidiabetic, anti-inflammatory, antidepressant, cytotoxic, and antimicrobial activities. Despite extensive research, further studies are needed to understand new components and mechanisms of action, requiring more detailed investigations. This thorough exploration could facilitate the advancement and utilization of H. ascyron. Full article
(This article belongs to the Section Medicinals, Herbs, and Specialty Crops)
Show Figures

Figure 1

Figure 1
<p>Morphology of <span class="html-italic">Hypericum ascyron</span> L.: aboveground part (<b>A</b>) and root (<b>B</b>).</p>
Full article ">Figure 2
<p>The general geographical distribution of <span class="html-italic">Hypericum ascyron</span> L. in China.</p>
Full article ">Figure 3
<p>Chemical structures of phloroglucinols isolated or identified from <span class="html-italic">Hypericum ascyron</span> L.</p>
Full article ">Figure 3 Cont.
<p>Chemical structures of phloroglucinols isolated or identified from <span class="html-italic">Hypericum ascyron</span> L.</p>
Full article ">Figure 3 Cont.
<p>Chemical structures of phloroglucinols isolated or identified from <span class="html-italic">Hypericum ascyron</span> L.</p>
Full article ">Figure 3 Cont.
<p>Chemical structures of phloroglucinols isolated or identified from <span class="html-italic">Hypericum ascyron</span> L.</p>
Full article ">Figure 3 Cont.
<p>Chemical structures of phloroglucinols isolated or identified from <span class="html-italic">Hypericum ascyron</span> L.</p>
Full article ">Figure 4
<p>Chemical structures of xanthones and dibenzo-1,4-dioxane derivatives isolated or identified from <span class="html-italic">Hypericum ascyron</span> L.</p>
Full article ">Figure 4 Cont.
<p>Chemical structures of xanthones and dibenzo-1,4-dioxane derivatives isolated or identified from <span class="html-italic">Hypericum ascyron</span> L.</p>
Full article ">Figure 5
<p>Chemical structures of flavonoids isolated or identified from <span class="html-italic">Hypericum ascyron</span> L.</p>
Full article ">Figure 6
<p>Chemical structures of phenolics isolated or identified from <span class="html-italic">Hypericum ascyron</span> L.</p>
Full article ">Figure 7
<p>Chemical structures of steroids and triterpenoids isolated or identified from <span class="html-italic">Hypericum ascyron</span> L.</p>
Full article ">Figure 7 Cont.
<p>Chemical structures of steroids and triterpenoids isolated or identified from <span class="html-italic">Hypericum ascyron</span> L.</p>
Full article ">Figure 8
<p>Chemical structures of volatile components isolated or identified from <span class="html-italic">Hypericum ascyron</span> L.</p>
Full article ">Figure 8 Cont.
<p>Chemical structures of volatile components isolated or identified from <span class="html-italic">Hypericum ascyron</span> L.</p>
Full article ">Figure 9
<p>Chemical structures of other components isolated or identified from <span class="html-italic">Hypericum ascyron</span> L.</p>
Full article ">
22 pages, 3459 KiB  
Article
Spasmolytic, Antimicrobial, and Antioxidant Activities of Spray-Dried Extracts of Gentiana asclepiadea L. with In Silico Pharmacokinetic Analysis
by Miloš S. Jovanović, Milica Milutinović, Suzana Branković, Tatjana Mihajilov-Krstev, Milica Randjelović, Bojana Miladinović, Nada Ćujić Nikolić, Katarina Šavikin and Dušanka Kitić
Plants 2024, 13(11), 1445; https://doi.org/10.3390/plants13111445 - 23 May 2024
Viewed by 933
Abstract
This study aimed to evaluate the spasmolytic activity of an underground parts extract of Gentiana asclepiadea L. (Gentianaceae), assess its antioxidant and antimicrobial activities, and explore the impact of extract encapsulation on the aforementioned bioactivities. An extract encapsulated by spray drying with whey [...] Read more.
This study aimed to evaluate the spasmolytic activity of an underground parts extract of Gentiana asclepiadea L. (Gentianaceae), assess its antioxidant and antimicrobial activities, and explore the impact of extract encapsulation on the aforementioned bioactivities. An extract encapsulated by spray drying with whey protein, pure extract, and pure whey protein were comparatively tested. The main compounds identified via HPLC-DAD analysis underwent in silico ADME assessment. The spasmolytic effect was tested on a model of spontaneous rat ileum contractions, and the mechanism of action was further evaluated on acetylcholine-, KCl-, CaCl2-, BaCl2-, histamine-, N(ω)-nitro-L-arginine methyl ester-, and glibenclamide-modified contractions. The most abundant compounds were secoiridoids (dominantly gentiopicroside), followed by C-glycosylated flavonoids and xanthones. Both pure and encapsulated extracts achieved significant spasmolytic effects, despite the spasmogenic activity of pure whey protein. The extract may exert its spasmolytic effect through multiple pathways, predominantly by antagonizing the Ca2+ channel and opening the K+ channel, while the nitric oxide pathway appears not to be involved. The antimicrobial and antioxidant activities of the pure extract were moderate. The extract stabilized by encapsulation retained all of the tested bioactivities of the unencapsulated extract. The obtained results suggest that G. asclepiadea has potential for use in the treatment of some gastrointestinal complaints and that the encapsulated extract could be a valuable functional ingredient in pharmaceutical and food products. Full article
(This article belongs to the Special Issue Plant-Derived Natural Products: Development and Utilization)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Effects of <span class="html-italic">Gentiana asclepiadea</span> L. pure extract (E), extract microencapsulated with whey protein (EWP), pure whey protein (WP), and papaverine on spontaneous contractions of the rat ileum. (* <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 indicate significant differences compared to spontaneous contractions in Tyrode’s solution according to Student’s <span class="html-italic">t</span>-test).</p>
Full article ">Figure 2
<p>Inhibitory effects of <span class="html-italic">Gentiana asclepiadea</span> L. pure extract (E) and extract microencapsulated with whey protein (EWP) on the acetylcholine (Ach)-induced contractions of the rat ileum: (<b>A</b>) the values of control, Ach + E (0.5 mg/mL), and Ach + E (1.5 mg/mL); (<b>B</b>) the values of control, Ach + EWP (0.6 mg/mL), and Ach + EWP (1.9 mg/mL); and (<b>C</b>) the values of control and Ach + positive control atropine (140 nmol/L). (* <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 vs. control, according to the Student’s <span class="html-italic">t</span>-test).</p>
Full article ">Figure 3
<p>Inhibitory effects of <span class="html-italic">Gentiana asclepiadea</span> L. pure extract (E), extract encapsulated with whey protein (EWP), and verapamil on the potassium chloride (KCl)-induced contractions of the rat ileum. (* <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 vs. control, according to the Student’s <span class="html-italic">t</span>-test).</p>
Full article ">Figure 4
<p>Inhibitory effects of <span class="html-italic">Gentiana asclepiadea</span> L. pure extract (E) and extract microencapsulated with whey protein (EWP) on the calcium chloride (CaCl<sub>2</sub>)-induced contractions of the rat ileum: (<b>A</b>) the values of control, CaCl<sub>2</sub> + E (0.5 mg/mL), and CaCl<sub>2</sub> + E (1.5 mg/mL); (<b>B</b>) the values of control, CaCl<sub>2</sub> + EWP (0.6 mg/mL), and CaCl<sub>2</sub> + EWP (1.9 mg/mL); and (<b>C</b>) the values of control and CaCl<sub>2</sub> + positive control verapamil (0.3 µm). (* <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 vs. control, according to the Student’s <span class="html-italic">t</span>-test).</p>
Full article ">Figure 5
<p>Inhibitory effects of <span class="html-italic">Gentiana asclepiadea</span> L. pure extract (E) and extract microencapsulated with whey protein (EWP) on the barium chloride (BaCl<sub>2</sub>)-induced contractions of the rat ileum: (<b>A</b>) the values of control, BaCl<sub>2</sub> + E (0.5 mg/mL), and BaCl<sub>2</sub> + E (1.5 mg/mL); and (<b>B</b>) the values of control, BaCl<sub>2</sub> + EWP (0.6 mg/mL), and BaCl<sub>2</sub> + EWP (1.9 mg/mL). (** <span class="html-italic">p</span> &lt; 0.01 vs. control, according to the Student’s <span class="html-italic">t</span>-test).</p>
Full article ">Figure 6
<p>Inhibitory effects of <span class="html-italic">Gentiana asclepiadea</span> L. pure extract (E) and extract microencapsulated with whey protein (EWP) on the histamine-induced contractions of the rat ileum: (<b>A</b>) the values of control, histamine + E (0.5 mg/mL), and histamine + E (1.5 mg/mL); and (<b>B</b>) the values of control, histamine + EWP (0.6 mg/mL), and histamine + EWP (1.9 mg/mL). (** <span class="html-italic">p</span> &lt; 0.01 vs. control, according to the Student’s <span class="html-italic">t</span>-test).</p>
Full article ">Figure 7
<p>Effects of <span class="html-italic">Gentiana asclepiadea</span> L. pure extract (E) and extract microencapsulated with whey protein (EWP) on the rat ileum contractions in the presence of N(<span class="html-italic">ω</span>)-nitro-L-arginine methyl ester (L-NAME): (<b>A</b>) increasing concentration of E (control) and increasing concentration of E + L-NAME; and (<b>B</b>) increasing concentration of EWP (control) and increasing concentration of EWP + L-NAME.</p>
Full article ">Figure 8
<p>Effects of <span class="html-italic">Gentiana asclepiadea</span> L. pure extract (E) and extract microencapsulated with whey protein (EWP) on the rat ileum contractions induced by KCl (25 mmol/L) in the presence of glibenclamide: (<b>A</b>) KCl + E (control) and KCl + E + glibenclamide; and (<b>B</b>) KCl + EWP (control) and KCl + EWP + glibenclamide. (* <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 vs. control, according to the Student’s <span class="html-italic">t</span>-test).</p>
Full article ">Figure 9
<p>The chemical structures of the assessed compounds from the underground parts of <span class="html-italic">Gentiana asclepiadea</span>, along with bioavailability radars displaying physicochemical parameters relevant to the prediction of their oral bioavailability.</p>
Full article ">
22 pages, 2276 KiB  
Review
Molecular Pathways of Genistein Activity in Breast Cancer Cells
by Evangelia K. Konstantinou, Aristea Gioxari, Maria Dimitriou, George I. Panoutsopoulos and Athanasios A. Panagiotopoulos
Int. J. Mol. Sci. 2024, 25(10), 5556; https://doi.org/10.3390/ijms25105556 - 20 May 2024
Cited by 6 | Viewed by 1664
Abstract
The most common malignancy in women is breast cancer. During the development of cancer, oncogenic transcription factors facilitate the overproduction of inflammatory cytokines and cell adhesion molecules. Antiapoptotic proteins are markedly upregulated in cancer cells, which promotes tumor development, metastasis, and cell survival. [...] Read more.
The most common malignancy in women is breast cancer. During the development of cancer, oncogenic transcription factors facilitate the overproduction of inflammatory cytokines and cell adhesion molecules. Antiapoptotic proteins are markedly upregulated in cancer cells, which promotes tumor development, metastasis, and cell survival. Promising findings have been found in studies on the cell cycle-mediated apoptosis pathway for medication development and treatment. Dietary phytoconstituents have been studied in great detail for their potential to prevent cancer by triggering the body’s defense mechanisms. The underlying mechanisms of action may be clarified by considering the role of polyphenols in important cancer signaling pathways. Phenolic acids, flavonoids, tannins, coumarins, lignans, lignins, naphthoquinones, anthraquinones, xanthones, and stilbenes are examples of natural chemicals that are being studied for potential anticancer drugs. These substances are also vital for signaling pathways. This review focuses on innovations in the study of polyphenol genistein’s effects on breast cancer cells and presents integrated chemical biology methods to harness mechanisms of action for important therapeutic advances. Full article
(This article belongs to the Special Issue Phenolic Compounds in Human Diseases)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Chemical structure of genistein.</p>
Full article ">Figure 2
<p>Biosynthetic pathway of genistein.</p>
Full article ">Figure 3
<p>The mechanisms of action of genistein in BC cells. By modifying Bcl-2 family proteins, genistein triggers apoptosis via a mitochondrial-mediated, classical caspase-dependent mechanism. Altering the cycle regulating proteins causes cell cycle arrest. It deactivates the MAPK (ERK1/2) and PI3K/AKT signaling pathways. In addition, genistein controls epigenetic regulation, inhibits angiogenesis, invasion, and cell migration, and modifies the expression of numerous miRNAs. Genistein raises the Bax/Bcl-2 ratio, causing apoptosis through autophagy-dependent pathways, and preventing oxidative stress by altering the expression of antioxidant enzymes. Additionally, genistein inhibits cell proliferation by downregulating CCNG1 GADD45A, NF-κB, Bcl-2, TNFR, ESR1, NCOA2, and NCOA3, and upregulating genes like p53 and CDKN1A. Research conducted in vitro has demonstrated that GNT can reduce tumorigenic processes by upregulating the expression and activity of the GSTP1 and RARβ2 genes. Through the downregulation of the proteins COX, TPA, and EROD, genistein can also inhibit angiogenesis. Fis1 and Opa1 mRNA expression can be decreased by genistein through mitochondrial-dependent pathways, according to in vitro research [<a href="#B211-ijms-25-05556" class="html-bibr">211</a>]. In figure, the symbol (↑) indicates an increase and the symbol (↓) indicates a decrease.</p>
Full article ">
15 pages, 1703 KiB  
Review
Xanthone Derivatives and Their Potential Usage in the Treatment of Telangiectasia and Rosacea
by Katarzyna Brezdeń and Anna M. Waszkielewicz
Appl. Sci. 2024, 14(10), 4037; https://doi.org/10.3390/app14104037 - 9 May 2024
Viewed by 873
Abstract
Xanthone derivatives, a class of natural compounds abundantly found in plants such as mangosteen (Garcinia mangostana) and certain herbs, have garnered substantial interest due to their diverse pharmacological properties, including antioxidant, anti-inflammatory, and anti-cancer activities. Recent investigations have unveiled their potential [...] Read more.
Xanthone derivatives, a class of natural compounds abundantly found in plants such as mangosteen (Garcinia mangostana) and certain herbs, have garnered substantial interest due to their diverse pharmacological properties, including antioxidant, anti-inflammatory, and anti-cancer activities. Recent investigations have unveiled their potential as modulators of enzymatic activity, prompting exploration into their effects on hyaluronidase-mediated hyaluronic acid (HA) degradation, and their effects in topical treatment of telangiectasia and rosacea. Telangiectasia and rosacea are common dermatological conditions characterized by chronic skin inflammation, vascular abnormalities, and visible blood vessels, resulting in significant cosmetic concerns and impaired quality of life for affected individuals. This review aims to provide a comprehensive overview of the current understanding regarding the interplay between the mechanisms of action by which xanthone derivatives exert their therapeutic effects, including the inhibition of pro-inflammatory cytokines, modulation of oxidative stress pathways, and regulation of vascular endothelial growth factors. Furthermore, we will discuss the implications of harnessing xanthone derivatives as therapeutic agents for mitigating vascular dysfunction and its associated pathologies, thereby offering insights into future research directions and therapeutic strategies in the field of vascular biology. Full article
(This article belongs to the Special Issue Development of Innovative Cosmetics)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Structure of xanthone (dibenzo-γ-pyrone).</p>
Full article ">Figure 2
<p>Structure of gambogic acid: (Z)-4-((1S,3aR,5S,11R,14aS)-8-hydroxy-2,2,11-trimethyl-13-(3-methylbut-2-en-1-yl)-11-(4-methylpent-3-en-1-yl)-4,7-dioxo-1,2,5,7-tetrahydro-11H-1,5-methanofuro[3,2-g]pyrano[3,2-b]xanthen-3a(4H)-yl)-2-methylbut-2-enoic acid).</p>
Full article ">Figure 3
<p>Structure of α-mangostin (1,3,6-trihydroxy-7-methoxy-2,8-bis(3-methylbut-2-en-1-yl)-xanthone).</p>
Full article ">Figure 4
<p>Structure of mangiferin: 1,3,6,7-tetrahydroxy-2-((2S,3R,4R,5S,6R)-3,4,5-trihydroxy-6-(hydroxymethyl)tetrahydro-2H-pyran-2-yl)-xanthone.</p>
Full article ">Figure 5
<p>Davitin A (1,8-dihydroxy-2,5,6-trimethoxyxanthone) structure.</p>
Full article ">Figure 6
<p>Structure of <math display="inline"><semantics> <mrow> <mi mathvariant="bold-italic">γ</mi> </mrow> </semantics></math>-mangostin (1,3,6,7-tetrahydroxy-2,8-bis(3-methylbut-2-en-1-yl)-xanthone).</p>
Full article ">Figure 7
<p>Structures of xanthone derivatives isolated from <span class="html-italic">Garcinia delpyana</span>.</p>
Full article ">Figure 8
<p>Structure of mangostanin ((R)-4,8-dihydroxy-2-(2-hydroxypropan-2-yl)-7-methoxy-6-(3-methylbut-2-en-1-yl)-2,3-dihydro-5H-furo[3,2-b]xanthen-5-one).</p>
Full article ">Scheme 1
<p>Mechanism of action of hyaluronidase. Adapted from Lee A. et al. [<a href="#B11-applsci-14-04037" class="html-bibr">11</a>] and Ponnuraj K. et al. [<a href="#B12-applsci-14-04037" class="html-bibr">12</a>].</p>
Full article ">
15 pages, 2031 KiB  
Article
Unraveling a Historical Mystery: Identification of a Lichen Dye Source in a Fifteenth Century Medieval Tapestry
by Rachel M. Lackner, Solenn Ferron, Joël Boustie, Françoise Le Devehat, H. Thorsten Lumbsch and Nobuko Shibayama
Heritage 2024, 7(5), 2370-2384; https://doi.org/10.3390/heritage7050112 - 1 May 2024
Cited by 1 | Viewed by 1298
Abstract
As part of a long-term campaign to document, study, and conserve the Heroes tapestries from The Cloisters collection at The Metropolitan Museum of Art, organic colorant analysis of Julius Caesar (accession number 47.101.3) was performed. Analysis with liquid chromatography–quadrupole time-of-flight mass spectrometry (LC-qToF-MS) [...] Read more.
As part of a long-term campaign to document, study, and conserve the Heroes tapestries from The Cloisters collection at The Metropolitan Museum of Art, organic colorant analysis of Julius Caesar (accession number 47.101.3) was performed. Analysis with liquid chromatography–quadrupole time-of-flight mass spectrometry (LC-qToF-MS) revealed the presence of several multiply chlorinated xanthones produced only by certain species of lichen. Various lichen dye sources have been documented in the literature for centuries and are classified as either ammonia fermentation method (AFM) or boiling water method (BWM) dyes based on their method of production. However, none of these known sources produce the distinctive metabolites present in the tapestry. LC-qToF-MS was also used to compare the chemical composition of the dyes in the tapestry with that of several species of crustose lichen. Lichen metabolites, including thiophanic acid and arthothelin, were definitively identified in the tapestry based on comparison with lichen xanthone standards and a reference of Lecanora sulphurata, confirming the presence of a lichen source. This finding marks the first time that lichen xanthones have been identified in a historic object and the first evidence that BWM lichen dyes may have been used prior to the eighteenth century. Full article
(This article belongs to the Special Issue Dyes in History and Archaeology 42)
Show Figures

Figure 1

Figure 1
<p>Several classes of secondary metabolites found in lichen.</p>
Full article ">Figure 2
<p><span class="html-italic">Julius Caesar (from the Heroes Tapestries);</span> wool warp, wool wefts; South Netherlandish, ca. 1400–1410; 165 1/2 × 93 11/16 in. (420.4 × 238 cm); The Metropolitan Museum of Art, New York, Gift of John D. Rockefeller Jr., 1947 (47.101.3). Image © The Metropolitan Museum of Art. The following annotations to the original photograph were added by the author: the individual fragments from which the tapestry was recreated are numbered. White boxes correspond to areas from which the dark brown fibers were sampled.</p>
Full article ">Figure 3
<p>Basic structure of lichen metabolites derived from the norlichexanthone scaffold.</p>
Full article ">Figure 4
<p>Isotopic pattern, exact mass and [M − H]<sup>−</sup> formula, possible chemical structure, and number of possible isomers for the components at 22.7, 23.8, 24.5, and 25.0 min: (<b>a</b>), 25.8 min (<b>b</b>), and 26.4 and 27.5 min (<b>c</b>).</p>
Full article ">Figure 5
<p>Extracted ion chromatogram (EIC) for colorants and lichen xanthones identified in the dark brown samples from fragments 3 (<b>a</b>), 4 (<b>b</b>), and 5 (<b>c</b>). The components are numbered corresponding to <a href="#heritage-07-00112-t001" class="html-table">Table 1</a>. The colors for compounds <b>13</b>–<b>15</b> and <b>17</b>–<b>19</b> correspond to the chlorinated xanthones in <a href="#heritage-07-00112-f004" class="html-fig">Figure 4</a>. The white arrows in the image indicate the location from where the samples were taken (see <a href="#heritage-07-00112-f002" class="html-fig">Figure 2</a>).</p>
Full article ">Figure 6
<p>UV-visible spectrum (a) and MS/MS fragmentation spectrum (b) for thiophanic acid, <b>17</b>.</p>
Full article ">Figure 7
<p>Extracted ion chromatogram (EIC) of acetone extract of <span class="html-italic">Lecanora sulphurata</span>.</p>
Full article ">
19 pages, 5939 KiB  
Article
Metabolomics-Guided Discovery of New Dimeric Xanthones from Co-Cultures of Mangrove Endophytic Fungi Phomopsis asparagi DHS-48 and Phomopsis sp. DHS-11
by Jingwan Wu, Dandan Chen, Qing Li, Ting Feng and Jing Xu
Mar. Drugs 2024, 22(3), 102; https://doi.org/10.3390/md22030102 - 23 Feb 2024
Cited by 3 | Viewed by 1910
Abstract
The co-culture strategy, which mimics natural ecology by constructing an artificial microbial community, is a useful tool for the activation of biosynthetic gene clusters (BGCs) to generate new metabolites, as well as to increase the yield of respective target metabolites. As part of [...] Read more.
The co-culture strategy, which mimics natural ecology by constructing an artificial microbial community, is a useful tool for the activation of biosynthetic gene clusters (BGCs) to generate new metabolites, as well as to increase the yield of respective target metabolites. As part of our project aiming at the discovery of structurally novel and biologically active natural products from mangrove endophytic fungi, we selected the co-culture of a strain of Phomopsis asparagi DHS-48 with another Phomopsis genus fungus DHS-11, both endophyted in mangrove Rhizophora mangle considering the impart of the taxonomic criteria and ecological data. The competition interaction of the two strains was investigated through morphology observation and scanning electron microscopy (SEM), and it was found that the mycelia of the DHS-48 and DHS-11 compacted and tangled with each other with an interwoven pattern in the co-culture system. A new approach that integrates HPLC chromatogram, 1HNMR spectroscopy, UPLC-MS-PCA, and molecular networking enabled the targeted isolation of the induced metabolites, including three new dimeric xanthones phomoxanthones L-N (13), along with six known analogs (49). Their planar structures were elucidated by an analysis of their HRMS, MS/MS, and NMR spectroscopic data and the absolute configurations based on ECD calculations. These metabolites showed broad cytotoxic activity against the cancer cells assessed, of which compounds 79 displayed significant cytotoxicity towards human liver cells HepG-2 with IC50 values ranging from 4.83 μM to 12.06 μM. Compounds 16 exhibited weak immunosuppressive activity against the proliferation of ConA-induced (T-cell) and LPS-induced (B-cell) murine splenic lymphocytes. Therefore, combining co-cultivation with a metabolomics-guided strategy as a discovery tool will be implemented as a systematic strategy for the quick discovery of target bioactive compounds. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Structures of the isolated compounds <b>1</b>–<b>9</b>.</p>
Full article ">Figure 2
<p>Mycelia morphology observation of <span class="html-italic">Phomopsis asparagi</span> DHS-48 and <span class="html-italic">Phomopsis</span> sp. DHS-11 in PDA medium’s co-culture. Colony morphology of (<b>A</b>) DHS-48, (<b>B</b>) DHS-11, (<b>C</b>) DHS-48, and DHS-11 in co-culture. Scanning electron micrographs of (<b>D</b>) DHS-48, (<b>E</b>) DHS-11, (<b>F</b>) DHS-48, and DHS-11 in co-culture (<b>F</b>). Enlarged areas of (<b>G</b>). The red arrow points to strain DHS-48, the blue arrow points to strain DHS-11, and the yellow arrow points to the junction of the two species.</p>
Full article ">Figure 3
<p>HPLC chromatograms of the EtOAc extracts deriving from (<b>A</b>) the whole co-culture of DHS-48 and DHS-11 and the monocultures of (<b>B</b>) DHS-48 and (<b>C</b>) DHS-11. * Compounds <b>1</b>–<b>3</b> in (<b>A</b>) represent the new compounds stimulated by the co-culture.</p>
Full article ">Figure 4
<p>Molecular network (MN) of DHS-48 (green), DHS-11 (yellow), and their co-culture (red). (<b>A</b>) All 33 clusters containing at least two nodes are numbered. Thickness of the edges between nodes indicates the degree of similarity between their respective MS/MS spectra. (<b>B</b>) Enlarged cluster corresponding to xantone dimers (<b>1</b>–<b>9</b>) generated by GNPS, of which compounds <b>3</b>, <b>5</b>–<b>9</b> are annotated in cluster 1, while compounds <b>1</b>, <b>2</b>, and <b>4</b> were displayed as singletons in the MN. * represent the new compounds <b>1</b>–<b>3</b> stimulated by the co-culture.</p>
Full article ">Figure 5
<p>Key COSY and HMBC correlations of compounds <b>1</b>–<b>3</b>.</p>
Full article ">Figure 6
<p>Key NOESY correlations of compounds <b>1</b>–<b>3</b>.</p>
Full article ">Figure 7
<p>Experimental and calculated electronic circular dichroism (ECD) spectra of <b>1</b>–<b>3</b>.</p>
Full article ">Scheme 1
<p>Workflow of targeted isolation of new dimeric xanthones from co-cultures of mangrove endophytic fungi <span class="html-italic">Phomopsis asparagi</span> DHS-48 and <span class="html-italic">Phomopsis</span> sp. DHS-11 based on integrated metabolomics-guided discovery.</p>
Full article ">
Back to TopTop