[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

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
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
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

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
remove_circle_outline
remove_circle_outline

Search Results (10,673)

Search Parameters:
Keywords = antibacterial activity

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
16 pages, 1563 KiB  
Article
Enhancing Antibiotic Efficacy and Combating Biofilm Formation: Evaluating the Synergistic Potential of Origanum vulgare Essential Oil against Multidrug-Resistant Gram-Negative Bacteria
by Bilal Saoudi, Karim Bariz, Sarah Saci, Yousra Belounis, Hakima Ait Issad, Mohamed Abbaci, Mohamed Abou Mustapha, El-Hafid Nabti, Rawaf Alenazy, Mohammed Sanad Alhussaini, Abdulrahman A. I. Alyahya, Mohammed Alqasmi, Maryam S. Alhumaidi, Fawaz M. Almufarriji and Karim Houali
Microorganisms 2024, 12(8), 1651; https://doi.org/10.3390/microorganisms12081651 - 12 Aug 2024
Abstract
Multidrug-resistant (MDR) Gram-negative bacteria remain a global public health issue due to the barrier imposed by their outer membrane and their propensity to form biofilms. It is becoming imperative to develop new antibacterial strategies. In this context, this study aims to evaluate the [...] Read more.
Multidrug-resistant (MDR) Gram-negative bacteria remain a global public health issue due to the barrier imposed by their outer membrane and their propensity to form biofilms. It is becoming imperative to develop new antibacterial strategies. In this context, this study aims to evaluate the antibacterial efficacy of Origanum vulgare essential oil (OEO), alone and in combination with antibiotics, as well as its antibiofilm action against multidrug-resistant Gram-negative strains. OEO components were identified by gas chromatography-mass spectrometry (GC-MS), and antibacterial activity was assessed using the agar diffusion test and the microdilution method. Interactions between OEO and antibiotics were examined using the checkerboard method, while antibiofilm activity was analyzed using the crystal violet assay. Chemical analysis revealed that carvacrol was the major compound in OEO (61.51%). This essential oil demonstrated activity against all the tested strains, with inhibition zone diameters (IZDs) reaching 32.3 ± 1.5 mm. The combination of OEO with different antibiotics produced synergistic and additive effects, leading to a reduction of up to 98.44% in minimum inhibitory concentrations (MICs). In addition, this essential oil demonstrated an ability to inhibit and even eradicate biofilm formation. These results suggest that OEO could be exploited in the development of new molecules, combining its metabolites with antibiotics. Full article
(This article belongs to the Special Issue Healthcare-Associated Infections and Antimicrobial Therapy)
Show Figures

Figure 1

Figure 1
<p>Checkerboard assays using OEO and antibiotics. Blue wells indicate growth inhibition. Pink wells indicate growth. H1 well is used for sterility control, and the 12th-column is for growth control. H2–H11 contains antibiotic alone, while G1–A1 contains the OEO alone. All the other wells contain combinations of antibiotic and OEO. (<b>A</b>) Synergistic combination between OEO and CIP against <span class="html-italic">A. baumannii</span> 14889 strain. (<b>B</b>) Additive effect observed between OEO and CTX against <span class="html-italic">E. coli</span> 45 strain.</p>
Full article ">Figure 2
<p>Percentages of inhibition of biofilm formation (blue) and eradication of preformed biofilm (orange) by OEO against tested bacterial strains.</p>
Full article ">
14 pages, 16172 KiB  
Article
Topical Application of Cha-Koji, Green Tea Leaves Fermented with Aspergillus Luchuensis var Kawachii Kitahara, Promotes Acute Cutaneous Wound Healing in Mice
by Yasuhiro Katahira, Jukito Sonoda, Miu Yamagishi, Eri Horio, Natsuki Yamaguchi, Hideaki Hasegawa, Izuru Mizoguchi and Takayuki Yoshimoto
Sci. Pharm. 2024, 92(3), 44; https://doi.org/10.3390/scipharm92030044 - 12 Aug 2024
Abstract
“Koji” is one of the most well-known probiotic microorganisms in Japan that contribute to the maintenance of human health. Although the beneficial effects of some probiotics on ulcer healing have been demonstrated, there have been no reports on the wound healing effects of [...] Read more.
“Koji” is one of the most well-known probiotic microorganisms in Japan that contribute to the maintenance of human health. Although the beneficial effects of some probiotics on ulcer healing have been demonstrated, there have been no reports on the wound healing effects of koji to date. In the present study, we investigated the effects of “cha-koji”, green tea leaves fermented with Aspergillus luchuensis, on cutaneous wound healing, using a linear incision wound mouse model. Topical application of autoclave-sterilized cha-koji suspension on the dorsal incision wound area healed the wound significantly faster and, notably, with less scarring than did the green tea or the control distilled water treatment. Further in vitro experiments revealed that the accelerated effects of cha-koji could be attributed to its increased anti-bacterial activity, enhanced epidermal cell proliferation and migration, augmented expression of the anti-inflammatory cytokine transforming growth factor-β1, reduced expression of inflammatory cytokine interleukin-6 in macrophages, and decreased endoplasmic reticulum stress. In addition, we conducted a skin sensitizing potential test, which revealed that cha-koji had no adverse effects that posed a sensitizing risk. Thus, cha-koji may have a potent therapeutic effect on cutaneous wound healing, opening up a new avenue for its clinical application as a medical aid. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Autoclave-sterilized cha-koji suspension promotes cutaneous wound healing. A cutaneous incision wound was made on the dorsal skin of the mice; coated with gauze soaked in green tea, cha-koji suspension, or distilled water; and covered with a transparent film dressing, Tegaderm, daily until day 13 (<b>A</b>,<b>B</b>). Photographs of the incision were taken over time. Representative photographs of the healing process are shown (<b>C</b>). The length of the incision wound in each photograph was measured using FIJI, and the relative length at each time point to the initial length of 10 mm was calculated (<b>D</b>). Data are shown as the mean ± SD (<span class="html-italic">n</span> = 3) and are representative of five independent experiments. <span class="html-italic">p</span>-values were determined via two-way analysis of variance with Tukey’s multiple comparison test. The red and blue asterisks marked on day 6, 8, 10, and 12 in the figure (<b>D</b>) mean that there are significant differences between cha-koji and green tea and between cha-koji and distilled water, respectively. * <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. Scale bar = 5 mm.</p>
Full article ">Figure 2
<p>Autoclave-sterilized cha-koji suspension showing anti-bacterial activity. <span class="html-italic">E. coli</span> was cultured in LB containing various concentrations of autoclave-sterilized cha-koji, green tea (0.5, 1.0, 2.0, 4.0, and 5.0%) or distilled water for 8 h at 37 °C (<b>A</b>). As a positive control, raw green tea was used. Bacterial growth was examined according to the turbidity of the LB medium, which was determined by measuring optical density at 600 nm (OD600) (<b>B</b>). Data are shown as the mean ± SD (<span class="html-italic">n</span> = 3) and are representative of three independent experiments. <span class="html-italic">p</span>-values were determined via one-way analysis of variance with Tukey’s multiple comparison test. *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 3
<p>Autoclave-sterilized cha-koji suspension promotes skin epidermal cell proliferation and migration. Keratinocyte PAM212 cells (<b>A</b>,<b>B</b>) or fibroblast NIH3T3 cells (<b>C</b>,<b>D</b>) were seeded in 24-well plates and cultured to semi-confluence in DMEM containing 10% FBS. The medium was then exchanged with DMEM containing 0.05% green tea or 0.05% cha-koji and 1.0% FBS. A cross-shaped scratch was made in the center of the well, and photographs were taken over time. Representative photographs are shown (<b>A</b>,<b>C</b>). The remaining wound area size was determined using FIJI, and the ratio relative to the initial wound area on day 0 was calculated (<b>B</b>,<b>D</b>). Data are shown as the mean ± SD (<span class="html-italic">n</span> = 6) and are representative of four (<b>A</b>,<b>B</b>) and three (<b>C</b>,<b>D</b>) independent experiments. <span class="html-italic">p</span>-values were determined via one-way analysis of variance with Tukey’s multiple comparison test. ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001. Scale bar = 500 µm.</p>
Full article ">Figure 4
<p>Autoclave-sterilized cha-koji suspension promotes the expression of the anti-inflammatory cytokine TGF-β1. Macrophage RAW264.7 cells were stimulated with 0.3% cha-koji, 0.3% green tea, and distilled water for 24 and 48 h. RT-qPCR was then performed to analyze the mRNA expression of <span class="html-italic">TGF-β1</span> (<b>A</b>,<b>B</b>) and <span class="html-italic">IL-6</span> (<b>C</b>,<b>D</b>). <span class="html-italic">HPRT</span> was used as an internal control, and the relative expression of <span class="html-italic">TGF-β1</span> or <span class="html-italic">IL-6</span> to <span class="html-italic">HPRT</span> was calculated. Data are shown as the mean ± SD (<span class="html-italic">n</span> = 4−6) and are representative of three (<b>A</b>,<b>B</b>) and five (<b>C</b>,<b>D</b>) independent experiments. <span class="html-italic">p</span>-values were determined via one-way analysis of variance with Dunnett’s multiple comparison test. * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 5
<p>Autoclave-sterilized cha-koji suspension suppresses ER stress. Fibroblast NIH3T3 cells were stimulated with tunicamycin (1.0 μg/mL) in the presence or absence of 0.25% cha-koji and 0.25% green tea for 12 h. RNA was extracted, and RT-qPCR analysis was performed to examine the mRNA expression of ER stress-related factors, namely, <span class="html-italic">HSPA5</span> (<b>A</b>), <span class="html-italic">XBP1</span> (<b>B</b>), and <span class="html-italic">CHOP</span> (<b>C</b>). <span class="html-italic">HPRT</span> was used as an internal control, and its relative expression of <span class="html-italic">HPRT</span> was calculated. Data are shown as the mean ± SD (<span class="html-italic">n</span> = 3) and are representative of two independent experiments. <span class="html-italic">p</span>-values were determined via one-way analysis of variance with Dunnett’s multiple comparison test. After 48 h, photographs of each well were taken (<b>D</b>), and cell growth activity was determined by measuring the remaining viable cell area relative to the tunicamycin-untreated, distilled water-treated cell area using FIJI (<b>E</b>). Data are shown as the mean ± SD (<span class="html-italic">n</span> = 3) and are representative of three independent experiments. <span class="html-italic">p</span>-values were determined via one-way analysis of variance with Dunnett’s multiple comparison test. * <span class="html-italic">p</span> &lt; 0.05 and *** <span class="html-italic">p</span> &lt; 0.001. Scale bar = 250 µm.</p>
Full article ">Figure 6
<p>Autoclave-sterilized cha-koji suspension poses no potential skin sensitizing risk. Human monocytic THP-1 cells were stimulated with cha-koji or green tea suspension (0.25, 1.0, and 4.0%) together with the positive control, conditioned medium from rhododenol-treated melanoma SK-MEL-37 cells. After 24 h, CD86 expression in the THP-1 cells was analyzed via flow cytometry using anti-CD86 (red-shaded histogram) or a control antibody (blue-shaded histogram). Conditioned medium from the rhododenol-treated melanoma (SK-MEL-37) was used as a positive control. Representative histograms for the cell surface expression of CD86 are shown (<b>A</b>). The RFI values of each sample were calculated and compared (<b>B</b>). An RFI higher than 150% is considered to be positive for skin sensitizing potential.</p>
Full article ">
20 pages, 5019 KiB  
Article
Enhancing Deer Sous Vide Meat Shelf Life and Safety with Eugenia caryophyllus Essential Oil against Salmonella enterica
by Miroslava Kačániová, Stefania Garzoli, Anis Ben Hsouna, Zhaojun Ban, Joel Horacio Elizondo-Luevano, Maciej Ireneusz Kluz, Rania Ben Saad, Peter Haščík, Natália Čmiková, Božena Waskiewicz-Robak, Ján Kollár and Alessandro Bianchi
Foods 2024, 13(16), 2512; https://doi.org/10.3390/foods13162512 - 12 Aug 2024
Abstract
Modern lifestyles have increased the focus on food stability and human health due to evolving industrial goals and scientific advancements. Pathogenic microorganisms significantly challenge food quality, with Salmonella enterica and other planktonic cells capable of forming biofilms that make them more resistant to [...] Read more.
Modern lifestyles have increased the focus on food stability and human health due to evolving industrial goals and scientific advancements. Pathogenic microorganisms significantly challenge food quality, with Salmonella enterica and other planktonic cells capable of forming biofilms that make them more resistant to broad-spectrum antibiotics. This research examined the chemical composition and antibacterial and antibiofilm properties of the essential oil from Eugenia caryophyllus (ECEO) derived from dried fruits. GC-MS analyses identified eugenol as the dominant component at 82.7%. Additionally, the study aimed to extend the shelf life of sous vide deer meat by applying a plant essential oil and inoculating it with S. enterica for seven days at 4 °C. The essential oil demonstrated strong antibacterial activity against S. enterica. The ECEO showed significant antibiofilm activity, as indicated by the MBIC crystal violet test results. Data from MALDI-TOF MS analysis revealed that the ECEO altered the protein profiles of bacteria on glass and stainless-steel surfaces. Furthermore, the ECEO was found to have a beneficial antibacterial effect on S. enterica. In vacuum-packed sous vide red deer meat samples, the anti-Salmonella activity of the ECEO was slightly higher than that of the control samples. These findings underscore the potential of the ECEO’s antibacterial and antibiofilm properties in food preservation and extending the shelf life of meat. Full article
Show Figures

Figure 1

Figure 1
<p>GC-MS chromatogram of the ECEO.</p>
Full article ">Figure 2
<p>Representative MALDI-TOF mass spectra of <span class="html-italic">S. enterica</span>: (<b>A</b>) 3rd day; (<b>B</b>) 5th day; (<b>C</b>) 7th day; (<b>D</b>) 9th day; (<b>E</b>) 12th day; (<b>F</b>) 14th day. SE = <span class="html-italic">S. enterica</span>; G = glass; S = stainless-steel; and PC = planktonic cells.</p>
Full article ">Figure 2 Cont.
<p>Representative MALDI-TOF mass spectra of <span class="html-italic">S. enterica</span>: (<b>A</b>) 3rd day; (<b>B</b>) 5th day; (<b>C</b>) 7th day; (<b>D</b>) 9th day; (<b>E</b>) 12th day; (<b>F</b>) 14th day. SE = <span class="html-italic">S. enterica</span>; G = glass; S = stainless-steel; and PC = planktonic cells.</p>
Full article ">Figure 2 Cont.
<p>Representative MALDI-TOF mass spectra of <span class="html-italic">S. enterica</span>: (<b>A</b>) 3rd day; (<b>B</b>) 5th day; (<b>C</b>) 7th day; (<b>D</b>) 9th day; (<b>E</b>) 12th day; (<b>F</b>) 14th day. SE = <span class="html-italic">S. enterica</span>; G = glass; S = stainless-steel; and PC = planktonic cells.</p>
Full article ">Figure 3
<p>Dendrogram of <span class="html-italic">S. enterica</span> generated using MSPs of the planktonic cells and the control. SE = <span class="html-italic">S. enterica</span>; C = glass; S = stainless-steel; and PC = planktonic cells.</p>
Full article ">Figure 4
<p>Total viable count (log CFU/g) of sous vide deer meat samples after 1 and 7 days of storage, treated in a water bath at temperatures between 50 and 65 °C for 5 to 20 min. Data are the mean (bars indicate ± SD) of 3 deer meat samples. Control: deer meat samples placed in polyethylene bags without vacuum. Control vacuum: deer meat samples vacuum-packed in polyethylene bags. Essential oil: deer meat samples treated with 1% ECEO and vacuum-packed. <span class="html-italic">Salmonella enterica</span>: deer meat samples inoculated with <span class="html-italic">S. enterica</span> and vacuum-packed. Essential oil + <span class="html-italic">Salmonella enterica:</span> deer meat samples treated with 1% ECEO and inoculated with <span class="html-italic">S. enterica</span> and vacuum-packed.</p>
Full article ">Figure 5
<p>Total coliform bacteria (log CFU/g) of sous vide deer meat samples after 1 and 7 days of storage, treated in a water bath at temperatures between 50 and 65 °C for 5 to 20 min. Data are the mean (bars indicate ± SD) of 3 deer meat samples. Control: deer meat samples placed in polyethylene bags without vacuum. Control vacuum: deer meat samples vacuum-packed in polyethylene bags. Essential oil: deer meat samples treated with 1% ECEO and vacuum-packed. <span class="html-italic">Salmonella enterica</span>: deer meat samples inoculated with <span class="html-italic">S. enterica</span> and vacuum-packed. Essential oil + <span class="html-italic">Salmonella enterica:</span> deer meat samples treated with 1% ECEO and inoculated with <span class="html-italic">S. enterica</span> and vacuum-packed.</p>
Full article ">Figure 6
<p><span class="html-italic">Salmonella enterica</span> count (log CFU/g) of sous vide deer meat samples after 1 and 7 days of storage, treated in a water bath at temperatures between 50 and 65 °C for 5 to 20 min. Data are the mean (bars indicate ± SD) of 3 deer meat samples. Control: deer meat samples placed in polyethylene bags without vacuum. Control vacuum: deer meat samples vacuum-packed in polyethylene bags. Essential oil: deer meat samples treated with 1% ECEO and vacuum-packed. <span class="html-italic">Salmonella enterica</span>: deer meat samples inoculated with <span class="html-italic">S. enterica</span> and vacuum-packed. Essential oil + <span class="html-italic">Salmonella enterica:</span> deer meat samples treated with 1% ECEO and inoculated with <span class="html-italic">S. enterica</span> and vacuum-packed.</p>
Full article ">Figure 7
<p>Krona chart: Isolated species, genera, and families from deer sous vide meat at 1 day.</p>
Full article ">Figure 8
<p>Krona chart: Isolated species, genera, and families from deer sous vide meat after 7 days.</p>
Full article ">
16 pages, 1908 KiB  
Article
Synthesis of Second-Generation Analogs of Temporin-SHa Peptide Having Broad-Spectrum Antibacterial and Anticancer Effects
by Arif Iftikhar Khan, Shahzad Nazir, Muhammad Nadeem ul Haque, Rukesh Maharjan, Farooq-Ahmad Khan, Hamza Olleik, Elise Courvoisier-Dezord, Marc Maresca and Farzana Shaheen
Antibiotics 2024, 13(8), 758; https://doi.org/10.3390/antibiotics13080758 (registering DOI) - 11 Aug 2024
Viewed by 317
Abstract
Antimicrobial peptides (AMPs) are a promising class of therapeutic alternatives with broad-spectrum activity against resistant pathogens. Small AMPs like temporin-SHa (1) and its first-generation analog [G10a]-SHa (2) possess notable efficacy against Gram-positive and Gram-negative bacteria. In an effort to [...] Read more.
Antimicrobial peptides (AMPs) are a promising class of therapeutic alternatives with broad-spectrum activity against resistant pathogens. Small AMPs like temporin-SHa (1) and its first-generation analog [G10a]-SHa (2) possess notable efficacy against Gram-positive and Gram-negative bacteria. In an effort to further improve this antimicrobial activity, second-generation analogs of 1 were synthesised by replacing the natural glycine residue at position-10 of the parent molecule with atypical amino acids, such as D-Phenylalanine, D-Tyrosine and (2-Naphthyl)-D-alanine, to study the effect of hydrophobicity on antimicrobial efficacy. The resultant analogs (36) emerged as broad-spectrum antibacterial agents. Notably, the [G10K]-SHa analog (4), having a lysine substitution, demonstrated a 4-fold increase in activity against Gram-negative (Enterobacter cloacae DSM 30054) and Gram-positive (Enterococcus faecalis DSM 2570) bacteria relative to the parent peptide (1). Among all analogs, [G10f]-SHa peptide (3), featuring a D-Phe substitution, showed the most potent anticancer activity against lung cancer (A549), skin cancer (MNT-1), prostate cancer (PC-3), pancreatic cancer (MiaPaCa-2) and breast cancer (MCF-7) cells, achieving an IC50 value in the range of 3.6–6.8 µM; however, it was also found to be cytotoxic against normal cell lines as compared to [G10K]-SHa (4). Peptide 4 also possessed good anticancer activity but was found to be less cytotoxic against normal cell lines as compared to 1 and 3. These findings underscore the potential of second-generation temporin-SHa analogs, especially analog 4, as promising leads to develop new broad-spectrum antibacterial and anticancer agents. Full article
(This article belongs to the Section Antimicrobial Peptides)
Show Figures

Figure 1

Figure 1
<p>UPLC profiles of the synthesised peptides; (<b>A</b>) temporin-SHa; (<b>B</b>) [G10a]-SHa; (<b>C</b>) [G10f]-SHa; (<b>D</b>) [G10y]-SHa; (<b>E</b>) [G10n]-SHa and (<b>F</b>) [G10K]-SHa.</p>
Full article ">Figure 2
<p>Circular dichroism of temporin SHa, [G10a]-Sha, and newly synthesised second-generation analogs of [G10a]-SHa in 20 mM SDS.</p>
Full article ">Figure 3
<p>Antiproliferative effect of SHa derivatives on human cancer cells. The antiproliferative effect of SHa derivatives was measured on dividing cancer cells, as explained in <a href="#sec4-antibiotics-13-00758" class="html-sec">Section 4</a> (Temporin-SHa (<b>1</b>): open black circles, [G10a]-SHa (<b>2</b>): closed red circles, [G10f]-SHa (<b>2</b>): closed green squares, [G10K]-SHa (<b>3</b>): closed black diamonds, [G10n]-SHa (<b>4</b>): inverted open purple triangles, [G10y]-SHa (<b>5</b>): closed blue triangles). Results are expressed as a percentage of cell proliferation, the untreated cells giving 100% proliferation (means ± SD, <span class="html-italic">n</span> = 3).</p>
Full article ">Figure 4
<p>Antiproliferative effect of SHa derivatives on human normal/non cancerous cells. The antiproliferative effect of SHa derivatives was measured on dividing normal cells, as explained in <a href="#sec4-antibiotics-13-00758" class="html-sec">Section 4</a> (temporin-SHa (<b>1</b>): open black circles, [G10a]-SHa (<b>2</b>): closed red circles, [G10f]-SHa (<b>2</b>): closed green squares, [G10K]-SHa (<b>3</b>): closed black diamonds, [G10n]-SHa (<b>4</b>): inverted open purple triangles, [G10y]-SHa (<b>5</b>): closed blue triangles). Results are expressed as a percentage of cell proliferation, the untreated cells giving 100% proliferation (means ± SD, <span class="html-italic">n</span> = 3).</p>
Full article ">Figure 5
<p>Cytotoxic effect of SHa derivatives on human normal and cancer lung cells. The cytotoxic effect of temporin-SHa (<b>1</b>) derivatives was measured on confluent/non-dividing cells, as explained in <a href="#sec4-antibiotics-13-00758" class="html-sec">Section 4</a>, using human lung cancer (A549 cells) and normal cells (BEAS-2B cells); ([G10a]-SHa (<b>2</b>): closed red circles, [G10f]-SHa (<b>2</b>): closed green squares, [G10K]-SHa (<b>3</b>): closed black diamonds, [G10n]-SHa (<b>4</b>): inverted open purple triangles, [G10y]-SHa (<b>5</b>): closed blue triangles). Results are expressed as a percentage of cell proliferation, the untreated cells giving 100% proliferation (means ± SD, <span class="html-italic">n</span> = 3).</p>
Full article ">Scheme 1
<p>Synthesis and structure of temporin-SHa (<b>1</b>), its first-generation [G10a]-SHa peptide (<b>2</b>) and newly synthesised second-generation analogs (<b>3</b>–<b>6</b>).</p>
Full article ">
26 pages, 4780 KiB  
Article
Synthesis, Structural Properties and Biological Activities of Novel Hydrazones of 2-, 3-, 4-Iodobenzoic Acid
by Izabela Czyżewska, Liliana Mazur, Anna Biernasiuk, Anna Hordyjewska and Łukasz Popiołek
Molecules 2024, 29(16), 3814; https://doi.org/10.3390/molecules29163814 - 11 Aug 2024
Viewed by 389
Abstract
Nowadays, searching for novel antimicrobial agents is crucial due to the increasing number of resistant bacterial strains. Moreover, cancer therapy is a major challenge for modern medicine. Currently used cytostatics have a large number of side effects and insufficient therapeutic effects. Due to [...] Read more.
Nowadays, searching for novel antimicrobial agents is crucial due to the increasing number of resistant bacterial strains. Moreover, cancer therapy is a major challenge for modern medicine. Currently used cytostatics have a large number of side effects and insufficient therapeutic effects. Due to the above-mentioned facts, we undertook research to synthesize novel compounds from the acylhydrazone group aimed at obtaining potential antimicrobial and anticancer agents. As a starting material, we employed hydrazides of 2-, 3- or 4-iodobenzoic acid, which gave three series of acylhydrazones in the condensation reaction with various aldehydes. The chemical structure of all obtained compounds was confirmed by IR, 1H NMR, and 13C NMR. The structure of selected compounds was determined by single-crystal X-ray diffraction analysis. Additionally, all samples were characterized using powder X-ray diffraction. The other issue in this research was to examine the possibility of the solvent-free synthesis of compounds using mechanochemical methods. The biological screening results revealed that some of the newly synthesized compounds indicated a beneficial antimicrobial effect even against MRSA—the methicillin-resistant Staphylococcus aureus ATCC 43300 strain. In many cases, the antibacterial activity of synthesized acylhydrazones was equal to or better than that of commercially available antibacterial agents that were used as reference substances in this research. Significantly, the tested compounds do not show toxicity to normal cell lines either. Full article
Show Figures

Figure 1

Figure 1
<p>PXRD patterns of compound <b>13</b>: (<b>a</b>) simulated from the SCXRD data; (<b>b</b>) experimental after synthesis from solution; (<b>c</b>–<b>e</b>) experimental after liquid-assisted grinding (LAG) for 30, 60 and 90 min, respectively, using ethanol as a solvent; (<b>f</b>) experimental after LAG for 90 min using acetonitrile.</p>
Full article ">Figure 2
<p>Perspective view of the molecules constituting the asymmetric part in crystals <b>13</b>, <b>13</b>∙<b>ACN</b>, <b>20</b> and <b>26a</b> with the atom-numbering scheme. Thermal ellipsoids are drawn at the 50% probability level. Dashed lines indicate the hydrogen bonds.</p>
Full article ">Figure 3
<p>Molecular overlay of the conformers found in: (<b>a</b>) polymorphic modifications <b>26a</b> (red line), <b>26b</b> (green line) and <b>26c</b> (dark blue line); (<b>b</b>) unsolvated crystal <b>13</b> (molecule 13A—blue line, molecule 13B—green line) and its solvate <b>13</b>∙<b>ACN</b> (pink line).</p>
Full article ">Figure 4
<p>Part of the crystal structure of <b>13</b>·<b>ACN</b> in view along the <span class="html-italic">a</span> axis, showing the formation of channels filled in by the solvent molecules.</p>
Full article ">Figure 5
<p>Part of the crystal structure of <b>9</b> showing (<b>a</b>) supramolecular chains stabilized via strong N1–H1n∙∙∙O1A/N1A–H1nA∙∙∙O1 (x, y + 1, <span class="html-italic">z</span>) hydrogen bonds and weak C–H∙∙∙O/π interactions; (<b>b</b>) crystal packing viewed along the <span class="html-italic">b</span> axis with marked 2D layer parallel to the (−102) crystallographic plane. Molecules <b>9-A</b> and <b>9-B</b> are marked in green and blue, respectively. Dashed lines indicate hydrogen bonds.</p>
Full article ">Figure 6
<p>(<b>a</b>) Part of the crystal structure of <b>26a</b> showing hydrogen-bonding motifs; (<b>b</b>) crystal packing in <b>26a</b> viewed along the <span class="html-italic">b</span> axis; (<b>c</b>) crystal packing in <b>26b</b> viewed down the <span class="html-italic">a</span> axis; (<b>d</b>) hydrogen-bonding patterns in crystal <b>26c</b>, (<b>e</b>) part of the crystal structure of <b>26c</b> in view along the <span class="html-italic">a</span> axis. Dashed lines indicate inter- and intramolecular interactions.</p>
Full article ">Scheme 1
<p>Synthesis of the hydrazides of 2-, 3- or 4-iodobenzoic acid.</p>
Full article ">Scheme 2
<p>Synthesis of the acylhydrazones of 2-, 3- or 4-iodobenzoic acid.</p>
Full article ">
14 pages, 2319 KiB  
Article
Improved Anti-Oxidant and Anti-Bacterial Capacities of Skim Milk Fermented by Lactobacillus plantarum
by Ying Wang, Bingtian Zhao, Yun Ding, Nan Liu, Cheng Yang and Yajuan Sun
Molecules 2024, 29(16), 3800; https://doi.org/10.3390/molecules29163800 - 10 Aug 2024
Viewed by 257
Abstract
Milk, on account of its abundant protein content, is recognized as a vital source of bioactive substances. In this study, the bioactive ingredients in milk were obtained by a combination of protease hydrolysis and fermentation with Lactobacillus plantarum. The compositions of protease [...] Read more.
Milk, on account of its abundant protein content, is recognized as a vital source of bioactive substances. In this study, the bioactive ingredients in milk were obtained by a combination of protease hydrolysis and fermentation with Lactobacillus plantarum. The compositions of protease hydrolysate (PM) and fermentation supernatant (FM) were determined, and their anti-oxidant and anti-bacterial activities were evaluated. Using LC-MS/MS, the molecular weights and sequences of the peptides were characterized, among which a total of 25 bioactive peptides were identified. The DPPH radical scavenging results demonstrated that FM exhibited an enhanced anti-oxidant capacity compared to PM. The bacterial survival rate results revealed that FM had a remarkable anti-bacterial ability compared to PM. Additionally, the anti-bacterial component and potential anti-bacterial mechanisms were determined. The results of cytoplasmic membrane depolarization, cell membrane permeability, and morphological observation indicated that FM could interact with bacterial membranes to achieve its anti-bacterial effect. These findings suggested that FM, as a bioactive substance of natural origin, holds potential applications in the functional food, pharmaceutical, and cosmetic industries. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) The molecular weight distributions and (<b>B</b>) Venn diagram of PM and FM.</p>
Full article ">Figure 2
<p>The DPPH scavenging ability of PM and FM. * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 indicate that the DPPH scavenging capacity of FM was significantly and extremely significantly higher than that of PM, respectively.</p>
Full article ">Figure 3
<p>The bacterial survival rate of <span class="html-italic">Staphylococcus aureus</span> (<b>A</b>,<b>B</b>) and <span class="html-italic">Escherichia coli</span> (<b>C</b>,<b>D</b>) after 24 h of incubation with PM or FM at the final concentrations of 50, 100, and 200 mg/mL (<b>A</b>,<b>C</b>), with 100 and 200 mg/mL of FM or corresponding concentrations of organic acid (<b>B</b>,<b>D</b>). Different letters (a–e) indicate significant differences.</p>
Full article ">Figure 4
<p>The cytoplasmic membrane potential of <span class="html-italic">Staphylococcus aureus</span> (<b>A</b>) and <span class="html-italic">Escherichia coli</span> (<b>B</b>) treated with samples. (*) indicates sample addition.</p>
Full article ">Figure 5
<p>Live–dead fluorescence microscope images from <span class="html-italic">Staphylococcus aureus</span> (<b>A</b>) and <span class="html-italic">Escherichia coli</span> (<b>B</b>) with 200 mg/mL of PM and FM. The ratio of red fluorescence intensity (dead bacteria) to green fluorescence intensity (live bacteria) of <span class="html-italic">Staphylococcus aureus</span> (<b>C</b>) and <span class="html-italic">Escherichia coli</span> (<b>D</b>). The different superscript letters (a–c) indicate that the results possessed significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 6
<p>SEM images of <span class="html-italic">Staphylococcus aureus</span> (<b>A</b>) and <span class="html-italic">Escherichia coli</span> (<b>B</b>) in the absence and presence of PM or FM.</p>
Full article ">
14 pages, 3333 KiB  
Article
Discovery of Antibacterial Compounds with Potential Multi-Pharmacology against Staphylococcus Mur ligase Family Members by In Silico Structure-Based Drug Screening
by Mio Teshima, Kohei Monobe, Saya Okubo and Shunsuke Aoki
Molecules 2024, 29(16), 3792; https://doi.org/10.3390/molecules29163792 - 10 Aug 2024
Viewed by 279
Abstract
Staphylococcus aureus (S. aureus) is a major bacterial infection in humans, leading to severe disease and causing death. The stagnation of antibiotic development in recent decades has made it difficult to combat drug-resistant infections. In this study, we performed an in [...] Read more.
Staphylococcus aureus (S. aureus) is a major bacterial infection in humans, leading to severe disease and causing death. The stagnation of antibiotic development in recent decades has made it difficult to combat drug-resistant infections. In this study, we performed an in silico structure-based drug screening (SBDS) targeting the S. aureus MurE (saMurE) enzyme involved in cell wall synthesis of S. aureus. saMurE is an enzyme that is essential for the survival of S. aureus but not present in humans. SBDS identified nine saMurE inhibitor candidates, Compounds 19, from a structural library of 154,118 compounds. Among them, Compound 2 showed strong antibacterial activity against Staphylococcus epidermidis (S. epidermidis) used as a model bacterium. Amino acid sequence homology between saMurE and S. epidermidis MurE is 87.4%, suggesting that Compound 2 has a similar inhibitory effect on S. aureus. Compound 2 showed an IC50 value of 301 nM for S. epidermidis in the dose-dependent growth inhibition assay. Molecular dynamics simulation showed that Compound 2 binds stably to both S. aureus MurD and S. aureus MurF, suggesting that it is a potential multi-pharmacological pharmacological inhibitor. The structural and bioactivity information of Compound 2, as well as its potential multiple-target activity, could contribute to developing new antimicrobial agents based on MurE inhibition. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>saMurE inhibitor screening strategy.</p>
Full article ">Figure 2
<p>Growth inhibitory effect of Compounds <b>1</b>–<b>9</b> on bacteria (<span class="html-italic">S. epidermidis</span>). A total of 0.3% DMSO and 100 μM ampicillin (AMP) were used as samples for comparison. Compounds <b>1</b>–<b>9</b> (100 μM). The vertical axis is the mean +/− SEM of the results of four independent experiments. Dunnett’s test: ****; <span class="html-italic">p</span> &lt; 0.0001; **; <span class="html-italic">p</span> &lt; 0.0021; *; <span class="html-italic">p</span> &lt; 0.0332; n.s. = not significant.</p>
Full article ">Figure 3
<p>Determination of 50% growth inhibition concentration against bacteria (<span class="html-italic">S. epidermidis</span>). The vertical axis shows the relative bacterial growth rate. The horizontal axis shows the molar concentration of Compound <b>2</b>.</p>
Full article ">Figure 4
<p>MDS results for the saMurE–Compound <b>2</b> complex: (<b>A</b>) Transition of ligand RMSD value (nm). Ligand RMSD values were calculated by comparison with the post-equilibration pose. (<b>B</b>) Radius (nm) of gyration during MDS. (<b>C</b>) Number of intermolecular hydrogen bonds.</p>
Full article ">Figure 5
<p>ProLIF analysis for the saMurE–Compound <b>2</b> interaction: (<b>A</b>) Interacting residues throughout the MDS timeframe. In the screening process, the piperazine group of Compound <b>2</b> is protonated. (<b>B</b>) A major (≥60% probability of presence) interaction residue group was observed throughout the entire period.</p>
Full article ">Figure 6
<p>Toxicity verification of Compound <b>2</b> on mammalian-derived cells: (<b>A</b>) COS-7 cells; (<b>B</b>) HepG2 cells. Negative control was 0.3% DMSO and positive control was 50 μM triclosan (TCS). Concentration of Compound <b>2</b> was 100 μM. Dunnett’s test: ***; <span class="html-italic">p</span> &lt; 0.0002; n.s. = not significant.</p>
Full article ">Figure 7
<p>MDS results for saMur ligase–Compound <b>2</b> complexes: (<b>A</b>) ligand RMSD values (nm); (<b>B</b>) radius (nm) of gyration during MDS; (<b>C</b>) number of intermolecular hydrogen bonds. saMurC (black), saMurD (red), saMurF (green).</p>
Full article ">
21 pages, 4169 KiB  
Article
Kalanchoe tomentosa: Phytochemical Profiling, and Evaluation of Its Biological Activities In Vitro, In Vivo, and In Silico
by Jorge L. Mejía-Méndez, Gildardo Sánchez-Ante, Yulianna Minutti-Calva, Karen Schürenkämper-Carrillo, Diego E. Navarro-López, Ricardo E. Buendía-Corona, Ma. del Carmen Ángeles González-Chávez, Angélica Lizeth Sánchez-López, J. Daniel Lozada-Ramírez, Eugenio Sánchez-Arreola and Edgar R. López-Mena
Pharmaceuticals 2024, 17(8), 1051; https://doi.org/10.3390/ph17081051 - 9 Aug 2024
Viewed by 440
Abstract
In this work, the leaves of K. tomentosa were macerated with hexane, chloroform, and methanol, respectively. The phytochemical profiles of hexane and chloroform extracts were unveiled using GC/MS, whereas the chemical composition of the methanol extract was analyzed using UPLC/MS/MS. The antibacterial activity [...] Read more.
In this work, the leaves of K. tomentosa were macerated with hexane, chloroform, and methanol, respectively. The phytochemical profiles of hexane and chloroform extracts were unveiled using GC/MS, whereas the chemical composition of the methanol extract was analyzed using UPLC/MS/MS. The antibacterial activity of extracts was determined against gram-positive and gram-negative strains through the minimal inhibitory concentration assay, and in silico studies were implemented to analyze the interaction of phytoconstituents with bacterial peptides. The antioxidant property of extracts was assessed by evaluating their capacity to scavenge DPPH, ABTS, and H2O2 radicals. The toxicity of the extracts was recorded against Artemia salina nauplii and Caenorhabditis elegans nematodes. Results demonstrate that the hexane and chloroform extracts contain phytosterols, triterpenes, and fatty acids, whereas the methanol extract possesses glycosidic derivatives of quercetin and kaempferol together with sesquiterpene lactones. The antibacterial performance of extracts against the cultured strains was appraised as weak due to their MIC90 values (>500 μg/mL). As antioxidants, treatment with extracts executed high and moderate antioxidant activities within the range of 50–300 μg/mL. Extracts did not decrease the viability of A. salina, but they exerted a high toxic effect against C. elegans during exposure to treatment. Through in silico modeling, it was recorded that the flavonoids contained in the methanol extract can hamper the interaction of the NAM/NAG peptide, which is of great interest since it determines the formation of the peptide wall of gram-positive bacteria. This study reports for the first time the biological activities and phytochemical content of extracts from K. tomentosa and proposes a possible antibacterial mechanism of glycosidic derivatives of flavonoids against gram-positive bacteria. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) FTIR spectra of the hexane (H), chloroform (Cl), and methanol (M) extracts, and (<b>B</b>) quercetin calibration curve to estimate TFC.</p>
Full article ">Figure 2
<p>Mass spectra and chemical structures of identified compounds in the methanol extract from <span class="html-italic">K. tomentosa</span>: not identified (<b>1</b>), kaempferol 3-<span class="html-italic">O</span>-rutinoside (<b>2</b>), quercetin-<span class="html-italic">O</span>-hexoside (<b>3</b>), kaempferide-3-glucuronide (<b>4</b>), eriodictyol-7-<span class="html-italic">O</span>-hexoside (<b>5</b>), deacetoxy (7)-7-oxokhivorinic acid (<b>6</b>), kaempferol 3-<span class="html-italic">O</span>-α-L-arabinopyranosyl-(1→2) α-L-rhamnopyranoside (<b>7</b>), kaempferin (<b>8</b>), ganolucidic acid C (<b>9</b>), kaempferol (<b>10</b>), spiraeoside (<b>11</b>), apigenin (<b>12</b>), linoleic acid (<b>13</b>), and heliannuol A (<b>14</b>).</p>
Full article ">Figure 2 Cont.
<p>Mass spectra and chemical structures of identified compounds in the methanol extract from <span class="html-italic">K. tomentosa</span>: not identified (<b>1</b>), kaempferol 3-<span class="html-italic">O</span>-rutinoside (<b>2</b>), quercetin-<span class="html-italic">O</span>-hexoside (<b>3</b>), kaempferide-3-glucuronide (<b>4</b>), eriodictyol-7-<span class="html-italic">O</span>-hexoside (<b>5</b>), deacetoxy (7)-7-oxokhivorinic acid (<b>6</b>), kaempferol 3-<span class="html-italic">O</span>-α-L-arabinopyranosyl-(1→2) α-L-rhamnopyranoside (<b>7</b>), kaempferin (<b>8</b>), ganolucidic acid C (<b>9</b>), kaempferol (<b>10</b>), spiraeoside (<b>11</b>), apigenin (<b>12</b>), linoleic acid (<b>13</b>), and heliannuol A (<b>14</b>).</p>
Full article ">Figure 3
<p>Antibacterial activity of extracts from <span class="html-italic">K. tomentosa</span> against (<b>A</b>) <span class="html-italic">E. coli</span>, (<b>B</b>) <span class="html-italic">K. pneumoniae</span>, and (<b>C</b>) <span class="html-italic">S. aureus</span>. Positive controls comprehended fosfomycin (F), amikacin (AK), and vancomycin (V), respectively. The mean ± SD of three independent experiments is shown.</p>
Full article ">Figure 4
<p>Docking analysis and hydrogen bonding of compounds <b>2</b>, <b>7</b>, <b>9</b>, and <b>11</b> with the NAM/NAG-peptide subunits of the <span class="html-italic">S. aureus</span> cell wall. Green cylinders represent hydrogen bonds, whereas other colors show elements: oxygen (red), nitrogen (blue), and carbon (ligand).</p>
Full article ">Figure 5
<p>DPPH (<b>A</b>), ABTS (<b>B</b>), and H<sub>2</sub>O<sub>2</sub> (<b>C</b>) scavenging activity of extracts from <span class="html-italic">K. tomentosa</span>. Ascorbic acid (AC) was utilized as a positive control. The mean ± SD of three independent experiments is shown. Different letters indicate significant differences among extracts, which were observed at <span class="html-italic">p</span> values significantly below &lt;0.001 evaluated by a one-way ANOVA, followed by a post hoc multiple comparison with Tukey’s test.</p>
Full article ">Figure 6
<p>Toxicity of extracts from <span class="html-italic">K. tomentosa</span> against (<b>A</b>) <span class="html-italic">A. salina</span> nauplii and (<b>B</b>–<b>D</b>) <span class="html-italic">C. elegans</span>. Results presented for <span class="html-italic">A. salina</span>, and <span class="html-italic">C. elegans</span> were obtained after 24 and 6 h of exposure to treatment, respectively.</p>
Full article ">
23 pages, 5705 KiB  
Review
Recent Advances in Alkaloids from Papaveraceae in China: Structural Characteristics and Pharmacological Effects
by Meixian Zhang, Jing Yang, Yanping Sun and Haixue Kuang
Molecules 2024, 29(16), 3778; https://doi.org/10.3390/molecules29163778 - 9 Aug 2024
Viewed by 326
Abstract
The Papaveraceae plant family serves as a botanical reservoir for a variety of medicinal compounds that have been traditionally utilized in Chinese medicine for numerous generations. Growing attention towards the pharmaceutical potential of Papaveraceae has resulted in the identification of many alkaloids, which [...] Read more.
The Papaveraceae plant family serves as a botanical reservoir for a variety of medicinal compounds that have been traditionally utilized in Chinese medicine for numerous generations. Growing attention towards the pharmaceutical potential of Papaveraceae has resulted in the identification of many alkaloids, which have attracted significant attention from the scientific community because of their structural complexity and wide range of biological activities, such as analgesic, antihypertensive, antiarrhythmic, anti-inflammatory, antibacterial, anti-tumor, anti-cancer, and other activities, making them potential candidates for medical use. The primary objective of this review is to analyze the existing literature on the historical use of Papaveraceae plants, focusing on their alkaloid structures and relationship with pharmacological effects, as well as provide a theoretical basis for their clinical application, with the goal of unveiling the future potential of Papaveraceae plants. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The number of various alkaloids in the <span class="html-italic">papaver</span> family. I: benzophenanthridine (including dihydrobenzo[c]phenanthridines—113, hexahydrobenzo[c]phenanthridine—40, quaternary benzophenanthridine—13, and dimeric dihydrobenzophenanthridines—27); II: protoberberine (including berberines—52 and protoberberines—54); III: aporphine (including aporphines—53, proaporphines—6, isoproaporphines—2, and aporphine dimers—1); IV: benzylisoquinoline (including 1-benzylisoquinolines—36 and spirobenzylisoquinolines—11); V: benphthaleoquinoline (including phthaloisoquinolines—29, cleavage-ring phthaloisoquinolines—7, and dimeric cleavage-ring phthaloisoquinolines—1); VI: simple isoquinoline (20); VII: protopine (9); VIII: morphine (13); and IX: others (62).</p>
Full article ">Figure 2
<p>The known alkaloids in different genera.</p>
Full article ">Figure 3
<p>Benzophenanthridine alkaloids’ parent nucleus and representative compounds. I parent nucleus of benzophenanthridine alkaloids; chelidonine and corynoline are type of I-i; sanguinarine and chelerythrineand are type of I-ii; dihydrosanguinarine and demethylchelerythrine are type of I-iii; and chelidimerine and sanguidimerine are type of I-iv.</p>
Full article ">Figure 4
<p>Protoberberine alkaloids’ parent nucleus and the representative compounds. II parent nucleus of protoberberine alkaloids; berberine, coptisine, columbamine, berberrubine, dihydrocoptisine, rupestrines B, and rupestrines C are type of II-i; and corydaline, <span class="html-small-caps">d</span>,<span class="html-small-caps">l</span>-tetrahydropalmatine, yuanhunine, 8-oxocoptisine, 1-nitroapocavidine, 4-nitroisoapocavidine are type of II-ii.</p>
Full article ">Figure 5
<p>Aporphine alkaloids’ parent nucleus and the representative compounds. III parent nucleus of aporphine alkaloids; corydine, isocorydine, glaucine, magnoflorine, menisperine, corunine are type of III-i; (-)-pronuciferine, mecambrine, glaziovine are type of III-ii; isocorydione and demethylsonodione are type of III-iii; and dactylidine are type of III-iv.</p>
Full article ">Figure 6
<p>Benzylisoquinoline alkaloids’ parent nucleus and the representative compounds. IV parent nucleus of benzylisoquinoline alkaloids; papaverine, codamine, armepavine, dl-coclaurine, 7-isoquinoline, hendersine B are type of IV-i; and fumaricine, (13<span class="html-italic">S</span>,14<span class="html-italic">S</span>)-tomentelline E, hendersine D, hendersine E are type of IV-ii.</p>
Full article ">Figure 7
<p>Benphthaleoquinoline alkaloids’ parent nucleus and the representative compounds. V parent nucleus of benphthaleoquinoline alkaloids; corlumidine, mucroniferanine B, mucroniferanine C, torulosine, <span class="html-italic">N</span>-methyldemethyltorulosine are type of V-i; bicucullinine, leonticine, <span class="html-italic">N</span>-methylhydrasteine are type of V-ii, dactylicapnosine is type of V-iii.</p>
Full article ">Figure 8
<p>Simple-isoquinoline alkaloids’ parent nucleus and the representative compounds. VI parent nucleus of simple-isoquinoline alkaloids.</p>
Full article ">Figure 9
<p>Protopine alkaloids’ parent nucleus and the compounds. VII parent nucleus of protopine alkaloids.</p>
Full article ">Figure 10
<p>Morphine alkaloids’ parent nucleus and the compounds. VIII parent nucleus of morphine alkaloids.</p>
Full article ">Figure 11
<p>Other alkaloids and their representative compounds.</p>
Full article ">Figure 12
<p>The pharmacological effects of <span class="html-italic">Papaveraceae</span>.</p>
Full article ">
18 pages, 8359 KiB  
Article
Membrane Damage and Metabolic Disruption as the Mechanisms of Linalool against Pseudomonas fragi: An Amino Acid Metabolomics Study
by Jiaxin Cai, Haiming Chen, Runqiu Wang, Qiuping Zhong, Weijun Chen, Ming Zhang, Rongrong He and Wenxue Chen
Foods 2024, 13(16), 2501; https://doi.org/10.3390/foods13162501 - 9 Aug 2024
Viewed by 471
Abstract
Pseudomonas fragi (P. fragi) is usually detected in low-temperature meat products, and seriously threatens food safety and human health. Therefore, the study investigated the antibacterial mechanism of linalool against P. fragi from membrane damage and metabolic disruption. Results from field-emission transmission [...] Read more.
Pseudomonas fragi (P. fragi) is usually detected in low-temperature meat products, and seriously threatens food safety and human health. Therefore, the study investigated the antibacterial mechanism of linalool against P. fragi from membrane damage and metabolic disruption. Results from field-emission transmission electron microscopy (FETEM) and atomic force microscopy (AFM) showed that linalool damage membrane integrity increases surface shrinkage and roughness. According to Fourier transform infrared (FTIR) spectra results, the components in the membrane underwent significant changes, including nucleic acid leakage, carbohydrate production, protein denaturation and modification, and fatty acid content reduction. The data obtained from amino acid metabolomics indicated that linalool caused excessive synthesis and metabolism of specific amino acids, particularly tryptophan metabolism and arginine biosynthesis. The reduced activities of glucose 6-phosphate dehydrogenase (G6PDH), malate dehydrogenase (MDH), and phosphofructokinase (PFK) suggested that linalool impair the respiratory chain and energy metabolism. Meanwhile, genes encoding the above enzymes were differentially expressed, with pfkB overexpression and zwf and mqo downregulation. Furthermore, molecular docking revealed that linalool can interact with the amino acid residues of G6DPH, MDH and PFK through hydrogen bonds. Therefore, it is hypothesized that the mechanism of linalool against P. fragi may involve cell membrane damage (structure and morphology), disturbance of energy metabolism (TCA cycle, EMP and HMP pathway) and amino acid metabolism (cysteine, glutamic acid and citrulline). These findings contribute to the development of linalool as a promising antibacterial agent in response to the food security challenge. Full article
Show Figures

Figure 1

Figure 1
<p>AFM images and FETEM micrographs of <span class="html-italic">P. fragi</span> treated with or without linalool for 4 h. MIC: cells treated with 1.5 mL/L of linalool. Ethanol: cells treated with ethanol. Control: cells treated with water.</p>
Full article ">Figure 2
<p>Calcein-AM/PI dual-stained confocal laser scanning microscopy (<b>A</b>). Representative FTIR spectra (4000–400 cm<sup>−1</sup>) of <span class="html-italic">P. fragi</span> (<b>B</b>). MIC: cells treated with 1.5 mL/L of linalool. Ethanol: cells treated with ethanol. Control: cells treated with water.</p>
Full article ">Figure 3
<p>SDS-PAGE analysis of intracellular soluble proteins of <span class="html-italic">P. fragi</span> (<b>A</b>). DNA release results based on gel electrophoresis in <span class="html-italic">P. fragi</span> (<b>B</b>). MIC: cells treated with 1.5 mL/L of linalool. Ethanol: cells treated with ethanol. Control: cells treated with water.</p>
Full article ">Figure 4
<p>Inhibitory effect of linalool on the enzymatic activity: G6DPH activity (<b>A</b>), MDH activity (<b>B</b>), PFK activity (<b>C</b>). MIC: cells treated with 1.5 mL/L of linalool. Ethanol: cells treated with ethanol. Control: cells treated with water.</p>
Full article ">Figure 5
<p>Score plots of PCA (<b>A</b>) and OPLS-DA (<b>B</b>). Volcano plot (<b>C</b>). Upregulated, downregulated and non-significant differential metabolites are represented by red, blue, grey dots, respectively. Heat map (<b>D</b>).</p>
Full article ">Figure 6
<p>Bubble plot (<b>A</b>); each bubble represents a metabolic pathway. KEGG pathway classification (<b>B</b>). Enrichment analysis of the amino acids present as a tree map (<b>C</b>).</p>
Full article ">Figure 7
<p>Pathway analysis of enzymes and amino acids related to amino acid metabolism, EMP (green arrows), HMP (orange arrows) and TCA cycle (purple arrows) in <span class="html-italic">P. fragi</span> with linalool treatment; red color represents upregulation of amino acids and blue color represents downregulation of amino acids.</p>
Full article ">Figure 8
<p>Validation of selected DEGs using real-time PCR. *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 9
<p>Molecular docking perspective of linalool with the binding sites of G6PDH (<b>A</b>), MDH (<b>B</b>) and PFK (<b>C</b>). Ligand: green color, macromolecule: pink color, receptor: orange color.</p>
Full article ">
21 pages, 9431 KiB  
Article
Anti-Skin Aging Potential, Antibacterial Activity, Inhibition of Single-Stranded DNA-Binding Protein, and Cytotoxic Effects of Acetone-Extracted Passiflora edulis (Tainung No. 1) Rind Extract on Oral Carcinoma Cells
by Yen-Hua Huang and Cheng-Yang Huang
Plants 2024, 13(16), 2194; https://doi.org/10.3390/plants13162194 - 8 Aug 2024
Viewed by 265
Abstract
The passion fruit, Passiflora edulis, recognized for its rich nutritional properties, has long been used for its varied ethnobotanical applications. This study investigates the therapeutic potential of P. edulis var. Tainung No. 1 rind extracts by examining their polyphenolic content (TPC), total [...] Read more.
The passion fruit, Passiflora edulis, recognized for its rich nutritional properties, has long been used for its varied ethnobotanical applications. This study investigates the therapeutic potential of P. edulis var. Tainung No. 1 rind extracts by examining their polyphenolic content (TPC), total flavonoid content (TFC), anti-skin aging activities against key enzymes such as elastase, tyrosinase, and hyaluronidase, and their ability to inhibit bacterial growth, single-stranded DNA-binding protein (SSB), and their cytotoxic effects on oral carcinoma cells. The acetone extract from the rind exhibited the highest levels of TPC, TFC, anti-SSB, and antibacterial activities. The antibacterial effectiveness of the acetone-extracted rind was ranked as follows: Escherichia coli > Pseudomonas aeruginosa > Staphylococcus aureus. A titration curve for SSB inhibition showed an IC50 value of 313.2 μg/mL, indicating the potency of the acetone extract in inhibiting SSB. It also significantly reduced the activity of enzymes associated with skin aging, particularly tyrosinase, with a 54.5% inhibition at a concentration of 100 μg/mL. Gas chromatography–mass spectrometry (GC–MS) analysis tentatively identified several major bioactive compounds in the acetone extract, including stigmast-5-en-3-ol, vitamin E, palmitic acid, stigmasterol, linoleic acid, campesterol, and octadecanoic acid. Molecular docking studies suggested some of these compounds as potential inhibitors of tyrosinase and SSB. Furthermore, the extract demonstrated anticancer potential against Ca9-22 oral carcinoma cells by inhibiting cell survival, migration, and proliferation and inducing apoptosis. These results underscore the potential of P. edulis (Tainung No. 1) rind as a promising candidate for anti-skin aging, antibacterial, and anticancer applications, meriting further therapeutic investigation. Full article
(This article belongs to the Special Issue Biological Activities of Plant Extracts 2023)
Show Figures

Figure 1

Figure 1
<p>Molecular docking analysis of tyrosinase. (<b>A</b>) The crystal structure of tyrosinase (PDB ID 2Y9X) is shown, with the complexed molecule 2-hydroxycyclohepta-2,4,6-trien-1-one removed for clarity. Active site copper ions are highlighted in brown. (<b>B</b>) Docking analysis demonstrating that four compounds from the extract exhibit higher binding affinities than kojic acid: campesterol (cyan), stigmasterol (orange), stigmast-5-en-3-ol (deepsalmon), and vitamin E (light magenta). These compounds have the capacity to dock into the tyrosinase active site, potentially obstructing substrate access and inhibiting enzyme activity through various binding poses. (<b>C</b>–<b>F</b>) Binding modes of stigmasterol, stigmast-5-en-3-ol, campesterol, and vitamin E with tyrosinase, illustrating their interactions within the enzyme’s active site. The blue dashed lines indicate hydrogen bonds, and the black dashed lines indicate hydrophobic interactions. The numbers labeled on the dashed lines represent the distances in angstroms.</p>
Full article ">Figure 2
<p>Inhibition of SSB by <span class="html-italic">P. edulis</span> rind extracts. (<b>A</b>) SSB binding to ssDNA. SSB from <span class="html-italic">K. pneumoniae</span> across a range of concentrations (0, 18, 37, 77, 155, 310, 625, 1250, 2500, and 5000 nM) was incubated with a biotinylated dT35 oligonucleotide. A streptavidin–horseradish peroxidase conjugate was used to detect the ssDNA and the resulting complexes. C indicates the formed complex. (<b>B</b>–<b>D</b>) Inhibition of ssDNA-binding activity of SSB by <span class="html-italic">P. edulis</span> rind extracts obtained using acetone (<b>B</b>), methanol (<b>C</b>), and ethanol (<b>D</b>). SSB (310 nM) was treated with varying concentrations of each extract (0, 31, 63, 125, 250, 500, 1000, 2000, and 3000 μg/mL) to assess inhibition of binding activity.</p>
Full article ">Figure 3
<p>Molecular docking analysis of SSB. (<b>A</b>) Crystal structure of <span class="html-italic">K. pneumoniae</span> SSB. <span class="html-italic">K. pneumoniae</span> SSB is a homotetramer. The charge distribution pattern is shown for clarity to indicate possible binding sites. (<b>B</b>) Docking analysis depicting the four most prevalent compounds from the extract, each individually docked into SSB: stigmasterol (orange), stigmast-5-en-3-ol (deepsalmon), campesterol (cyan), and vitamin E (lightmagenta). (<b>C</b>) The structure of <span class="html-italic">P. aeruginosa</span> SSB bound by ssDNA dT20. The ssDNA within the complex crystal structure of the <span class="html-italic">P. aeruginosa</span> SSB tetramer is highlighted in yellow. Given the unavailability of an ssDNA-complexed structure for <span class="html-italic">K. pneumoniae</span> SSB, the complexed structure of the <span class="html-italic">P. aeruginosa</span> SSB is utilized for comparative analysis of the ssDNA-binding mode in <span class="html-italic">K. pneumoniae</span> SSB. (<b>D</b>) Superimposed structures of ssDNA bound by <span class="html-italic">P. aeruginosa</span> SSB alongside the docked compounds bound by <span class="html-italic">K. pneumoniae</span> SSB, suggesting potential ssDNA-binding sites in <span class="html-italic">K. pneumoniae</span> SSB. (<b>E</b>–<b>G</b>) Binding modes of stigmasterol (<b>E</b>), stigmast-5-en-3-ol (<b>F</b>), and campesterol (<b>G</b>) to <span class="html-italic">K. pneumoniae</span> SSB are illustrated, highlighting their interactions within the binding sites.</p>
Full article ">Figure 4
<p>Anticancer potential of acetone-extracted <span class="html-italic">P. edulis</span> rind on Ca9-22 gingival carcinoma cells. (<b>A</b>) Overview of the rind extract’s influence on Ca9-22 cell viability, migration, proliferation, and nuclear condensation. (<b>B</b>) Results from the Trypan blue exclusion assay, illustrating cell viability post-treatment with varying concentrations of the rind extract. (<b>C</b>) Hoechst staining analysis depicting the extent of apoptosis and DNA fragmentation across different concentrations of the rind extract. (<b>D</b>) Wound healing assay images capturing Ca9-22 cell migration before and 24 h post-extraction treatment at various concentrations. (<b>E</b>) Clonogenic assay results evaluating the ability of Ca9-22 cells to form colonies under different concentrations of the rind extract, reflecting their survival and proliferative capabilities. Statistical significance relative to the control is indicated by * for <span class="html-italic">p</span> &lt; 0.05, ** for <span class="html-italic">p</span> &lt; 0.01, and *** for <span class="html-italic">p</span> &lt; 0.001. A control medium containing 1% DMSO served as the negative control and caused deaths at the rate of 0%, reduced migration by 0%, suppressed proliferation and colony formation by 0%, and did not induce apoptosis in Ca9-22 cells.</p>
Full article ">
23 pages, 3779 KiB  
Article
Vancomycin-Conjugated Polyethyleneimine-Stabilized Gold Nanoparticles Attenuate Germination and Show Potent Antifungal Activity against Aspergillus spp.
by Aishwarya Nikhil, Atul Kumar Tiwari, Ragini Tilak, Saroj Kumar, Prahlad Singh Bharti, Prem C. Pandey, Roger J. Narayan and Munesh Kumar Gupta
Appl. Sci. 2024, 14(16), 6926; https://doi.org/10.3390/app14166926 - 7 Aug 2024
Viewed by 329
Abstract
Antifungal drug resistance in filamentous fungi, particularly Aspergillus species, is increasing worldwide. Therefore, new antifungal drugs or combinations of drugs are urgently required to overcome this public health situation. In the present study, we examined the antifungal activity of vancomycin-functionalized AuNPs. These functionalized [...] Read more.
Antifungal drug resistance in filamentous fungi, particularly Aspergillus species, is increasing worldwide. Therefore, new antifungal drugs or combinations of drugs are urgently required to overcome this public health situation. In the present study, we examined the antifungal activity of vancomycin-functionalized AuNPs. These functionalized AuNPs were characterized, and their antifungal activity and associated killing mechanism were investigated using conventional methodologies against the conidia of A. fumigatus and A. flavus. The differential antifungal activity of vancomycin-functionalized Au-NPs against the conidia of Aspergillus species is dependent on structural differences in the conidial cell wall. The results demonstrated potent fungicidal activity against A. fumigatus, with a MIC value of 4.68 µg/mL, 93% germination inhibition, and 38.4% killing rate within 8 h of exposure. However, the activity against A. flavus was fungistatic; a MIC value of 18.7 µg/mL and 35% conidial germination inhibition, followed by 28.4% killing rate, were noted under similar conditions. Furthermore, endogenous reactive oxygen species (ROS) accumulation was 37.4 and 23.1% in conidial populations of A. fumigatus and A. flavus, respectively. Raman spectroscopy analysis confirmed the possible (but not confirmed) binding of functionalized AuNPs with the chitin and galactomannan components of the cell wall. A potential strategy that involves the exploration of antibacterial drugs using AuNPs as efficient drug carriers may also be appropriate for countering emerging drug resistance in filamentous fungi. Full article
(This article belongs to the Section Nanotechnology and Applied Nanosciences)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Physical characterization of PEI-AuNP@Van nanoparticles. (<b>a</b>) UV-Vis spectrum, (<b>b</b>) TEM micrograph, (<b>c</b>) high resolution image of nanoparticles, (<b>d</b>) mean nanoparticle size, (<b>e</b>) XRD diffractogram, and (<b>f</b>) dynamic light scattering data. (<b>B</b>) Zeta potential distribution histograms for non-functionalized gold nanoparticles (PEI-AuNPs) and vancomycin-functionalized gold nanoparticles (PEI-AuNP@Van).</p>
Full article ">Figure 1 Cont.
<p>(<b>A</b>) Physical characterization of PEI-AuNP@Van nanoparticles. (<b>a</b>) UV-Vis spectrum, (<b>b</b>) TEM micrograph, (<b>c</b>) high resolution image of nanoparticles, (<b>d</b>) mean nanoparticle size, (<b>e</b>) XRD diffractogram, and (<b>f</b>) dynamic light scattering data. (<b>B</b>) Zeta potential distribution histograms for non-functionalized gold nanoparticles (PEI-AuNPs) and vancomycin-functionalized gold nanoparticles (PEI-AuNP@Van).</p>
Full article ">Figure 2
<p>Fluorescence spectroscopic confirmation of vancomycin functionalization of PEI-stabilized gold nanoparticles. Adopted and modified under CC BY (2023) [<a href="#B33-applsci-14-06926" class="html-bibr">33</a>].</p>
Full article ">Figure 3
<p>Antifungal evaluation plates (<b>a</b>) against <span class="html-italic">A. fumigatus</span> and (<b>b</b>) <span class="html-italic">A. flavus</span> (NC = PEI-AuNPs control; NP = PEI-AuNP@Van; PC = positive control). (<b>c</b>) MIC values of non-functionalized AuNPs (PEI-AuNPs), vancomycin-functionalized AuNPs (PEI-AuNP@Van), positive control (voriconazole), vancomycin, and negative control (DW) against <span class="html-italic">A. fumigatus</span> and <span class="html-italic">A. flavus</span>.</p>
Full article ">Figure 4
<p>Conidial germination inhibition assay.</p>
Full article ">Figure 5
<p>Confocal micrograph of PEI-AuNP@Van treated conidia of <span class="html-italic">Aspergillus</span>. (<b>a</b>) Bright field image of <span class="html-italic">A. fumigatus</span> treated with PEI-AuNP@Van and (<b>b</b>) stained with PI; (<b>c</b>) blue panel showing surface adsorbed functionalized nanoparticles, and (<b>d</b>) merged panel; (<b>e</b>) bright field image of <span class="html-italic">A. flavus</span> treated with PEI-AuNP@Van and (<b>f</b>) stained with PI; (<b>g</b>) blue panel showing surface adsorbed functionalized nanoparticles, and (<b>h</b>) merged panel.</p>
Full article ">Figure 6
<p>Histograms of dead conidia after 8 h of exposure to PEI-AuNP@Van. (<b>a</b>,<b>c</b>) represent the untreated controls of <span class="html-italic">A. fumigatus</span> and <span class="html-italic">A. flavus</span>, respectively; (<b>b</b>,<b>d</b>) represent PEI-AuNP@Van-treated conidia of <span class="html-italic">A. fumigatus</span> and <span class="html-italic">A. flavus</span>.</p>
Full article ">Figure 7
<p>Endogenous ROS accumulation in PEI-AuNP@Van-treated conidia. (<b>a</b>,<b>c</b>) show the untreated control conidia of <span class="html-italic">A. fumigatus</span> and <span class="html-italic">A. flavus</span>, respectively; (<b>b</b>,<b>d</b>) represent the PEI-AuNP@Van-treated conidia.</p>
Full article ">Figure 8
<p>Raman spectrums of voriconazole and PEI-AuNP@Van-treated conidia of <span class="html-italic">A. flavus</span> and <span class="html-italic">A. fumigatus</span> along with an untreated control. (<b>ai</b>) Untreated conidia of <span class="html-italic">A. flavus</span>, (<b>aii</b>) treated with voriconazole, (<b>aiii</b>) treated with PEI-AuNP@Van, (<b>bi</b>) untreated conidia of <span class="html-italic">A. fumigatus</span>, (<b>bii</b>) treated with voriconazole, and (<b>biii</b>) treated with PEI-AuNP@Van.</p>
Full article ">Figure 9
<p>TEM micrograph of PEI-AuNP@Van-treated conidia of <span class="html-italic">A. fumigatus</span> and <span class="html-italic">A. flavus</span>: (<b>a</b>) represents the untreated conidia of <span class="html-italic">A. fumigatus</span>, and (<b>b</b>) treated with PEI-AuNP@Van; (<b>c</b>) represents the untreated conidia of <span class="html-italic">A. flavus</span>, and (<b>d</b>) treated with PEI-AuNP@Van.</p>
Full article ">
21 pages, 3340 KiB  
Article
Development of Xanthoangelol-Derived Compounds with Membrane-Disrupting Effects against Gram-Positive Bacteria
by Siyu Yang, Fangquan Liu, Yue Leng, Meiyue Zhang, Lei Zhang, Xuekun Wang and Yinhu Wang
Antibiotics 2024, 13(8), 744; https://doi.org/10.3390/antibiotics13080744 - 7 Aug 2024
Viewed by 398
Abstract
Infections caused by multidrug-resistant pathogens have emerged as a serious threat to public health. To develop new antibacterial agents to combat such drug-resistant bacteria, a class of novel amphiphilic xanthoangelol-derived compounds were designed and synthesized by mimicking the structure and function of antimicrobial [...] Read more.
Infections caused by multidrug-resistant pathogens have emerged as a serious threat to public health. To develop new antibacterial agents to combat such drug-resistant bacteria, a class of novel amphiphilic xanthoangelol-derived compounds were designed and synthesized by mimicking the structure and function of antimicrobial peptides (AMPs). Among them, compound 9h displayed excellent antimicrobial activity against the Gram-positive strains tested (MICs = 0.5–2 μg/mL), comparable to vancomycin, and with low hemolytic toxicity and good membrane selectivity. Additionally, compound 9h demonstrated rapid bactericidal effects, low resistance frequency, low cytotoxicity, and good plasma stability. Mechanistic studies further revealed that compound 9h had good membrane-targeting ability and was able to destroy the integrity of bacterial cell membranes, causing an increase in intracellular ROS and the leakage of DNA and proteins, thus accelerating bacterial death. These results make 9h a promising antimicrobial candidate to combat bacterial infection. Full article
(This article belongs to the Topic Antimicrobial Agents and Nanomaterials)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Structure of CAS-13, XF-73, and PMX-30063.</p>
Full article ">Figure 2
<p>Design concept for amphiphilic xanthoangelol-derived compounds by mimicking the structure and the biological function of AMPs.</p>
Full article ">Figure 3
<p>Plasma stability and cytotoxicity of compound <b>9h.</b> Plasma stability (<b>A</b>) and bactericidal activity in complex mammalian fluids (<b>B</b>), and in vitro cytotoxicity (<b>C</b>) of <b>9h</b> toward LO2 cells. Data are expressed as mean ± standard deviation (n = 3).</p>
Full article ">Figure 4
<p>Time-kill kinetics (<b>A</b>) and bacterial resistance study (<b>B</b>) of <b>9h</b> against <span class="html-italic">S. aureus</span> ATCC43300. Data are expressed as mean ± standard deviation (n = 3).</p>
Full article ">Figure 5
<p>Antibiofilm activity of <b>9h</b>. (<b>A</b>) Inhibition of <span class="html-italic">S. aureus</span> biofilm formation by <b>9h</b>. (<b>B</b>) Eradication of the preformed <span class="html-italic">S. aureus</span> biofilm by <b>9h</b>. Error bars represent standard deviation from the mean of triplicate readout. * <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, and **** <span class="html-italic">p</span> &lt; 0.0001, compared with the control group.</p>
Full article ">Figure 6
<p>Fluorescence and electron scanning microscopy. (<b>A</b>) SEM images of the cell membrane of <span class="html-italic">S. aureus</span> cells, scar bar: 1.00 um. (<b>B</b>) Fluorescence micrographs of <span class="html-italic">S. aureus</span> cells stained with DAPI and PI and treated with <b>9h,</b> scar bar: 50 um.</p>
Full article ">Figure 7
<p>Antibacterial mechanism of compound <b>9h</b>. (<b>A</b>) Cytoplasmic membrane depolarization against <span class="html-italic">S. aureus</span> by compound <b>9h</b>. (<b>B</b>) Cell membrane permeabilization against <span class="html-italic">S. aureus</span> by compound <b>9h</b>. Data are expressed as mean ± standard deviation (n = 3).</p>
Full article ">Figure 8
<p>Antibacterial mechanism of compound <b>9h</b> against <span class="html-italic">S. aureus</span> ATCC43300. (<b>A</b>) Intracellular ROS changes after the treatment of <b>9h</b>. (<b>B</b>) DNA leakage caused by compound <b>9h</b>. (<b>C</b>) Protein leakage caused by compound <b>9h</b>. Data are expressed as mean ± standard deviation (n = 3). ns, not significant. * <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, and **** <span class="html-italic">p</span> &lt; 0.0001, compared with the control group.</p>
Full article ">Scheme 1
<p>Reagents and conditions: (a) pyridine, 150 °C, 10 h; (b) K<sub>2</sub>CO<sub>3</sub>, 1,3-dibromopropane, CH<sub>3</sub>CN, 60 °C, 8 h; (c) RH, K<sub>2</sub>CO<sub>3,</sub> CH<sub>3</sub>CN, 60 °C, 8 h.</p>
Full article ">Scheme 2
<p>Reagents and conditions: (a) pyridine, 150 °C, 18 h; (b) K<sub>2</sub>CO<sub>3</sub>, CH<sub>3</sub>CN, 60 °C, 10 h; (c) RH, K<sub>2</sub>CO<sub>3,</sub> CH<sub>3</sub>CN, rt, 8 h.</p>
Full article ">
24 pages, 1559 KiB  
Review
Potential Possibilities of Using Peat, Humic Substances, and Sulfurous Waters in Cosmetology
by Ewelina Maria Błońska-Sikora, Marta Klimek-Szczykutowicz, Monika Michalak, Katarzyna Kulik-Siarek and Małgorzata Wrzosek
Appl. Sci. 2024, 14(16), 6912; https://doi.org/10.3390/app14166912 - 7 Aug 2024
Viewed by 632
Abstract
Balneology is one of the oldest fields of medicine related to the use of natural raw materials (medicinal waters, medicinal gases, peloids, climatic values) in the treatment, prevention, and rehabilitation of many diseases but also increasingly in cosmetology. Currently, balneotherapy (spa therapy) combines [...] Read more.
Balneology is one of the oldest fields of medicine related to the use of natural raw materials (medicinal waters, medicinal gases, peloids, climatic values) in the treatment, prevention, and rehabilitation of many diseases but also increasingly in cosmetology. Currently, balneotherapy (spa therapy) combines tradition and modernity. The interest in spa treatments, the popularity of a healthy lifestyle, as well as the constant search for active substances of a natural origin for cosmetics make peloids, medicinal, and mineral waters very popular in the cosmetics industry. The main aim of this review was to present current, scientifically proven knowledge about the potential use of peat, huic substances, and sulfurous water in cosmetology. The work describes the potential possibilities of using medicinal waters, especially sulfurous waters, as well as peats and humic compounds, which are the source of active substances with biological activity e.g., antibacterial, anti-inflammatory, and antioxidant, and possess a positive effect on psoriasis, atopic dermatitis, or acne. The therapeutic effects of these substances have been well documented in the literature; however, the validity of their use in cosmetology requires further confirmation. Full article
Show Figures

Figure 1

Figure 1
<p>Divisions of balneotherapy [<a href="#B4-applsci-14-06912" class="html-bibr">4</a>,<a href="#B5-applsci-14-06912" class="html-bibr">5</a>,<a href="#B6-applsci-14-06912" class="html-bibr">6</a>].</p>
Full article ">Figure 2
<p>Factors influencing the effectiveness of peat treatments [<a href="#B93-applsci-14-06912" class="html-bibr">93</a>,<a href="#B94-applsci-14-06912" class="html-bibr">94</a>,<a href="#B95-applsci-14-06912" class="html-bibr">95</a>].</p>
Full article ">Figure 3
<p>Division of medicinal waters [<a href="#B98-applsci-14-06912" class="html-bibr">98</a>].</p>
Full article ">
16 pages, 1798 KiB  
Article
Assessment of Photoactivated Chlorophyllin Production of Singlet Oxygen and Inactivation of Foodborne Pathogens
by Cristina Pablos, Javier Marugán, Rafael van Grieken, Jeremy W. J. Hamilton, Nigel G. Ternan and Patrick S. M. Dunlop
Catalysts 2024, 14(8), 507; https://doi.org/10.3390/catal14080507 - 6 Aug 2024
Viewed by 461
Abstract
Singlet oxygen (1O2) is known to have antibacterial activity; however, production can involve complex processes with expensive chemical precursors and/or significant energy input. Recent studies have confirmed the generation of 1O2 through the activation of photosensitizer molecules [...] Read more.
Singlet oxygen (1O2) is known to have antibacterial activity; however, production can involve complex processes with expensive chemical precursors and/or significant energy input. Recent studies have confirmed the generation of 1O2 through the activation of photosensitizer molecules (PSs) with visible light in the presence of oxygen. Given the increase in the incidence of foodborne diseases associated with cross-contamination in food-processing industries, which is becoming a major concern, food-safe additives, such as chlorophyllins, have been studied for their ability to act as PSs. The fluorescent probe Singlet Oxygen Sensor Green (SOSG®) was used to estimate 1O2 formation upon the irradiation of traditional PSs (rose bengal (RB), chlorin 6 (ce6)) and novel chlorophyllins, sodium magnesium (NaChl) and sodium copper (NaCuChl), with both simulated-solar and visible light. NaChl gave rise to a similar 1O2 production rate when compared to RB and ce6. Basic mixing was shown to introduce sufficient oxygen to the PS solutions, preventing the limitation of the 1O2 production rate. The NaChl-based inactivation of Gram-positive S. aureus and Gram-negative E. coli was demonstrated with a 5-log reduction with UV–Vis light. The NaChl-based inactivation of Gram-positive S. aureus was accomplished with a 2-log reduction after 105 min of visible-light irradiation and a 3-log reduction following 150 min of exposure from an initial viable bacterial concentration of 106 CFU mL−1. CHS-NaChl-based photosensitization under visible light enhanced Gram-negative E. coli inactivation and provided a strong bacteriostatic effect preventing E. coli proliferation. The difference in the ability of NaChl and CHS-NaChl complexes to inactivate Gram-positive and Gram-negative bacteria was confirmed to result from the cell wall structure, which impacted PS–bacteria attachment and therefore the production of localized singlet oxygen. Full article
(This article belongs to the Special Issue Photocatalysis towards a Sustainable Future)
Show Figures

Figure 1

Figure 1
<p>Green fluorescence enhancement due to <sup>1</sup>O<sub>2</sub> generation arising from the photoirradiation of a 0.5 μM PS solution in deionized water and 2 µM SOSG under UV–Vis radiation.</p>
Full article ">Figure 2
<p>Singlet oxygen production rate as a function of initial solution pH during irradiation with UV–Vis photons. Photosensitizer initial concentration = 0.5 µM.</p>
Full article ">Figure 3
<p>Effect of the increase in oxygen in the reaction solution on the singlet oxygen production rate in deionized water throughout the irradiation with UV–Vis light of different photosensitizers at 0.5 µM as the initial concentration. Significant differences (<span class="html-italic">p</span> ≤ 0.05) have been indicated with an asterisk.</p>
Full article ">Figure 4
<p>Inactivation of <span class="html-italic">S. aureus</span> and <span class="html-italic">E. coli</span> (figure inset) by chlorophyllin-based photosensitization in ¼ strength Ringer’s solution under UV–Vis light. PSs: 0.5 µM NaChl, and the complex 0.1% CHS-NaChl 0.00004% (0.5 µM).</p>
Full article ">Figure 5
<p>Inactivation of <span class="html-italic">S. aureus</span> by chlorophyllin-based photosensitization in ¼ strength Ringer’s solution under visible light: (<b>a</b>) PS: 0.5 µM NaChl. Bacterial initial concentration: 10<sup>6</sup> CFU∙mL<sup>−1</sup>. (<b>b</b>) PSs: 0.5 µM NaChl and 0.5 µM NaChl under a high concentration of oxygen corresponding to an increase of 20%; 0.5 µM NaCuChl; and the complex CHS (0.1%)-NaChl (0.00004%, ca. 0.5 µM). Bacterial initial concentration: 10<sup>3</sup> CFU∙mL<sup>−1</sup>.</p>
Full article ">Figure 6
<p>Inactivation of <span class="html-italic">E. coli</span> by chlorophyllin-based photosensitization in ¼ strength Ringer’s solution under visible light. PSs: 0.5 µM NaChl and the complex CHS (0.1%)-NaChl (0.00004%, ca. 0.5 µM).</p>
Full article ">
Back to TopTop