[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (695)

Search Parameters:
Keywords = enzyme labels

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
24 pages, 5556 KiB  
Article
Differential Mitochondrial Redox Responses to the Inhibition of NAD+ Salvage Pathway of Triple Negative Breast Cancer Cells
by Jack Kollmar, Junmei Xu, Diego Gonzalves, Joseph A. Baur, Lin Z. Li, Julia Tchou and He N. Xu
Cancers 2025, 17(1), 7; https://doi.org/10.3390/cancers17010007 - 24 Dec 2024
Abstract
Background/Objectives: Cancer cells rely on metabolic reprogramming that is supported by altered mitochondrial redox status and an increased demand for NAD+. Over expression of Nampt, the rate-limiting enzyme of the NAD+ biosynthesis salvage pathway, is common in breast cancer [...] Read more.
Background/Objectives: Cancer cells rely on metabolic reprogramming that is supported by altered mitochondrial redox status and an increased demand for NAD+. Over expression of Nampt, the rate-limiting enzyme of the NAD+ biosynthesis salvage pathway, is common in breast cancer cells, and more so in triple negative breast cancer (TNBC) cells. Targeting the salvage pathway has been pursued for cancer therapy. However, TNBC cells have heterogeneous responses to Nampt inhibition, which contributes to the diverse outcomes. There is a lack of imaging biomarkers to differentiate among TNBC cells under metabolic stress and identify which are responsive. We aimed to characterize and differentiate among a panel of TNBC cell lines under NAD-deficient stress and identify which subtypes are more dependent on the NAD salvage pathway. Methods: Optical redox imaging (ORI), a label-free live cell imaging microscopy technique was utilized to acquire intrinsic fluorescence intensities of NADH and FAD-containing flavoproteins (Fp) thus the mitochondrial redox ratio Fp/(NADH + Fp) in a panel of TNBC cell lines. Various fluorescence probes were then added to the cultures to image the mitochondrial ROS, mitochondrial membrane potential, mitochondrial mass, and cell number. Results: Various TNBC subtypes are sensitive to Nampt inhibition in a dose- and time-dependent manner, they have differential mitochondrial redox responses; furthermore, the mitochondrial redox indices linearly correlated with mitochondrial ROS induced by various doses of a Nampt inhibitor. Moreover, the changes in the redox indices correlated with growth inhibition. Additionally, the redox state was found fully recovered after removing the Nampt inhibitor. Conclusions: This study supports the utility of ORI in rapid metabolic phenotyping of TNBC cells under NAD-deficient stress to identify responsive cells and biomarkers of treatment response, facilitating combination therapy strategies. Full article
(This article belongs to the Section Methods and Technologies Development)
Show Figures

Figure 1

Figure 1
<p>The critical role of NAD as a co-enzyme and co-substrate (<b>A</b>) and experimental schematic (<b>B</b>), where NAM stands for nicotinamide, NMN for nicotinamide mononucleotide, TCA for tricarboxylic acid cycle, LDH for lactate dehydrogenase.</p>
Full article ">Figure 2
<p>Typical images of MDA-MB-468 cells subjected to various imaging. (<b>A</b>–<b>E</b>) white light image and raw images of Fp and NADH, mitochondrial membrane potential (MMP) represented by TMRE intensity, and mitochondrial mass represented by MitoView Green intensity, respectively, for control. (<b>F</b>–<b>J</b>) display the processed images of those shown in (<b>A</b>–<b>E</b>). (<b>K</b>–<b>O</b>) the processed images of MDA-MB-468 cells treated with 50 nM FK866 for 48 h. The intensity bars for the raw images are in arbitrary unit and are set for better visualization of signal dynamic range. The intensity bars for the pseudocolored images represent the pixel values of the corresponding images with the redder color indicating higher values. The numbers in white color in the processed images are the means and standard deviations of the respective images. (<b>P</b>) the mean values and standard deviations of the redox indices of images shown in (<b>F</b>–<b>H</b>) (control) and (<b>K</b>–<b>M</b>) (treated); (<b>Q</b>) the mean values and standard deviations of the MMP shown in images (<b>I</b>,<b>N</b>); (<b>R</b>) the mean values and standard deviations of the mitochondrial mass of images shown in (<b>J</b>,<b>O</b>). The error bars represent the standard deviations of the pixel values in the corresponding images.</p>
Full article ">Figure 3
<p>Dose-dependent mitochondrial redox responses observed in the TNBC cells. (<b>A</b>) Redox responses of MDA-MB-468 cells; (<b>B</b>) Redox responses of MDA-MB-436 cells; (<b>C</b>) Redox responses of MDA-MB-453 cells treated with 0–100 nM of FK866 for 48 h. Bars: mean ± SD, black circles indicating individual dishes. *, <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, ****, <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 4
<p>The redox responses represented as the percentage change from baseline for Fp, NADH, and the redox ratio in five human TNBC cell lines treated with 1 nM FK866 for 48 h. Bars: mean ± SD, black circles indicating individual dishes. Asterisks indicate the comparison between the redox indices of control (untreated) and treated cells. *, <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, ****, <span class="html-italic">p</span> &lt; 0.0001. Note: the data for HCC1806 were extracted from our previous report [<a href="#B35-cancers-17-00007" class="html-bibr">35</a>].</p>
Full article ">Figure 5
<p>Heatmaps of the adjusted <span class="html-italic">p</span> values for multiple comparisons between cell lines of the corresponding redox index changes shown in <a href="#cancers-17-00007-f004" class="html-fig">Figure 4</a> and <a href="#cancers-17-00007-t002" class="html-table">Table 2</a>.</p>
Full article ">Figure 6
<p>Nampt inhibitor dose-dependent redox changes observed in E0771 cells after 20 h treatment with (<b>A</b>) various doses of FK866 and (<b>B</b>) various doses of GMX1778. Bars: mean ± SD, black circles indicating individual dishes. *, <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, ****, <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 7
<p>Temporal ORI responses of TNBC cells to FK866 inhibition. (<b>A</b>) MDA-MB-231 cells and (<b>B</b>) MDA-MB-436 cells to 1 nM FK866 treatment for 24 or 48 h treatment; (<b>C</b>) HCC1806 cells were first treated with 1 nM FK866 for 48 h then allowed 24 h to recover (Rec) after removal of FK866. Bars: mean ± SD, black circles indicating individual dishes. Asterisks by themselves indicate the comparisons between the values of control (untreated) and treatment. Asterisks with brackets in (<b>B</b>) indicate the comparisons between the values of 24 h and 48 h treatment. *, <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, ****, <span class="html-italic">p</span> &lt; 0.0001. <span class="html-italic">p</span> values for multiple comparisons between different time points in (<b>B</b>) were adjusted by Bonferroni post hoc method.</p>
Full article ">Figure 8
<p>NR rescue effects on MDA-MB-453 cells under 100 nM FK866 for 48 h. (<b>A</b>) Both NR and NAM are converted to NMN but via different enzymes. (<b>B</b>) NR (1 mM) rescue effect. Bars: mean ± SD, black circles indicating individual dishes. *, <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, ****, <span class="html-italic">p</span> &lt; 0.0001. Note: all redox indices were normalized to their respective control.</p>
Full article ">Figure 9
<p>ORI responses of AT-3 cells to Nampt inhibition and NR rescue effect. 500 nM GMX1778, 1 mM NR, or 500 nM GMX1778, 1 mM NR treatment, or NR and NAM co-treatment for 24 h. Bars: mean ± SD, black circles indicating individual dishes. *, <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; 0.0001.</p>
Full article ">Figure 10
<p>(<b>A</b>) ORI responses and changes in mitochondrial ROS of AT-3 cells treated with various doses of GMX1778 for 24 h. (<b>B</b>–<b>D</b>) The significant linear correlations between mitochondrial ROS and Fp, NADH, and the redox ratio, respectively, determined from data shown in (<b>A</b>), where R<sup>2</sup> and <span class="html-italic">p</span> values are indicated on the graphs. Bars: mean ± SD, black circles indicating individual dishes. Dashed lines: 95% confidence intervals *, <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, ****, <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 11
<p>The effects of FK866 on the mitochondrial membrane potential and mitochondrial mass indicated by fluorescence probes TMRE or MitoView Green. (<b>A</b>) A large decrease (~40%) in the mitochondrial membrane potential in MDA-MB-468 cells treated with 50 nM FK866 for 48 h; (<b>B</b>) insignificant change in the mitochondrial mass of MDA-MB-468; (<b>C</b>) dose-dependent increase in the mitochondrial mass of E0771 cells treated with either 1 nM or 100 nM FK866 for 24 h. Bars: mean ± SD, black circles indicating individual FOVs. ****, <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 12
<p>The correlations between the mitochondrial redox indices and cell growth at various doses of FK866 and two treatment periods for E0771 cells. (<b>A</b>) The dose-dependent changes of the mitochondrial redox indices (note that this figure is the same as that shown in <a href="#cancers-17-00007-f006" class="html-fig">Figure 6</a>A); (<b>B</b>) The dose-dependent cell growth retardation due to 20 h exposure to various doses of FK866; (<b>C</b>–<b>E</b>) The significant linear correlations between the redox indices and cell growth under 20 h exposure to various doses of FK866, where R<sup>2</sup> and <span class="html-italic">p</span> values are indicated on the graphs. (<b>F</b>,<b>G</b>) The dose-dependent changes of the mitochondrial redox indices and suppressed cell growth due to 48 h treatment with various doses of FK866, respectively. Bars: mean ± SD, black circles indicating individual dishes. *, <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, ****, <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 13
<p>Effects of 1 nM FK866 on NAD<sup>+</sup> and NADH of E0771 cell homogenates after 24 h exposure. Data were obtained with two technical and two biological replicates. Bars: mean ± SD. *, <span class="html-italic">p</span> &lt; 0.05, **, <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure A1
<p>To confirm the redox responses to metabolic modulations, we performed redox titration using MDA-MB-468 cells under normal growth conditions. We observed the expected changes, i.e., a decrease in NADH due to uncoupled oxidative phosphorylation from the electron transport chain by mitochondrial uncoupler FCCP, followed by an increase in NADH due to the inhibition of complex I and III by rotenone and antimycin A (ROT/AA), respectively. Bars: mean ± SD, N = 3–4 FOVs. *, <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; 0.0001.</p>
Full article ">Figure A2
<p>Temporal redox changes of E0771 cells due to various doses of FK866 treatment. Bars: mean ± SD. *, <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.</p>
Full article ">Figure A3
<p>LADH A specific inhibitor FX11 (5 µM) added to MDA-MB-231 cells exposed to 1 nM FK866 for 24 h immediately yielded a large spike of NADH and a reductive shift of the mitochondrial redox state. Bars: mean ± SD, black circles indicating individual samples. *, <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.</p>
Full article ">
12 pages, 1538 KiB  
Article
Application of the Chitooligosaccharides and Fluorescence Polarization Technique for the Assay of Active Lysozyme in Hen Egg White
by Liliya I. Mukhametova, Dmitry O. Zherdev, Sergei A. Eremin, Pavel A. Levashov, Hans-Christian Siebert, Yury E. Tsvetkov, Olga N. Yudina, Vadim B. Krylov and Nikolay E. Nifantiev
Biomolecules 2024, 14(12), 1589; https://doi.org/10.3390/biom14121589 - 12 Dec 2024
Viewed by 376
Abstract
This study describes the applicability of the fluorescence polarization assay (FPA) based on the use of FITC-labeled oligosaccharide tracers of defined structure for the measurement of active lysozyme in hen egg white. Depending on the oligosaccharide chain length of the tracer, this method [...] Read more.
This study describes the applicability of the fluorescence polarization assay (FPA) based on the use of FITC-labeled oligosaccharide tracers of defined structure for the measurement of active lysozyme in hen egg white. Depending on the oligosaccharide chain length of the tracer, this method detects both the formation of the enzyme-to-tracer complex (because of lectin-like, i.e., carbohydrate-binding action of lysozyme) and tracer splitting (because of chitinase activity of lysozyme). Evaluation of the fluorescence polarization dynamics enables simultaneous measurement of the chitinase and lectin activities of lysozyme, which is crucial for its detection in complex biological systems. Hen egg white lysozyme (HEWL), unlike human lysozyme (HL), formed a stable complex with the chitotriose tracer that underwent no further transformations. This fact allows for easy measurement of the carbohydrate-binding activity of the HEWL. The results of the lysozyme activity measurement for hen egg samples obtained through the FPA correlated with the results obtained using the traditional turbidimetry method. The FPA does not have the drawbacks of turbidimetry, which are associated with the need to use bacterial cells that cannot be precisely standardized. Additionally, FPA offers advantages such as rapid analysis, the use of compact equipment, and standardized reagents. Therefore, the new express technique for measuring the lysozyme activity is applicable for evaluating the complex biomaterial, including for the purposes of food product quality control. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Structure of oligosaccharide tracers <b>1–3</b>.</p>
Full article ">Figure 2
<p>Change in the FP signal of fluorescently labeled conjugates <b>1</b>–<b>3</b> in presence of hen egg white lysozyme (HEWL) (<b>a</b>) and human lysozyme (HL) (<b>b</b>). The black bars represent the fluorescence polarization (FP) of working solutions containing tracers <b>1</b>–<b>3</b> in the absence of proteins, while the gray bars indicate the FP immediately after the addition of the analyte with lysozyme. A significant increase in FP for tracer <b>1</b> was observed only after the addition of HEWL, but not HL. The time-dependent dynamics of the FP signal were monitored for tracers <b>1</b>–<b>3</b> in the presence of HEWL (<b>c</b>). The FP signal remained consistently high for tracer <b>1</b> but decreased over time for tracers <b>2</b> and <b>3</b>.</p>
Full article ">Figure 3
<p>FP signal of tracer <b>1</b> in the presence of different amounts of HEWL and approximation calibration curve for the variation in fluorescence polarization (mP) as a function of HEWL concentration. IC<sub>50</sub> = 76.9 ± 6.5 mg/mL; Hillslope = −1.3 ± 0.1, R<sup>2</sup> = 0.99.</p>
Full article ">Figure 4
<p>Dynamics of FP signal change in time for tracer <b>2</b> in the presence of different amounts of HEWL (<b>a</b>); dependence of FP decay rate on HEWL concentration (calibration dependence) and its linear approximation (R<sup>2</sup> = 0.98) (<b>b</b>).</p>
Full article ">Figure 5
<p>Comparison of FPA methods to determine lysozyme activity using the tracers <b>1</b> (<b>a</b>) and <b>2</b> (<b>b</b>), and the traditional turbidimetric method.</p>
Full article ">
20 pages, 10537 KiB  
Article
Growth, Quality, and Nitrogen Metabolism of Medicago sativa Under Continuous Light from Red–Blue–Green LEDs Responded Better to High Nitrogen Concentrations than Under Red–Blue LEDs
by Ren Chen, Yanqi Chen, Kunming Lin, Yiming Ding, Wenke Liu and Shurong Wang
Int. J. Mol. Sci. 2024, 25(23), 13116; https://doi.org/10.3390/ijms252313116 - 6 Dec 2024
Viewed by 412
Abstract
Alfalfa is a widely grown forage with a high crude protein content. Clarifying the interactions between light quality and nitrogen level on yield and nitrogen metabolism can purposely improve alfalfa productivity in plant factories with artificial light (PFAL). In this study, the growth, [...] Read more.
Alfalfa is a widely grown forage with a high crude protein content. Clarifying the interactions between light quality and nitrogen level on yield and nitrogen metabolism can purposely improve alfalfa productivity in plant factories with artificial light (PFAL). In this study, the growth, quality, and nitrogen metabolism of alfalfa grown in PFAL were investigated using three nitrate-nitrogen concentrations (10, 15, and 20 mM, labeled as N10, N15, and N20) and continuous light (CL) with two light qualities (red–blue and red–blue–green light, labeled as RB-C and RBG-C). The results showed that the adaptation performance of alfalfa to nitrogen concentrations differed under red–blue and red–blue–green CL. Plant height, stem diameter, leaf area, yield, Chl a + b, Chl a, Chl b, crude protein contents, and NiR activity under the RB-CN15 treatment were significantly higher than RB-CN10 and RB-CN20 treatments. The RB-CN20 treatment showed morphological damage, such as plant dwarfing and leaf chlorosis, and physiological damage, including the accumulation of proline, H2O2, and MDA. However, the difference was that under red–blue–green CL, the leaf area, yield, and Chl a + b, carotenoid, nitrate, and glutamate contents under RBG-CN20 treatment were significantly higher than in the RBG-CN10 and RBG-CN15 treatments. Meanwhile, the contents of soluble sugar, starch, and cysteine were significantly lower. However, the crude protein content reached 21.15 mg·g−1. The fresh yield, dry yield, stomatal conductance, leaf area, plant height, stem diameter, crude protein, GS, and free amino acids of alfalfa were positively correlated with increased green light. In addition, with the increase in nitrogen concentration, photosynthetic capacity, NiR, and GOGAT activities increased, promoting growth and improving feeding value. The growth, yield, photosynthetic pigments, carbon, nitrogen substances, and enzyme activities of alfalfa were significantly affected by the interaction between nitrogen concentration and light quality, whereas leaf/stem ratio and DPPH had no effect. In conclusion, RB-CN15 and RBG-CN20 are suitable for the production of alfalfa in PFAL, and green light can increase the threshold for the nitrogen concentration adaptation of alfalfa. Full article
(This article belongs to the Section Molecular Plant Sciences)
Show Figures

Figure 1

Figure 1
<p>Appearance of alfalfa under three nitrogen concentrations and CL with two light qualities.</p>
Full article ">Figure 2
<p>Growth characteristics and yield of alfalfa under three nitrogen concentrations with red–blue and red–blue–green CL. (<b>A</b>) Plant height, (<b>B</b>) stem diameter, (<b>C</b>) leaf area, (<b>D</b>) fresh yield, (<b>E</b>) dry yield, (<b>F</b>) fresh/dry ratio, (<b>G</b>) leaf/stem ratio, and (<b>H</b>) specific leaf area. Data are means ± SE; <span class="html-italic">n</span> = 5. Error bars with different letters show a significant difference (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>Photosynthetic pigment content, stomatal conductance, and <span class="html-italic">Fv</span>/<span class="html-italic">Fm</span> of alfalfa leaves under red and blue light and green light instead of red light. (<b>A</b>) Chl a + b content, (<b>B</b>) Chl a content, (<b>C</b>) Chl b content, (<b>D</b>) carotenoid content, (<b>E</b>) stomatal conductance, and (<b>F</b>) <span class="html-italic">Fv</span>/<span class="html-italic">Fm</span>. Data are means ± SE; <span class="html-italic">n</span> = 5. Error bars with different letters show a significant difference (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Soluble sugar (<b>A</b>), sucrose (<b>B</b>), starch (<b>C</b>), soluble protein (<b>D</b>), free amino acid (<b>E</b>), and crude protein (<b>F</b>) contents in alfalfa under three nitrogen concentrations and CL with two light qualities. Error bars with different letters show a significant difference (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>DPPH free radical clearance rate (<b>A</b>), Proline (<b>B</b>), H<sub>2</sub>O<sub>2</sub> (<b>C</b>), and MDA (<b>D</b>) contents in alfalfa under three nitrogen concentrations and CL with two light qualities. Error bars with different letters show a significant difference (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 6
<p>Nitrate (<b>A</b>) and ammonium (<b>B</b>) contents and activities of NR (<b>C</b>) and NiR (<b>D</b>) in alfalfa under three nitrogen concentrations and CL with two light qualities. Error bars with different letters show a significant difference (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 7
<p>Glutamate (<b>A</b>), cysteine (<b>B</b>), and lysine (<b>C</b>) contents and activities of GOGAT (<b>D</b>), GS (<b>E</b>), GDH (<b>F</b>), CS (<b>G</b>), and AK (<b>H</b>) in alfalfa under three nitrogen concentrations and CL with two light qualities. Error bars with different letters show a significant difference (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 8
<p>Correlation and heatmap analysis of growth characteristics, yield, photosynthetic pigment content, stomatal conductance, and <span class="html-italic">Fv</span>/<span class="html-italic">Fm</span> under three nitrogen concentrations and CL with two light qualities. Note. * Significant at the <span class="html-italic">p</span> &lt; 0.05 level of probability, ** significant at the <span class="html-italic">p</span> &lt; 0.01 level of probability, *** significant at the <span class="html-italic">p</span> &lt; 0.001 level of probability.</p>
Full article ">Figure 9
<p>Correlation and heatmap analysis of carbon and nitrogen substances and enzyme activities under three nitrogen concentrations and CL with two light qualities. Note. * Significant at the <span class="html-italic">p</span> &lt; 0.05 level of probability, ** significant at the <span class="html-italic">p</span> &lt; 0.01 level of probability, *** significant at the <span class="html-italic">p</span> &lt; 0.001 level of probability.</p>
Full article ">
16 pages, 3636 KiB  
Article
Molecular Decoration and Unconventional Double Bond Migration in Irumamycin Biosynthesis
by Vera A. Alferova, Anna A. Baranova, Olga A. Belozerova, Evgeny L. Gulyak, Andrey A. Mikhaylov, Yaroslav V. Solovev, Mikhail Y. Zhitlov, Arseniy A. Sinichich, Anton P. Tyurin, Ekaterina A. Trusova, Alexey V. Beletsky, Andrey V. Mardanov, Nikolai V. Ravin, Olda A. Lapchinskaya, Vladimir A. Korshun, Alexander G. Gabibov and Stanislav S. Terekhov
Antibiotics 2024, 13(12), 1167; https://doi.org/10.3390/antibiotics13121167 - 3 Dec 2024
Viewed by 563
Abstract
Irumamycin (Iru) is a complex polyketide with pronounced antifungal activity produced by a type I polyketide (PKS) synthase. Iru features a unique hemiketal ring and an epoxide group, making its biosynthesis and the structural diversity of related compounds particularly intriguing. In this study, [...] Read more.
Irumamycin (Iru) is a complex polyketide with pronounced antifungal activity produced by a type I polyketide (PKS) synthase. Iru features a unique hemiketal ring and an epoxide group, making its biosynthesis and the structural diversity of related compounds particularly intriguing. In this study, we performed a detailed analysis of the iru biosynthetic gene cluster (BGC) to uncover the mechanisms underlying Iru formation. We examined the iru PKS, including the domain architecture of individual modules and the overall spatial structure of the PKS, and uncovered discrepancies in substrate specificity and iterative chain elongation. Two potential pathways for the formation of the hemiketal ring, involving either an olefin shift or electrocyclization, were proposed and assessed using 18O-labeling experiments and reaction activation energy calculations. Based on our findings, the hemiketal ring is likely formed by PKS-assisted double bond migration and TE domain-mediated cyclization. Furthermore, putative tailoring enzymes mediating epoxide formation specific to Iru were identified. The revealed Iru biosynthetic machinery provides insight into the complex enzymatic processes involved in Iru production, including macrocycle sculpting and decoration. These mechanistic details open new avenues for a targeted architecture of novel macrolide analogs through synthetic biology and biosynthetic engineering. Full article
Show Figures

Figure 1

Figure 1
<p>Structures of some venturicidin-type compounds.</p>
Full article ">Figure 2
<p>Proposed scheme of Iru biosynthesis. (<b>A</b>) Comparison of <span class="html-italic">ven</span> [<a href="#B38-antibiotics-13-01167" class="html-bibr">38</a>] and <span class="html-italic">iru</span> (this work) BGCs. Homologous proteins are marked with colors with similarity (%) indicated on the labels. (<b>B</b>) The modular organization (module domains are highlighted with the same color code) of the <span class="html-italic">iru</span> PKS with the proposed scheme of backbone biosynthesis and tailoring steps.</p>
Full article ">Figure 3
<p>Front (<b>A</b>) and back (<b>B</b>) sides of the IruF/IruE complex. PKS domains in the AlphaFold3 model are colored green (KS), red (AT), yellow (KR), orange (DH), blue (T), and pink (TE). Domain numeric attributes indicate PKS modules. Arrows indicate the proposed pathway of the growing chain. Roman numerals indicate chain transfer steps.</p>
Full article ">Figure 4
<p>Comparison of the absolute configurations in the Iru backbone, derived from the KR domain types (blue) and previously established by NMR [<a href="#B28-antibiotics-13-01167" class="html-bibr">28</a>].</p>
Full article ">Figure 5
<p>Plausible mechanisms of hemiketal ring formation. Schematic representation of path 1 (<b>A</b>) cyclization through PKS-assisted double-bond shift and path 2 (<b>B</b>) through cycloaddition. The fragment of Iru backbone involved in hemiketal ring formation is highlighted with red. (<b>C</b>) Fragment of <span class="html-italic">iru</span> PKS DH domain alignment, presumably inactive domains are marked with asterisks. Key conserved motifs are highlighted with color.</p>
Full article ">Figure 6
<p>Free energy profiles for Path 1 (<b>A</b>) and Path 2 (<b>B</b>). Energy differences are given in kcal/mol.</p>
Full article ">Figure 7
<p>(<b>A</b>) MS/MS fragmentation pathway of <sup>18</sup>O-labeled Iru. Positive-mode mass spectrum of Iru from unlabeled media (<b>B</b>), H<sub>2</sub><sup>18</sup>O-enriched media (<b>C</b>). Positive-mode mass spectrum of a standard of Iru incubated in H<sub>2</sub><sup>18</sup>O for 24 h at pH 6.9 (<b>D</b>).</p>
Full article ">
17 pages, 2174 KiB  
Article
Lactate Dehydrogenase-B Oxidation and Inhibition by Singlet Oxygen and Hypochlorous Acid
by Lisa M. Landino and Emily E. Lessard
Oxygen 2024, 4(4), 432-448; https://doi.org/10.3390/oxygen4040027 - 24 Nov 2024
Viewed by 906
Abstract
Alterations in cellular energy metabolism are a hallmark of cancer and lactate dehydrogenase (LDH) enzymes are overexpressed in many cancers regardless of sufficient oxygen and functional mitochondria. Further, L-lactate plays signaling roles in multiple cell types. We evaluated the effect of singlet oxygen [...] Read more.
Alterations in cellular energy metabolism are a hallmark of cancer and lactate dehydrogenase (LDH) enzymes are overexpressed in many cancers regardless of sufficient oxygen and functional mitochondria. Further, L-lactate plays signaling roles in multiple cell types. We evaluated the effect of singlet oxygen and hypochlorous acid (HOCl) on pig heart LDH-B, which shares 97% homology with human LDH-B. Singlet oxygen was generated photochemically using methylene blue or the chlorophyll metabolites, pheophorbide A and chlorin e6. Singlet oxygen induced protein crosslinks observed by SDS-PAGE under reducing conditions and inhibited LDH-B activity. Ascorbate, hydrocaffeic acid, glutathione and sodium azide were employed as singlet oxygen scavengers and shown to protect LDH-B. Using fluorescein-modified maleimide, no changes in cysteine availability as a result of singlet oxygen damage were observed. This was in contrast to HOCl, which induced the formation of disulfides between LDH-B subunits, thereby decreasing LDH-B labeling with fluorescein. HOCl oxidation inhibited LDH-B activity; however, disulfide reduction did not restore it. LDH-B cysteines were resistant to millimolar H2O2, chloramines and Angeli’s salt. In the absence of pyruvate, LDH-B enhanced NADH oxidation in a chain reaction initiated by singlet oxygen that resulted in H2O2 formation. Once damaged by either singlet oxygen or HOCl, NADH oxidation by LDH-B was impaired. Full article
Show Figures

Figure 1

Figure 1
<p><sup>1</sup>O<sub>2</sub> oxidation of LDH-B. LDH-B (10 µM) samples containing 1 µM MB were irradiated for 0–90 s. (<b>A</b>) Samples were analyzed by SDS-PAGE under reducing conditions on 10% gels and stained with Coomassie blue. (<b>B</b>) LDH-B samples were prepared as in (<b>A</b>). After irradiation, samples were diluted 1:10 with 20 mM Tris pH 8.8. Kinetic assays (200 mL) contained 5 µL 1:10 LDH-B, 4 mM lithium lactate and 2 mM NAD<sup>+</sup> in 20 mM Tris pH 8.8. Data shown is the average of at least three independent experiments performed in duplicate. NAD<sup>+</sup> reduction was monitored at 340 nm in a 96-well plate at 30 °C for 4 min. (<b>C</b>) LDH-B samples prepared as in A were analyzed by native gel electrophoresis on 0.8% agarose gels. Activity stain contained 0.75 mM NAD<sup>+</sup>, 25 mM lithium lactate, 8–10 mg NBT, and 1–2 mg PMS in 100 mM Tris pH 8.6.</p>
Full article ">Figure 2
<p>Cysteine labeling of LDH-B with fluorescein. LDH-B (10 μM) samples containing 1 μM MB were irradiated for 0–120 s. Samples were treated with 500 μM M5F for 30 min at 37 °C and analyzed by SDS-PAGE under reducing conditions on 10% gels. Fluorescent images were captured using a Bio-Rad Chemi-doc XRS imaging system (<b>left</b>). After imaging, gels were stained with Coomassie blue (<b>right</b>).</p>
Full article ">Figure 3
<p>LDH-B protection from <sup>1</sup>O<sub>2</sub> by ascorbate and HCA. (<b>A</b>) LDH-B (10 μM) samples containing 1 μM MB were irradiated for 120 s. Ascorbate (100 or 250 μM) was added prior to irradiation. Samples were analyzed by SDS-PAGE under reducing conditions on 10% gels and stained with Coomassie blue. (<b>B</b>) LDH-B (10 μM) samples containing 1 μM MB or 15 μM pheoA were irradiated for 120 s. Ascorbate (250 or 500 μM) was added prior to irradiation. “D” indicates dark. Samples were analyzed by native gel electrophoresis on 0.8% agarose gels. Activity stain contained 0.75 mM NAD<sup>+</sup>, 25 mM lithium lactate, 8–10 mg NBT, and 1–2 mg PMS in 100 mM Tris pH 8.6. (<b>C</b>) LDH-B samples contained 1 μM MB, 2.5 μM chlorin e6, or 15 μM pheoA. Either 0.5 mM ascorbate or 1.5 mM HCA was added prior to irradiation. LDH-B activity was measured as in <a href="#oxygen-04-00027-f001" class="html-fig">Figure 1</a>B.</p>
Full article ">Figure 4
<p>LDH oxidation and inhibition by HOCl. LDH-B (10 μM) in 10 mM PB pH 7.4 was treated with up to 150 μM HOCl for 30 min at 30 °C. (<b>A</b>) Samples were analyzed by native gel electrophoresis on 0.8% agarose gels. Activity stain contained 0.75 mM NAD<sup>+</sup>, 25 mM lithium lactate, 8–10 mg NBT, and 1–2 mg PMS in 100 mM Tris pH 8.6. (<b>B</b>) Samples were analyzed by SDS-PAGE under reducing and nonreducing conditions (±βME). The gel was stained with Coomassie Blue. (<b>C</b>) Kinetic assays (200 μL) contained 5 μL 1:10 LDH-B, 4 mM lithium lactate, and 2 mM NAD<sup>+</sup> in 20 mM Tris pH 8.8. Data shown is the average of at least three independent experiments performed in duplicate. NAD<sup>+</sup> reduction was monitored at 340 nm in a 96-well plate at 30 °C for 4 min.</p>
Full article ">Figure 5
<p>Cysteine labeling after HOCl oxidation. LDH-B (10 μM) in 10 mM PB pH 7.4 was treated with up to 150 μM HOCl for 30 min at 30 °C. Excess HOCl was quenched with 0.2 mM S-methyl-cys. (<b>A</b>) Samples were treated with 500 μM M5F for 30 min at 37 °C and analyzed by SDS-PAGE under nonreducing conditions on 10% gels. Fluorescent images were captured using a Bio-Rad Chemi-doc XRS imaging system (<b>left</b>). After imaging, gels were stained with Coomassie blue (<b>right</b>). (<b>B</b>) Samples were prepared as in (<b>A</b>) except reactions were 60 μL. M5F-labeled LDH-B was precipitated with 80% ethanol. Fluorescein was quantitated at 495 nm relative to a fluorescein standard curve. These data are the average of two independent trials performed in duplicate.</p>
Full article ">Figure 6
<p>Enhanced NADH oxidation by LDH-A and LDH-B: activation by <sup>1</sup>O<sub>2</sub>. ll reactions (100 μL) contained 200 μM NADH in 50 mM PB pH 7.1 in a 96-well plate. <sup>1</sup>O<sub>2</sub> formation was initiated with 2.5 μM MB and 10 s of red light. Additions included LDH-A or LDH-B (10 μM). Absorbance at 340 nm was measured prior to and immediately after light exposure. These data represent the average of two independent experiments performed in triplicate.</p>
Full article ">
20 pages, 13573 KiB  
Article
Sparstolonin B Reduces Estrogen-Dependent Proliferation in Cancer Cells: Possible Role of Ceramide and PI3K/AKT/mTOR Inhibition
by Yağmur Dilber, Hanife Tuğçe Çeker, Aleyna Öztüzün, Bürke Çırçırlı, Esma Kırımlıoğlu, Zerrin Barut and Mutay Aslan
Pharmaceuticals 2024, 17(12), 1564; https://doi.org/10.3390/ph17121564 - 21 Nov 2024
Viewed by 419
Abstract
Background: The aim of this study was to determine the effect of Sparstolonin B (SsnB) on cell proliferation and apoptosis in human breast cancer (MCF-7) and human ovarian epithelial cancer (OVCAR-3) cell lines in the presence and absence of estradiol hemihydrate (ES). Phosphoinositol-3 [...] Read more.
Background: The aim of this study was to determine the effect of Sparstolonin B (SsnB) on cell proliferation and apoptosis in human breast cancer (MCF-7) and human ovarian epithelial cancer (OVCAR-3) cell lines in the presence and absence of estradiol hemihydrate (ES). Phosphoinositol-3 kinase (PI3K), phosphorylated protein kinase B alpha (p-AKT), phosphorylated mTOR (mechanistic target of rapamycin) signaling proteins, and sphingomyelin/ceramide metabolites were also measured within the scope of the study. Methods: The anti-proliferative effects of SsnB therapy were evaluated over a range of times and concentrations. Cell proliferation was determined by measuring the Proliferating Cell Nuclear Antigen (PCNA). PCNA was quantified by ELISA and cell distribution was assessed by immunofluorescence microscopy. MTT analysis was used to test the vitality of the cells, while LC-MS/MS was used to analyze the amounts of ceramides (CERs), sphingosine-1-phosphate (S1P), and sphingomyelins (SMs). TUNEL labeling was used to assess apoptosis, while immunofluorescence staining and enzyme-linked immunosorbent assay (ELISA) were used to measure the levels of PI3K, p-AKT, and p-mTOR proteins. Results: Sparstolonin B administration significantly decreased cell viability in MCF-7 and OVCAR-3 cells both in the presence and absence of ES, while it did not cause toxicity in healthy human fibroblasts. In comparison to controls, cancer cells treated with SsnB showed a significant drop in the levels of S1P, PI3K, p-AKT, and p-mTOR. In cancer cells cultured with SsnB, a significant increase in intracellular concentrations of C16-C24 CERs and apoptosis was observed. Conclusions: SsnB downregulated the levels of S1P, PI3K, p-AKT, and p-mTOR while reducing cell proliferation and promoting ceramide buildup and apoptosis. Full article
(This article belongs to the Section Pharmacology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>A</b>) Evaluation of cell viability by MTT analysis for 16–48 h in MCF-7 cells. Cells treated with estradiol hemihydrate (ES, 1, 10, and 100 nM) and DMSO, dimethyl sulfoxide (1 μL/mL). Data are representative of 7–8 separate measurements and values are given as mean ± SD. Statistical analysis was performed by two-way ANOVA and differences between groups were determined by Tukey’s multiple comparison test. *, <span class="html-italic">p</span> &lt; 0.05, when compared with the control group at the same time periods. #, <span class="html-italic">p</span> &lt; 0.05, when compared with all other groups at the same time period. §, <span class="html-italic">p</span> &lt; 0.05, when compared with 16 and 24 h within the same dose. (<b>B</b>) Evaluation of cell viability by MTT analysis for 16–48 h in OVCAR-3 cells. Data are representative of 6–8 separate measurements and values are given as mean ± SD. Statistical analysis was performed by two-way ANOVA and differences between groups were determined by Tukey’s multiple comparison test. *, <span class="html-italic">p</span> &lt; 0.05, when compared with the control and DMSO group at the same time period. #, <span class="html-italic">p</span> &lt; 0.05, when compared with the control group at the same time period. §, <span class="html-italic">p</span> &lt; 0.05, when compared with 24 and 48 h within the same dose. (<b>C</b>) Evaluation of cell viability in BJ cells by MTT analysis for 16–48 h. Data are representative of 6 separate measurements and values are given as mean ± SD. Statistical analysis was performed by two-way ANOVA. Differences between groups were determined by Tukey’s multiple comparison test. *, <span class="html-italic">p</span> &lt; 0.05, when compared with all other groups at the same time periods. §, <span class="html-italic">p</span> &lt; 0.05, when compared with 16 h within the same dose. (<b>D</b>) Light microscope image (10× magnification) of ES-applied MCF-7, OVCAR-3, and BJ cells after 48 h. While no change was observed in the control and DMSO (1 μL/mL) groups, significant proliferation was observed in MCF-7 and OVCAR-3 cells compared to the control as a result of 10 and 100 nM ES applications. It was observed that 100 nM ES application for 48 h caused deterioration in morphology, shrinkage, clustering, and toxicity in BJ cells.</p>
Full article ">Figure 2
<p>(<b>A</b>) The effect of Sparstolonin B(SsnB) on MCF-7 cell viability. Cell viability analysis was performed for 16–48 h. Cells treated with DMSO, dimethyl sulfoxide (1 μL/mL), and cells treated with SsnB. Data are representative of 6–8 separate experiments and values are given as mean ± SD. Statistical analysis was performed by two-way ANOVA and differences between groups were determined by Tukey’s multiple comparison test. *, <span class="html-italic">p</span> &lt; 0.05, compared to control, DMSO, 3.125–12.5 µM groups within the same time periods. #, <span class="html-italic">p</span> &lt; 0.05, compared to 25 µM group within the same time periods. §, <span class="html-italic">p</span> &lt; 0.05 when compared to 24 and 48 h within the same dose. (<b>B</b>) The effect of SsnB on OVCAR-3 cell viability. Data are representative of 7 separate experiments and values are given as mean ± SD. Statistical analysis was performed by two-way ANOVA and differences between groups were determined by Tukey’s multiple comparison test. *, <span class="html-italic">p</span> &lt; 0.001, vs. DMSO and control in all incubation periods. #, <span class="html-italic">p</span> &lt; 0.001, compared to 3.125–25 µM groups within the same time periods. ¶, <span class="html-italic">p</span> &lt; 0.05 vs. 3.125 and 6.25 µM groups within the same period. (<b>C</b>) Effect of SsnB on BJ cell viability. Data are representative of 7–8 separate experiments and values are given as mean ± SD. Statistical analysis was performed by two-way ANOVA and differences between groups were determined by Tukey’s multiple comparison test. *, <span class="html-italic">p</span> &lt; 0.05, vs. all groups in the same time periods. (<b>D</b>) Effect of 24 h SsnB treatment on cell viability in MCF-7 cells during 48 h ES proliferation. Data are representative of 8 separate experiments and values are given as mean ± SD. Cells treated with DMSO (1 μL/mL), estradiol hemihydrate (ES, 10 nM), and SsnB (25 μM). Incubation with SsnB started 24 h after ES treatment and was continued for 24 h. Statistical analysis was performed by one-way ANOVA and differences between groups were determined by Holm–Sidak’s multiple comparison test. *, <span class="html-italic">p</span> &lt; 0.0001, when compared to all groups. #, <span class="html-italic">p</span> &lt; 0.0001, compared with control, DMSO, and ES 10 nM groups. (<b>E</b>) Effect of 24 h SsnB treatment on cell viability in OVCAR-3 cells during 48 h ES proliferation. Data are representative of 8 separate experiments and values are given as mean ± SD. Statistical analysis was performed by one-way ANOVA and differences between groups were determined by Holm–Sidak multiple comparison test. *, <span class="html-italic">p</span> &lt; 0.05, compared with all groups. #, <span class="html-italic">p</span> &lt; 0.05, compared with control, DMSO, and ES 10 nM groups. (<b>F</b>) Light microscope image (10× magnification) of MCF-7 and OVCAR-3 cells after 48 h of ES (10 nM) application. Incubation with SsnB started 24 h after ES treatment and was continued for 24 h. Significant proliferation was observed in MCF-7 and OVCAR-3 cells compared to the control as a result of 10 nM ES application. SsnB application significantly decreased cell proliferation and caused significant changes in cell morphology in MCF-7 and OVCAR-3 cells compared to the control and ES groups. ES + SP application was observed to cause deterioration in morphology, shrinkage, clustering, and toxicity in MCF-7 and OVCAR-3 cells.</p>
Full article ">Figure 3
<p>(<b>A</b>) Representative immunofluorescent staining of proliferating cell nuclear antigen (PCNA) in MCF-7 and OVCAR-3 cells treated with either DMSO (1 μL/mL), ES (10 nM), or SsnB (25 μM). 10× magnification. Incubation with SsnB started 24 h after ES treatment and was continued for 24 h in the ES + SsnB group. (<b>B</b>) Quantitation of PCNA fluorescence staining in MCF-7 cells by ImageJ software (version 1.53k). Data shown are representative of 10 separate measurements and values are given as mean ± SD. Statistical analysis was performed by one-way ANOVA and differences between groups were determined by Tukey’s multiple comparisons analysis. *, <span class="html-italic">p</span> &lt; 0.0001, vs. control and DMSO groups. #, <span class="html-italic">p</span> &lt; 0.005, vs. control, DMSO, and ES 10 nM groups. ¶, <span class="html-italic">p</span> &lt;0.001 vs. ES + SsnB group. (<b>C</b>) Quantitation of PCNA fluorescence staining in OVCAR-3 cells by ImageJ software. Data shown are representative of 10 separate measurements and values are given as mean ± SD. Statistical analysis was performed by one-way ANOVA and differences between groups were determined by Tukey’s multiple comparisons analysis. *, <span class="html-italic">p</span> &lt; 0.0001, vs. control and DMSO groups. #, <span class="html-italic">p</span> &lt; 0.05, vs. control, DMSO, and ES 10 nM groups. ¶, <span class="html-italic">p</span> &lt;0.001 vs. ES + SsnB group. (<b>D</b>) PCNA protein levels in MCF-7 cells. Data shown are representative of 10 separate measurements and values are given as mean ± SD. Statistical analysis was performed by one-way ANOVA and differences between groups were determined by Tukey’s multiple comparisons analysis. *, <span class="html-italic">p</span> &lt; 0.0001, vs. control and DMSO groups. #, <span class="html-italic">p</span> &lt; 0.0001, vs. control, DMSO, and ES 10 nM groups. (<b>E</b>) PCNA protein levels in OVCAR-3 cells. Data shown are representative of 10 separate measurements and values are given as mean ± SD. Statistical analysis was performed by one-way ANOVA and differences between groups were determined by Tukey’s multiple comparisons analysis. *, <span class="html-italic">p</span> &lt; 0.0001, vs. control and DMSO groups. #, <span class="html-italic">p</span> &lt; 0.0001, vs. control, DMSO, and ES 10 nM groups. ¶, <span class="html-italic">p</span> &lt; 0.05 vs. ES + SsnB group. (<b>F</b>) Representative immunofluorescent staining of TUNEL staining in MCF-7 and OVCAR-3 cells. 40× objective lens was used to obtain double-labeled images. (<b>G</b>) Quantitation of TUNEL staining in MCF-7 cells with the ImageJ program. Values mean ± SD (<span class="html-italic">n</span> = 10). Statistical analysis was performed by one-way ANOVA and differences between groups were determined by Tukey’s multiple comparisons analysis. *, <span class="html-italic">p</span> &lt; 0.001, vs. control, DMSO, and ES 10 nM groups. #, <span class="html-italic">p</span> &lt; 0.001, vs. ES + SsnB. (<b>H</b>) Quantitation of TUNEL staining in OVCAR-3 cells with ImageJ program. Values are mean ± SD (<span class="html-italic">n</span> = 10). One-way ANOVA and Tukey’s multiple comparisons were used to determine statistical significance. *, <span class="html-italic">p</span> &lt; 0.0001, vs. control, DMSO, and ES 10 nM groups. #, <span class="html-italic">p</span> &lt; 0.05, vs. ES + SsnB.</p>
Full article ">Figure 4
<p>(<b>A</b>) Representative immunofluorescent staining of phosphatidylinositol 3-kinase (PI3K), phospho (Ser473) protein kinase AKT (p-AKT), and phospho (Ser2448) mammalian target of rapamycin (p-mTOR) in MCF-7 cells treated with either DMSO (1 μL/mL), ES (10 nM), or SsnB (25 μM). 10× magnification. Incubation with SsnB started 24 h after ES treatment and was continued for 24 h in the ES + SsnB group. (<b>B</b>) Quantitation of PI3K fluorescence staining in MCF-7 cells by ImageJ software. Data shown are representative of 9–10 separate measurements and values are given as mean ± SD. Statistical analysis was performed by one-way ANOVA and differences between groups were determined by Tukey’s multiple comparisons analysis. *, <span class="html-italic">p</span> &lt; 0.0001, vs. all groups. #, <span class="html-italic">p</span> &lt; 0.05, vs. all groups. (<b>C</b>) Quantitation of p-AKT fluorescence staining in MCF-7 cells by ImageJ software. Data shown are representative of 9–10 separate measurements and values are given as mean ± SD. Statistical analysis was performed by one-way ANOVA and differences between groups were determined by Tukey’s multiple comparisons analysis. *, <span class="html-italic">p</span> &lt; 0.0001, vs. all groups. **, <span class="html-italic">p</span>&lt; 0.05 vs. all groups. #, <span class="html-italic">p</span> &lt; 0.05, vs. all groups. (<b>D</b>) Quantitation of p-mTOR fluorescence staining in MCF-7 cells by ImageJ software. Data shown are representative of 10 separate measurements and values are given as mean ± SD. Statistical analysis was performed by one-way ANOVA and differences between groups were determined by Tukey’s multiple comparisons analysis. *, <span class="html-italic">p</span> &lt; 0.0001, vs. all groups. #, <span class="html-italic">p</span> &lt; 0.001, vs. all groups. (<b>E</b>) PI3K protein levels in MCF-7 cells. Data shown are representative of 7 separate measurements and values are given as mean ± SD. Statistical analysis was performed by one-way ANOVA and differences between groups were determined by Tukey’s multiple comparisons analysis. *, <span class="html-italic">p</span> &lt; 0.05, vs. all groups. #, <span class="html-italic">p</span> &lt; 0.05, vs. control and DMSO. (<b>F</b>) p-AKT protein levels in MCF-7 cells. Data shown are representative of 7 separate measurements and values are given as mean ± SD. Statistical analysis was performed by one-way ANOVA and differences between groups were determined by Tukey’s multiple comparisons analysis. *, <span class="html-italic">p</span> &lt; 0.05, vs. all groups. #, <span class="html-italic">p</span> &lt;0.001 vs. control, DMSO, and ES 10 nM. (<b>G</b>) p-mTOR protein levels in MCF-7 cells. Data shown are representative of 7 separate measurements and values are given as mean ± SD. Statistical analysis was performed by one-way ANOVA and differences between groups were determined by Tukey’s multiple comparisons analysis. *, <span class="html-italic">p</span> &lt; 0.05, vs. all groups.</p>
Full article ">Figure 5
<p>(<b>A</b>) Representative immunofluorescent staining of phosphatidylinositol 3-kinase (PI3K), phospho (Ser473) protein kinase AKT (p-AKT), and phospho (Ser2448) mammalian target of rapamycin (p-mTOR) in OVCAR-3 cells treated with either DMSO (1 μL/mL), ES (10 nM), or SsnB (25 μM). 10× magnification. Incubation with SsnB started 24 h after ES treatment and was continued for 24 h in the ES + SsnB group. (<b>B</b>) Quantitation of PI3K fluorescence staining in OVCAR-3 cells by ImageJ software. Data shown are representative of 10 separate measurements and values are given as mean ± SD. Statistical analysis was performed by one-way ANOVA and differences between groups were determined by Tukey’s multiple comparisons analysis. *, <span class="html-italic">p</span> &lt; 0.001, vs. all groups. **, <span class="html-italic">p</span> &lt; 0.05, vs. control and DMSO. #, <span class="html-italic">p</span> &lt; 0.05, vs. all groups. (<b>C</b>) Quantitation of p-AKT fluorescence staining in OVCAR-3 cells by ImageJ software. Data shown are representative of 10 separate measurements and values are given as mean ± SD. Statistical analysis was performed by one-way ANOVA and differences between groups were determined by Tukey’s multiple comparisons analysis. *, <span class="html-italic">p</span> &lt; 0.0001, vs. all groups. **, <span class="html-italic">p</span> &lt; 0.05 vs. control and DMSO. #, <span class="html-italic">p</span> &lt; 0.001, vs. all groups. (<b>D</b>) Quantitation of p-mTOR fluorescence staining in OVCAR-3 cells by ImageJ software. Data shown are representative of 10 separate measurements and values are given as mean ± SD. Statistical analysis was performed by one-way ANOVA and differences between groups were determined by Tukey’s multiple comparisons analysis. *, <span class="html-italic">p</span> &lt; 0.0001, vs. all groups. **, <span class="html-italic">p</span> &lt; 0.05, vs. control and DMSO #, <span class="html-italic">p</span> &lt; 0.05, vs. all groups. (<b>E</b>) PI3K protein levels in OVCAR-3 cells. Data shown are representative of 7 separate measurements and values are given as mean ± SD. Statistical analysis was performed by one-way ANOVA and differences between groups were determined by Tukey’s multiple comparisons analysis. *, <span class="html-italic">p</span> &lt; 0.05, vs. all groups. #, <span class="html-italic">p</span> &lt; 0.05, vs. control and DMSO. (<b>F</b>) p-AKT protein levels in OVCAR-3 cells. Data shown are representative of 7 separate measurements and values are given as mean ± SD. Statistical analysis was performed by one-way ANOVA and differences between groups were determined by Tukey’s multiple comparisons analysis. *, <span class="html-italic">p</span> &lt; 0.0001, vs. all groups. **, <span class="html-italic">p</span> &lt; 0.05 vs. all groups. #, <span class="html-italic">p</span> &lt; 0.001 vs. control and DMSO. (<b>G</b>) p-mTOR protein levels in OVCAR-3 cells. Data shown are representative of 7 separate measurements and values are given as mean ± SD. Statistical analysis was performed by one-way ANOVA and differences between groups were determined by Tukey’s multiple comparisons analysis. *, <span class="html-italic">p</span> &lt; 0.001, vs. all groups. **, <span class="html-italic">p</span> &lt; 0.05 vs. all groups. #, <span class="html-italic">p</span> &lt; 0.001 vs. control and DMSO.</p>
Full article ">Scheme 1
<p>Structure of Sparstolonin B [<a href="#B7-pharmaceuticals-17-01564" class="html-bibr">7</a>].</p>
Full article ">
15 pages, 2878 KiB  
Article
Validation and Optimization of a Stable Isotope-Labeled Substrate Assay for Measuring AGAT Activity
by Alex Lee, Lucas Anderson, Ilona Tkachyova, Michael B. Tropak, Dahai Wang and Andreas Schulze
Int. J. Mol. Sci. 2024, 25(23), 12490; https://doi.org/10.3390/ijms252312490 - 21 Nov 2024
Viewed by 488
Abstract
L-arginine: glycine amidinotransferase (AGAT) gained academic interest as the rate-limiting enzyme in creatine biosynthesis and its role in the regulation of creatine homeostasis. Of clinical relevance is the diagnosis of patients with AGAT deficiency but also the potential role of AGAT as therapeutic [...] Read more.
L-arginine: glycine amidinotransferase (AGAT) gained academic interest as the rate-limiting enzyme in creatine biosynthesis and its role in the regulation of creatine homeostasis. Of clinical relevance is the diagnosis of patients with AGAT deficiency but also the potential role of AGAT as therapeutic target for the treatment of another creatine deficiency syndrome, guanidinoacetate N-methyltransferase (GAMT) deficiency. Applying a stable isotope-labeled substrate method, we utilized ARG 15N2 (ARG-δ2) and GLY 13C215N (GLY-δ3) to determine the rate of 1,2-13C2,15N3 guanidinoacetate (GAA-δ5) formation to assess AGAT activity in various mouse tissue samples and human-derived cells. Following modification and optimization of the assay, we analyzed AGAT activity in several mouse organs. The Km and Vmax of AGAT in mouse kidney for GLY-δ3 were 2.06 mM and 6.48 ± 0.26 pmol/min/mg kidney, and those for ARG-δ2, they were 2.67 mM and 2.17 ± 0.49 pmol/min/mg kidney, respectively. Our results showed that mouse kidneys had the highest levels of enzymatic activity, followed by brain and liver, with 4.6, 1.8, and 0.4 pmol/min/mg tissue, respectively. Both the heart and muscle had no detectable levels of AGAT activity. We noted that due to interference with arginase in the liver, performing the enzyme assay in liver homogenates required the addition of Nor-NOHA, an arginase inhibitor. In immortalized human cell lines, we found the highest levels of AGAT activity in RH30 cells, followed by HepaRG, HAP1, and HeLa cells. AGAT activity was readily detectable in lymphoblasts and leukocytes from healthy controls. In our assay, AGAT activity was not detectable in HEK293 cells, in human fibroblasts, and in the lymphoblasts of a patient with AGAT deficiency. Our results demonstrate that this enzyme assay is capable of accurately quantifying AGAT activity from both tissues and cells for diagnostic purposes and research. Full article
(This article belongs to the Section Molecular Endocrinology and Metabolism)
Show Figures

Figure 1

Figure 1
<p>Schematic of the synthesis of ornithine and guanidinoacetate from GLY-δ3 and ARG-δ2. Green numbers indicate the locations of stable isotopes.</p>
Full article ">Figure 2
<p>GLY-δ3 and ARG-δ2 allow for the formation of GAA-δ5. Combinations of either unlabeled (ARG + GLY), one labeled (ARG-δ2 + GLY and ARG + GLY-δ3), or both labeled substrates (ARG-δ2 + GLY-δ3) at a final concentration of 30 mM were used for the AGAT enzyme reaction in mouse kidney lysates. GAA is present in kidney tissue endogenously, and its formation is observed in all aforementioned reactions. However, the formation of GAA-δ5 only occurs when both labeled substrates are present. Top panel shows the appearance of GAA peak (+MRM, transition 174.2/101) and lower panel GAA-δ5 (+MRM, transition 179/105).</p>
Full article ">Figure 3
<p>The formation of GAA-δ5 requires both labeled substrates. (<b>A</b>) Mouse kidneys were homogenized in potassium phosphate buffer, pH 7.4 and were incubated with different combinations of unlabeled and labeled substrates at a concentration of 7.5 mM in conjunction with 15 mM ornithine, and the concentrations of both GAA and GAA-δ5 were measured in positive-ionization MRM with the following transitions GAA 174.2/101, GAA-δ5 179/105, ornithine 189.2/70.1). The formation of GAA occurred only when both unlabeled substrates were present, while the synthesis of GAA-δ5 was exclusive to having both labeled substrates. In both reactions, the addition of ornithine resulted in a significant reduction of GAA and GAA-δ5 being produced. (<b>B</b>) Across all the different substrate combinations, the concentration of ornithine produced was relatively similar. (<b>C</b>) Dose–response curve showing the effect of ornithine on AGAT activity. Based on the concentration of ornithine produced in the enzyme reaction (<b>B</b>), it does not seem to have an inhibitory effect on AGAT activity. 0.0001 = 0 mM ornithine concentration.</p>
Full article ">Figure 4
<p>Optimization of reaction conditions for enzyme assay to measure AGAT activity. (<b>A</b>) Temperature. Mouse kidney samples were homogenized in potassium phosphate buffer, pH 7.4. Th reaction was performed using 1.8 mM of both labeled substrates, GLY-δ3 and ARG-δ2, for a duration of 1 h using the following temperatures: 4 °C, 22 °C, 37 °C, 42 °C, 55 °C, and 75 °C. GAA-δ5 was quantified to determine AGAT activity and was normalized to the amount of kidney used in the reaction. The data were fit to a Log-normal curve to visualize the correlation between temperature and AGAT activity. Observations showed peak activity to be approximately at 40 °C. to visualize the correlation between temperature and AGAT activity. Observations showed peak activity to be approximately at 40 °C. (<b>B</b>) pH. Mouse kidney samples were homogenized in water. Th reaction was performed using 1.8 mM of both labeled substrates, GLY-δ3 and ARG-δ2, for 1 h in potassium phosphate buffer under the following pH conditions: 6, 6.5, 7, 7.4, 8, 9, 10. AGAT activity was determined by normalizing GAA-δ5 concentrations to the amount of kidney used in the reaction. A Gaussian curve identified peak activity to occur when the pH of the environment was 7.6. (<b>C</b>) Time. Mouse kidney samples were homogenized in potassium phosphate buffer of pH 7.4. The reaction was performed using 1.8 mM of both labeled substrates, GLY-δ3and ARG-δ2, and incubated at 37 °C for durations of 0 min, 30 min, 60 min, 90 min, or 120 min. The concentration of GAA-δ5 was quantified and plotted against reaction duration. Results show a linear relationship between the duration of the enzyme assay and the amount of GAA-δ5 produced.</p>
Full article ">Figure 5
<p>Determining enzyme kinetics of AGAT in mouse kidney. Mouse kidneys were homogenized in potassium phosphate buffer, pH 7.4. and homogenates were incubated with various concentrations of one labeled substrate while the other substrate was held at a constant concentration of 1.9 mM. The reaction was performed at 37 °C for 1 h. The concentration of GAA-δ5 was determined and normalized to the amount of kidney used in the reaction. Graphing the data and fitting them to a Michaelis–Menten curve shows that the K<sub>m</sub> of AGAT for GLY-δ3 is 2.06 mM with a V<sub>max</sub> of 6.48 ± 0.26. In comparison, the K<sub>m</sub> of AGAT for ARG-δ2 is 2.67 mM with a V<sub>max</sub> of 2.17 ± 0.49.</p>
Full article ">Figure 6
<p>High levels of arginase in the liver can be inhibited through the use of nor-NOHA. Mouse tissues were homogenized in potassium phosphate buffer of pH 7.4 and were incubated with 7.5 mM of both GLY-δ3 and ARG-δ2 for 1 h at 37 °C. The concentrations of (<b>A</b>) GLY-δ3, (<b>B</b>) ARG-δ2, and (<b>C</b>) ornithine were compared across the various tissues. In the liver, the levels of ARG-δ2 were significantly lower, while the levels of ornithine were much higher when compared to the organs. In contrast, the concentration of GLY-δ3 was relatively constant across all tissues. In order to determine if nor-NOHA was an effective inhibitor of arginase activity, mouse livers were homogenized in potassium phosphate buffer, pH 7.4 and were incubated with various concentrations of nor-NOHA in combination with 7.5 mM of both GLY-δ3 and ARG-δ2 for 1 h at 37 °C. As the concentration of nor-NOHA within the reaction increased, the levels of (<b>D</b>) ARG-δ2 and (<b>E</b>) ornithine increased and decreased, respectively, to levels that were consistent with the other organs. (<b>F</b>) Quantification of AGAT activity shows that higher concentrations of nor-NOHA within the reaction increase AGAT activity until it plateaus at around 0.4 pmol/min/mg tissue.</p>
Full article ">Figure 7
<p>Determining AGAT activity in mouse organs. Mouse tissues were homogenized in potassium phosphate buffer, pH 7.4 and were incubated with 7.5 mM of both GLY-δ3 and ARG-δ2 for 1 h at 37 °C. The highest levels of AGAT activity were found in the kidney at 4.6 pmol/min/mg tissue, followed by the brain and liver. AGAT activity in the heart and muscle were below the detection limit (BDL). Reactions in liver samples were supplemented with 500 µM of nor-NOHA.</p>
Full article ">Figure 8
<p>Measuring AGAT activity in cell lines and patient derived samples. Cells were sonicated in potassium phosphate buffer, pH 7.4, and lysate was incubated with 7.5 mM of both GLY-δ3 and ARG-δ2 for 1 h at 37 °C. Penta-GAA levels were normalized to protein concentrations in order to determine AGAT activity in each sample. From the immortalized cell lines, RH30 cells had the highest levels of activity followed by HepaRG, HAP1, and HeLa cells. In the patient derived cells, AGAT activity was detected in all lymphocyte cells with the exception of AGAT D-, as it derived from a patient that is deficient in AGAT. The leukocyte sample had AGAT activity that was slightly lower than the lymphocyte cells. In addition, fibroblasts were also observed to have no detectable amounts of AGAT.</p>
Full article ">
16 pages, 4595 KiB  
Article
Effects of Two Trichoderma Strains on Apple Replant Disease Suppression and Plant Growth Stimulation
by Wen Du, Pengbo Dai, Mingyi Zhang, Guangzhu Yang, Wenjing Huang, Kuijing Liang, Bo Li, Keqiang Cao, Tongle Hu, Yanan Wang, Xianglong Meng and Shutong Wang
J. Fungi 2024, 10(11), 804; https://doi.org/10.3390/jof10110804 - 20 Nov 2024
Viewed by 676
Abstract
Fusarium oxysporum, the pathogen responsible for apple replant disease (ARD), is seriously threatening the apple industry globally. We investigated the antagonistic properties of Trichoderma strains against F. oxysporum HS2, aiming to find a biological control solution to minimize the dependence on chemical [...] Read more.
Fusarium oxysporum, the pathogen responsible for apple replant disease (ARD), is seriously threatening the apple industry globally. We investigated the antagonistic properties of Trichoderma strains against F. oxysporum HS2, aiming to find a biological control solution to minimize the dependence on chemical pesticides. Two of the thirty-one Trichoderma strains assessed through plate confrontation assays, L7 (Trichoderma atroviride) and M19 (T. longibrachiatum), markedly inhibited = F. oxysporum, with inhibition rates of 86.02% and 86.72%, respectively. Applying 1 × 106 spores/mL suspensions of these strains notably increased the disease resistance in embryonic mung bean roots. Strains L7 and M19 substantially protected Malus robusta Rehd apple rootstock from ARD; the plant height, stem diameter, leaf number, chlorophyll content, and defense enzyme activity were higher in the treated plants than in the controls in both greenhouse and field trials. The results of fluorescent labeling confirmed the effective colonization of these strains of the root soil, with the number of spores stabilizing over time. At 56 days after inoculation, the M19 and L7 spore counts in various soils confirmed their persistence. These results underscore the biocontrol potential of L7 and M19 against HS2, offering valuable insights into developing sustainable ARD management practices. Full article
(This article belongs to the Section Fungal Pathogenesis and Disease Control)
Show Figures

Figure 1

Figure 1
<p>Antagonistic observation of biocontrol <span class="html-italic">Trichoderma</span> strains L7 and M19 against <span class="html-italic">F. oxysporum</span> HS2. The morphology of L7 and M19 Petri dishes is observed on the (<b>left</b>), showing a distinct light yellow antagonistic zone forming at the mycelial intersection, with <span class="html-italic">Trichoderma</span> gradually covering <span class="html-italic">F. oxysporum</span> HS2. The red boxes indicate the confrontation observation zones. In the (<b>middle</b>), microscopic observation reveals that the test strains cause twisting, collapsing, and rupturing of HS2 mycelia during the parasitism process. On the (<b>right</b>), scanning electron microscope images show <span class="html-italic">Trichoderma</span> strains coiling around the mycelia of HS2.</p>
Full article ">Figure 2
<p>Identification of the tested <span class="html-italic">Trichoderma</span> isolates. (<b>A</b>) Colony morphology and microscopic observation of the tested <span class="html-italic">Trichoderma</span> isolates. L7 has circular and velvety colonies with light green conidia and slender mycelia. The phialides are slender, and the conidia are nearly spherical or ovoid, measuring 3.0–4.5 μm and 2.5–4.0 μm. M19 exhibits light green conidia with colonies radiating outward from the center, showing high sporulation rates centrally. The mycelia are tree-like, and the oval-shaped conidia measure 2.0–3.0 μm and 2.0–6.0 μm. (<b>B</b>) Phylogenetic trees of two <span class="html-italic">Trichoderma</span> strains constructed based on ITS sequences. Phylogenetic tree constructed by the neighbor-joining method based on ITS sequences. The percentage of replicate trees in which the associated taxa clustered together in the bootstrap test (1000 replicates) is shown next to the branches. The tree is drawn to scale, with branch lengths in the same units as those of the evolutionary distances used to infer the phylogenetic tree. The evolutionary distances were computed using the Poisson correction method and are in the units of the number of amino acid substitutions per site. Based on the tree, strain M19 clustered within the <span class="html-italic">T. longibrachiatum</span> branch, while L7 clustered within the <span class="html-italic">T. atroviride</span> branch. All positions containing gaps and missing data were eliminated. Evolutionary analyses were conducted in MEGA6.</p>
Full article ">Figure 3
<p>Effect of <span class="html-italic">Trichoderma</span> on the growth of <span class="html-italic">M. robusta</span> Rehd seedlings. (<b>A</b>) Determination of growth indexes of <span class="html-italic">Trichoderma</span> on <span class="html-italic">M. robusta</span> Rehd. Labels (<b>a</b>–<b>f</b>) represent seedling height, root length, fresh weight, root fresh weight, leaf number, and chlorophyll content, respectively. Values with superscript letters a and b are significanty diferent across columns (<span class="html-italic">p</span> &lt; 0.05). Results showed significant improvements in <span class="html-italic">M. robusta</span> Rehd seedling parameters after treatment with strains M19 and L7 compared to the control (CK). (<b>B</b>) The effect of <span class="html-italic">Trichoderma</span> on the growth of <span class="html-italic">M. robusta</span> Rehd. CK represents <span class="html-italic">M. robusta</span> Rehd seedlings treated with only water, L7 represents seedlings treated with L7 spore suspension, and M19 represents seedlings treated with M19 spore suspension.</p>
Full article ">Figure 4
<p>The effect of <span class="html-italic">Trichoderma</span> on the activity of defense enzymes in the roots of <span class="html-italic">M. robusta</span> seedlings. (<b>a</b>) SOD activity, (<b>b</b>) CAT activity, (<b>c</b>) PAL activity, and (<b>d</b>) root activity. CAT activity, SOD activity, PAL activity, and root vitality were all higher in <span class="html-italic">M. robusta</span> Rehd seedlings treated with the two <span class="html-italic">Trichoderma</span> strains compared to CK. Values with superscript letters a and b are significanty diferent across columns (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Effect of <span class="html-italic">Trichoderma</span> on <span class="html-italic">M. robusta</span> Rehd seedlings in normal cropping soil (60 days). (<b>A</b>) Growth of <span class="html-italic">M. robusta</span> Rehd seedlings in normal cropping soil for 60 days. CK represents <span class="html-italic">M. robusta</span> Rehd seedlings treated with only water, L7 represents seedlings treated with L7 spore suspension, and M19 represents seedlings treated with M19 spore suspension. The treatment of <span class="html-italic">Trichoderma</span> spore suspension in normal cropping soil significantly increased seedling height and demonstrated a strong growth-promoting effect. (<b>B</b>) Determination of physiological indexes of <span class="html-italic">M. robusta</span> Rehd seedlings growing in normal cropping soil for 60 days. Labels (<b>a</b>–<b>d</b>) represent seedling height, stem diameter, chlorophyll content, and leaf number, respectively. Values with superscript letters a and b are significanty diferent across columns (<span class="html-italic">p</span> &lt; 0.05). Significant enhancements in seedling height, leaf number, chlorophyll content, and root health were noted, indicating a strong growth-promoting effect.</p>
Full article ">Figure 6
<p>Effect of <span class="html-italic">Trichoderma</span> on <span class="html-italic">M. robusta</span> Rehd seedlings in continuous cropping soil (60 days). (<b>A</b>) Growth of <span class="html-italic">M. robusta</span> Rehd plants in continuous cropping soil for 60 days. CK represents <span class="html-italic">M. robusta</span> Rehd seedlings treated with only water, L7 represents seedlings treated with L7 spore suspension, and M19 represents seedlings treated with M19 spore suspension. The treatment of <span class="html-italic">Trichoderma</span> spore suspension in continuous cropping soil significantly increased seedling height and demonstrated a strong growth-promoting effect. (<b>B</b>) Determination of physiological indexes of <span class="html-italic">M. robusta</span> Rehd seedlings growing in continuous cropping soil for 60 days. Labels (<b>a</b>–<b>d</b>) represent seedling height, stem diameter, chlorophyll content, and leaf number, respectively. Values with superscript letters a, b and c are significanty diferent across columns (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 7
<p>Fluorescence observation and colonization status of two transformants in soil suspension and <span class="html-italic">M. robusta</span> Rehd root soil. (<b>A</b>) Fluorescence observed by two transformants in soil suspension. (<b>a</b>) Represents normal cropping soil; (<b>b</b>) represents continuous cropping soil; L7, M19, and MOCK are fluorescence of L7 transformant in soil, fluorescence of M19 transformant in soil, and CK of soil. Samples were taken after root drenching treatment, diluted 100 times, and fluorescence was observed under a fluorescence microscope. (<b>B</b>) Colonization status of two transformants in the soil of <span class="html-italic">M. robusta</span> Rehd root. Over time, the spore counts of the marked strains fluctuated before stabilizing. Notably, the colonization spore count of strain M19 was higher than that of strain L7 in both soil types.</p>
Full article ">
15 pages, 1932 KiB  
Article
Oxysophocarpine Prevents the Glutamate-Induced Apoptosis of HT–22 Cells via the Nrf2/HO–1 Signaling Pathway
by Ruiying Yuan, Dan Gao, Guibing Yang, Dongzhi Zhuoma, Zhen Pu, Yangzhen Ciren, Bin Li and Jianqing Yu
Curr. Issues Mol. Biol. 2024, 46(11), 13035-13049; https://doi.org/10.3390/cimb46110777 - 16 Nov 2024
Viewed by 727
Abstract
Oxysophocarpine (OSC), a quinolizidine alkaloid, shows neuroprotective potential, though its mechanisms are unclear. The aim of the present study was to investigate the neuroprotective effects of OSC through the nuclear factor erythroid 2−related factor 2 (Nrf2)/ heme oxygenase−1 (HO–1) signaling pathway using the [...] Read more.
Oxysophocarpine (OSC), a quinolizidine alkaloid, shows neuroprotective potential, though its mechanisms are unclear. The aim of the present study was to investigate the neuroprotective effects of OSC through the nuclear factor erythroid 2−related factor 2 (Nrf2)/ heme oxygenase−1 (HO–1) signaling pathway using the HT–22 cell line. Assessments of cell viability were conducted utilizing the 3−(4,5−dimethylthiazol−2−yl)−2,5−diphenyltetrazolium bromide (MTT) assay. Assessments of oxidative stress (OS) were conducted through the quantification of reactive oxygen species (ROS). The integrity of the mitochondrial membrane potential (MMP) was scrutinized using fluorescent probe technology. Apoptosis levels were quantified using terminal deoxynucleotidyl transferase dUTP nick end labeling (TUNEL) staining. The trafficking of Nrf2 within the cell nucleus was examined through immunofluorescence analysis. Furthermore, Western blotting (WB) was applied to evaluate the expression levels of proteins implicated in apoptosis and the Nrf2/HO–1 pathway. To further probe the influence of OSC on the overexpression of antioxidant enzymes, cells were subjected to transfection with HO–1 siRNA. The results showed that OSC inhibited glutamate-induced OS, as evidenced by reduced cell viability and ROS levels. Furthermore, the apoptotic condition induced by glutamate in HT–22 cells was significantly reduced following OSC treatment. More interestingly, the Nrf2/HO–1 signaling pathway was upregulated following OSC treatment. These results suggest that OSC can exert neuroprotective effects by regulating the Nrf2/HO–1 pathway to inhibit neuronal cell apoptosis, potentially aiding in the treatment of neurodegenerative diseases. Full article
(This article belongs to the Section Molecular Pharmacology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Molecular structure of oxysophocarpine.</p>
Full article ">Figure 2
<p>The effects of OSC on cell viability in HT–22 cells were assessed. HT–22 cells were exposed to varying concentrations (1.25, 2.5, 5, 10, 20 μM) of OSC for a period of 12 h. Cell viability was determined using the MTT assay. Each bar in the graph represents the mean ± standard deviation (SD), derived from three independent experiments (<span class="html-italic">n</span> = 3). “ns” stands for “not significant”, The bars marked with ## indicate a statistically significant difference compared to the control group (<span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 3
<p>The influence of OSC on glutamate-induced cytotoxicity and ROS production in HT–22 cells was assessed. Prior to a 24 h exposure to glutamate at a concentration of 20 mM, HT–22 cells were subjected to pretreatment with a range of OSC concentrations (1.25, 2.5, 5, 10 μM). Panel (<b>A</b>) illustrates the assessment of cell viability utilizing the MTT assay, while Panel (<b>B</b>) depicts the quantification of ROS production using the DCF Fluorescence intensity. Trolox, administered at 50 μM, served as a benchmark for a positive control. The data are presented as a percentage relative to untreated cell populations, with each bar signifying the mean ± SD derived from triplicate experiments. Statistical significance is denoted as follows: <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05 and <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 in contrast to the untreated control group; * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 in contrast to the group exposed solely to 20 mM glutamate. The presence or absence of treatments is indicated by “+” and “−”, respectively. ROS, reactive oxygen species.</p>
Full article ">Figure 4
<p>The impact of OSC on the modulation of the MMP and the expression of the apoptotic proteins BCL–2/BAX in glutamate-exposed HT–22 cells was investigated. HT–22 cells were pretreated with a range of concentrations (1.25, 2.5, 5, 10 μM) of OSC, followed by a 24 h exposure to glutamate at a concentration of 20 mM. (<b>A</b>) The MMP was evaluated using JC−1 staining, which was observed under a microscope at 200× magnification. Green fluorescence indicated mitochondrial depolarization, whereas red fluorescence represented normal polarization. (<b>B</b>) The levels of BCL–2/BAX were quantified through Western blotting (WB), with the expression levels normalized against actin as a loading control. The data, represented as mean values ± SD, were derived from three independent experiments (n = 3). Statistical significance is denoted as follows: <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05 and <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 indicate significant differences compared to the untreated control; * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 represent substantial differences from the group treated with glutamate alone (20 mM). The symbols “+” and “−” represent the inclusion or exclusion of the respective treatments.</p>
Full article ">Figure 5
<p>This study investigates the impact of OSC on the apoptotic response in HT-22 cells following glutamate exposure. (<b>A</b>) The apoptotic rate in HT–22 cells, subjected to 20 mM glutamate for 24 h with a prior treatment of OSC at concentrations of 1.25, 2.5, 5, and 10 μM, was ascertained using the TUNEL staining method. Apoptotic cells were identified by green fluorescence under a 200× microscope magnification. (<b>B</b>) The levels of apoptosis-related proteins, including cleaved caspase–3, caspase–3, cleaved caspase–9, and caspase–9, were assessed via WB analysis. The expression data were normalized against actin, a constitutively expressed protein. The results are expressed as a percentage relative to the control cells, which were not treated. Each bar graph displays the mean ± SD from three independent experiments (n = 3). Statistical significance is indicated as follows: <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05 and <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 indicate significant differences from the untreated control; * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 suggest substantial differences from the group treated solely with 20 mM glutamate. The inclusion or exclusion of treatments is indicated by “+” and “−”, respectively.</p>
Full article ">Figure 6
<p>Influence of OSC on Nrf2 translocation dynamics in HT–22 cells. The study examines the effect of OSC on the subcellular distribution of Nrf2 in HT–22 cells following exposure to a concentration of 10 μM for intervals of 0.5, 1, 1.5, or 2 h. (<b>A</b>,<b>B</b>) Nrf2 protein levels in both cytosolic and nuclear compartments were ascertained by WB analysis. This approach allows for the assessment of Nrf2 translocation from the cytoplasm to the nucleus in response to OSC treatment. (<b>C</b>) The visualization and quantification of Nrf2 translocation were further accomplished using immunofluorescence microscopy, providing a qualitative representation of protein movement within the cellular context. For the normalization of protein levels, cytosolic Nrf2 was referenced against actin, while nuclear Nrf2 was calibrated against lamin B1, ensuring the accuracy of the comparative analysis. Data are depicted as the mean ± SD derived from three independent experiments (n = 3). Statistical significance is represented by the following notations: <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05 indicate significant differences when compared to the control group without treatment.</p>
Full article ">Figure 7
<p>Modulation of HO–1 protein expression by OSC in HT–22 Cells. (<b>A</b>) Cells were exposed to a range of OSC concentrations (1.25, 2.5, 5, 10 μM) for a duration of 12 h to determine the dose-dependent effect on HO–1 expression. Cobalt protoporphyrin (CoPP), at a concentration of 20 μM, served as a positive control to validate the response. (<b>B</b>) To explore the time course of HO–1 induction, cells were treated with a fixed concentration of 10 μM OSC for varying periods. The protein expression of HO–1 was quantified using WB analysis, a method that allows for the detection and quantification of specific proteins. The results were normalized relative to actin, a reference protein, to adjust for any variations in protein loading. The data presentation follows the standard format, where each bar graph segment illustrates the mean ± SD from triplicate samples (n = 3), ensuring the reproducibility and reliability of the findings. Statistical significance is denoted by the symbols <span class="html-italic"><sup>#</sup> p</span> &lt; 0.05 and <span class="html-italic"><sup>##</sup> p</span> &lt; 0.01, which indicate significant differences in HO–1 expression levels when compared to the control group without treatment. The presence or absence of treatment is indicated by “+” and “−” signs, respectively.</p>
Full article ">Figure 8
<p>Impact of HO–1 knockdown on HT–22 cell response to OSC and glutamate challenge. This study delineates the consequences of HO–1 suppression in HT–22 cells under conditions designed to mimic OS. (<b>A</b>,<b>D</b>) The survival of HT–22 cells, following pretreatment with 10 μM OSC in conjunction with or without 50 μM SnPP and si−HO–1, was subsequently challenged with 20 mM glutamate for 24 h. The quantitative assessment of cell viability was performed using the MTT assay. (<b>B</b>,<b>E</b>) The production of ROS was evaluated through DCF fluorescence measurement, providing a quantitative assessment of intracellular ROS levels. (<b>C</b>) Representative WB images illustrate the levels of HO–1 protein expression in the treated cells, offering a visual confirmation of the HO–1 knockdown efficacy. Data are presented as the mean ± SD from three independent experiments (n = 3), ensuring the statistical robustness of the findings. Statistical significance is indicated as follows: <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05 and <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 denote significant differences relative to the untreated control group; * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 indicate substantial differences when compared to the group treated solely with 20 mM glutamate; <sup>%</sup> <span class="html-italic">p</span> &lt; 0.05 and <sup>%%</sup> <span class="html-italic">p</span> &lt; 0.01 signify substantial differences in comparison to the group treated with 10 μM OSC. The presence or absence of specific treatments is denoted by the symbols “+” and “−”.</p>
Full article ">
15 pages, 1242 KiB  
Article
Metabolic Effects of Sodium Thiosulfate During Resuscitation from Trauma and Hemorrhage in Cigarette-Smoke-Exposed Cystathionine-γ-Lyase Knockout Mice
by Maximilian Feth, Felix Hezel, Michael Gröger, Melanie Hogg, Fabian Zink, Sandra Kress, Andrea Hoffmann, Enrico Calzia, Ulrich Wachter, Peter Radermacher and Tamara Merz
Biomedicines 2024, 12(11), 2581; https://doi.org/10.3390/biomedicines12112581 - 12 Nov 2024
Viewed by 724
Abstract
Background: Acute and chronic pre-traumatic cigarette smoke exposure increases morbidity and mortality after trauma and hemorrhage. In mice with a genetic deletion of the H2S-producing enzyme cystathione-γ-lyase (CSE−/−), providing exogenous H2S using sodium thiosulfate (Na2S [...] Read more.
Background: Acute and chronic pre-traumatic cigarette smoke exposure increases morbidity and mortality after trauma and hemorrhage. In mice with a genetic deletion of the H2S-producing enzyme cystathione-γ-lyase (CSE−/−), providing exogenous H2S using sodium thiosulfate (Na2S2O3) improved organ function after chest trauma and hemorrhagic shock. Therefore, we evaluated the effect of Na2S2O3 during resuscitation from blunt chest trauma and hemorrhagic shock on CSE−/− mice with pre-traumatic cigarette smoke (CS) exposure. Since H2S is well established as being able to modify energy metabolism, a specific focus was placed on whole-body metabolic pathways and mitochondrial respiratory activity. Methods: Following CS exposure, the CSE−/− mice underwent anesthesia, surgical instrumentation, blunt chest trauma, hemorrhagic shock for over 1 h (target mean arterial pressure (MAP) ≈ 35 ± 5 mmHg), and resuscitation for up to 8 h comprising lung-protective mechanical ventilation, the re-transfusion of shed blood, fluid resuscitation, and continuous i.v. noradrenaline (NoA) to maintain an MAP ≥ 55 mmHg. At the start of the resuscitation, the mice randomly received either i.v. Na2S2O3 (0.45 mg/gbodyweight; n = 14) or the vehicle (NaCl 0.9%; n = 11). In addition to the hemodynamics, lung mechanics, gas exchange, acid–base status, and organ function, we quantified the parameters of carbohydrate, lipid, and protein metabolism using a primed continuous infusion of stable, non-radioactive, isotope-labeled substrates (gas chromatography/mass spectrometry) and the post-mortem tissue mitochondrial respiratory activity (“high-resolution respirometry”). Results: While the hemodynamics and NoA infusion rates did not differ, Na2S2O3 was associated with a trend towards lower static lung compliance (p = 0.071) and arterial PO2 (p = 0.089) at the end of the experiment. The direct, aerobic glucose oxidation rate was higher (p = 0.041) in the Na2S2O3-treated mice, which resulted in lower glycemia levels (p = 0.050) and a higher whole-body CO2 production rate (p = 0.065). The mitochondrial respiration in the heart, kidney, and liver tissue did not differ. While the kidney function was comparable, the Na2S2O3-treated mice showed a trend towards a shorter survival time (p = 0.068). Conclusions: During resuscitation from blunt chest trauma and hemorrhagic shock in CSE−/− mice with pre-traumatic CS exposure, Na2S2O3 was associated with increased direct, aerobic glucose oxidation, suggesting a switch in energy metabolism towards preferential carbohydrate utilization. Nevertheless, treatment with Na2S2O3 coincided with a trend towards worsened lung mechanics and gas exchange, and, ultimately, shorter survival. Full article
(This article belongs to the Special Issue Molecular Mechanisms and Therapeutics in Hemorrhagic Shock)
Show Figures

Figure 1

Figure 1
<p>Experimental setup, timeline, surgical instrumentation, and experimental protocol. MAP, mean arterial pressure; Na<sub>2</sub>S<sub>2</sub>O<sub>3</sub>, sodium thiosulfate; TxT, chest trauma; i.v., intravenous.</p>
Full article ">Figure 2
<p>Kaplan–Meier survival curve in the vehicle group (solid line) and Na<sub>2</sub>S<sub>2</sub>O<sub>3</sub> group (dotted line). Time “0” refers to the start of the observation period, with the initiation of blunt chest trauma. No significant differences were observed (log rank Mantel–Cox test, <span class="html-italic">p</span> = 0.068).</p>
Full article ">Figure 3
<p>The evaluation of metabolic pathways. The individual results as well as the median and interquartile range (endogenous glucose production rate) or mean ± standard deviation (glucose oxidation; glycerol, urea, and leucine production rates) are shown, according to the absence/presence of a normal data distribution for the metabolic parameters as assessed using stable, non-radioactive isotope-labeled substrates (glucose, glycerol, leucine, and urea) between mice in the vehicle (open circles) and the Na<sub>2</sub>S<sub>2</sub>O<sub>3</sub> (solid triangles) groups. Differences between the two treatment groups were analyzed using the Mann–Whitney U rank sum test or a Student’s <span class="html-italic">t</span>-test as appropriate.</p>
Full article ">Figure 4
<p>The evaluation of mitochondrial activity. The individual results as well as the median and interquartile range (liver OxPhos and ETC, kidney ETC) or mean ± standard deviation (heart OxPhos and ETC, kidney OxPhos) are shown according to the absence/presence of a normal data distribution for the parameters of mitochondrial respiratory activity in immediate postmortem specimens of the heart, liver, and kidney. The oxidative phosphorylation (OxPhos) (left panel) and maximal electron transfer capacity in the uncoupled state (ETC, right panel) are presented for mice in the vehicle (open circles) and Na<sub>2</sub>S<sub>2</sub>O<sub>3</sub> (solid triangles) groups. Differences between the two treatment groups were analyzed using the Mann–Whitney U rank sum test or a Student’s t-test as appropriate.</p>
Full article ">
16 pages, 8178 KiB  
Article
A New Probiotic Formulation Promotes Resolution of Inflammation in a Crohn’s Disease Mouse Model by Inducing Apoptosis in Mucosal Innate Immune Cells
by Carlo De Salvo, Abdullah Osme, Mahmoud Ghannoum, Fabio Cominelli and Luca Di Martino
Int. J. Mol. Sci. 2024, 25(22), 12066; https://doi.org/10.3390/ijms252212066 - 10 Nov 2024
Viewed by 819
Abstract
The interaction between gut-residing microorganisms plays a critical role in the pathogenesis of Crohn’s disease (CD), where microbiome dysregulation can alter immune responses, leading to unresolved local inflammation. The aim of this study is to analyze the immunomodulatory properties of a recently developed [...] Read more.
The interaction between gut-residing microorganisms plays a critical role in the pathogenesis of Crohn’s disease (CD), where microbiome dysregulation can alter immune responses, leading to unresolved local inflammation. The aim of this study is to analyze the immunomodulatory properties of a recently developed probiotic + amylase blend in the SAMP1/YitFc (SAMP) mouse model of CD-like ileitis. Four groups of SAMP mice were gavaged for 56 days with the following treatments: 1) probiotic strains + amylase (0.25 mg/100 µL PBS); 2) only probiotics; 3) only amylase; PBS-treated controls. Ilea were collected for GeoMx Digital Spatial Profiler (DSP) analysis and histological evaluation. Histology assessment for inflammation indicated a significantly reduced level of ileitis in mice administered the probiotics + amylase blend. DSP analysis showed decreased abundance of neutrophils and increased abundance of dendritic cells, regulatory T cells, and macrophages, with a significant enrichment of five intracellular pathways related to apoptosis, in probiotics + amylase-treated mice. Increased apoptosis occurrence was confirmed by (TdT)- deoxyuridine triphosphate (dUTP)-biotin nick end labeling assay. Our data demonstrate a beneficial role of the probiotic and amylase blend, highlighting an increased apoptosis of innate immunity-associated cell subsets, thus promoting the resolution of inflammation. Hence, we suggest that the developed probiotic enzyme blend may be a therapeutic tool to manage CD and therefore is a candidate formulation to be tested in clinical trials. Full article
(This article belongs to the Special Issue The Role of Microbiota in Immunity and Inflammation)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Amylase and probiotics are both necessary to ameliorate inflammation in SAMP mice. (<b>A</b>) Histology evaluation shows significant attenuation of ileitis in mice treated with amylase plus probiotic mix (P + A) in comparison with the mice treated with only amylase (A), only probiotic mix (P), or PBS (control) (C) (one-way ANOVA, 8.50 ± 2.46 vs. 13.20 ± 2.25 vs. 13.30 ± 4.08 vs. 15.20 ± 2.97; <span class="html-italic">p</span> &lt; 0.001). (<b>B</b>) Representative pictures of H&amp;E-stained sections indicate that mice treated with both amylase and probiotic mix have better-preserved architecture of the villi (red arrows) and less presence of inflammatory cells in the mucosal and submucosal layer (yellow arrows) compared to the other three groups. Data are represented as mean ± SEM and are representative of two independent experiments; *** <span class="html-italic">p</span> &lt; 0.001. N = 10/group. 10X + 1.25 original mag.</p>
Full article ">Figure 2
<p>Digital spatial profiling of genetic expression performed on mucosal ileal tissue in SAMP mice. (<b>A</b>) Schematic workflow showing immunohistochemistry (IHC)/immunofluorescence (IF) staining of slides embedded in paraffin with markers for PanCK, CD45, and nuclei in the ROIs. (<b>B</b>) Representative images of mucosal layers of stained tissues with segments superimposed and with probe counts indicating PanCK<sup>+</sup> cells (Cy3568 nm; yellow), CD45<sup>+</sup> cells (Texas red, 615 nm; red), and nuclear staining (FITC, 525 nm; green). N = 6 mice/group; N = 6 ROIs/mouse.</p>
Full article ">Figure 3
<p>Probiotics mix coupled with amylase exerts an enhanced immunomodulatory effect in the mucosal layer. (<b>A</b>) Principal component analysis showing the variance among ROIs based on cell type abundance beta estimates (PC1 variance: 22.3%; PC2 variance: 4.3%), highlighting significant difference in the mice administered the probiotics + amylase mix (P + A) compared to only amylase (A), only probiotics (P), and control (C) groups (PC1: Kruskal–Wallis test, 69.94 ± 3.94 vs. 5.39 ± 7.35 vs. −42.63 ± 5.89 vs. −30.00 ± 5.56; <span class="html-italic">p</span> &lt; 0.001; N = 36 ROIs/group). (<b>B</b>) Dendrogram showing estimated relative abundance of immune cell subsets in microenvironment segments, indicating a significant decrease in (<b>C</b>) neutrophils (one-way ANOVA, 1.22 × 10<sup>−3</sup> ± 0.14 × 10<sup>−3</sup> vs. 6.43 × 10<sup>−3</sup> ± 1.35 × 10<sup>−3</sup> vs. 5.30 × 10<sup>−3</sup> ± 0.61 × 10<sup>−3</sup> vs. 4.21 × 10<sup>−3</sup> ± 0.76 × 10<sup>−3</sup>; <span class="html-italic">p</span> &lt; 0.02) and increase in (<b>D</b>) dendritic cells (one-way ANOVA, 2.29 × 10<sup>−2</sup> ± 0.46 × 10<sup>−2</sup> vs. 1.12 × 10<sup>−2</sup> ± 0.33 × 10<sup>−2</sup> vs. 0.91 × 10<sup>−2</sup> ± 0.25 × 10<sup>−2</sup>; <span class="html-italic">p</span> &lt; 0.02), (<b>E</b>) innate lymphoid cells (ILC)s (one-way ANOVA, 7.19 × 10<sup>−2</sup> ± 0.42 × 10<sup>−2</sup> vs. 4.94 × 10<sup>−3</sup> ± 0.63 × 10<sup>−2</sup> vs. 5.35 × 10<sup>−2</sup> ± 0.36 × 10<sup>−2</sup> vs. 4.68 × 10<sup>−2</sup> ± 0.43 × 10<sup>−2</sup>; <span class="html-italic">p</span> &lt; 0.02), (<b>F</b>) macrophages (one-way ANOVA, 6.15 × 10<sup>−2</sup> ± 1.29 × 10<sup>−2</sup> vs. 2.28 × 10<sup>−2</sup> ± 0.58 × 10<sup>−2</sup> vs. 3.73 × 10<sup>−2</sup> ± 0.34 × 10<sup>−2</sup> vs. 2.21 × 10<sup>−2</sup> ± 0.46 × 10<sup>−2</sup>; <span class="html-italic">p</span> &lt; 0.02), and (<b>G</b>) regulatory T cells (Treg)s (one-way ANOVA, 1.05 × 10<sup>−2</sup> ± 0.12 × 10<sup>−2</sup> vs. 0.80 × 10<sup>−2</sup> ± 0.07 × 10<sup>−2</sup> vs. 0.44 × 10<sup>−2</sup> ± 0.08 × 10<sup>−2</sup> vs. 0.30 × 10<sup>−2</sup> ± 0.12 × 10<sup>−2</sup>; <span class="html-italic">p</span> &lt; 0.02) in the group treated with probiotic + amylase compared with control, only amylase, and only probiotics groups. Data are represented as mean ± SEM and are representative of two independent experiments. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.02; *** <span class="html-italic">p</span> &lt; 0.001. N = 6 mice/group; N = 6 ROIs/mouse.</p>
Full article ">Figure 4
<p>Probiotic + amylase blend stimulates apoptosis in immune cell subsets. (<b>A</b>–<b>C</b>) Volcano plots obtained from the pathway analysis of CD45<sup>+</sup> cells in ROIs of the four experimental groups. Statistically significant different apoptosis-related pathways are labeled in yellow (<span class="html-italic">p</span> &lt; 0.05; normalized enrichment score &gt; ±2) and in gray when not significantly different. Amylase and the probiotic mix are both necessary to significantly induce apoptosis in CD45<sup>+</sup> cells. Comparison between probiotics without amylase and the control group shows no significantly different apoptosis-related pathways between the two groups. (<b>D</b>) Heatmap of differentially expressed genes related to apoptosis pathways between the control group and probiotic + amylase, (<b>E</b>) only probiotics, and (<b>F</b>) only amylase groups. (<b>G</b>) Representative IF photomicrographs of TUNEL assay show the increased number of apoptotic TUNEL-positive cells (green, Alexa Fluor™ 488) and fewer neutrophils (red, Alexa Fluor™ 594) deep in the lamina propria of mice treated with the probiotic + amylase blend compared to the control group (DAPI, blue). 40X original mag. N = 6 mice/group.</p>
Full article ">
25 pages, 453 KiB  
Review
How to Keep Lactose Avoiders Healthy
by Zlatina Chengolova, Petar Shentov, Radina Ivanova and Reni Syarova
Dairy 2024, 5(4), 702-726; https://doi.org/10.3390/dairy5040052 - 6 Nov 2024
Viewed by 956
Abstract
A large portion of the world’s population has lactose intolerance. Fundamentally, this condition occurs when the small intestine does not produce enough of the lactase enzyme, which digests the disaccharide lactose in milk. Lactose avoiders might unconsciously decide to limit or exclude milk [...] Read more.
A large portion of the world’s population has lactose intolerance. Fundamentally, this condition occurs when the small intestine does not produce enough of the lactase enzyme, which digests the disaccharide lactose in milk. Lactose avoiders might unconsciously decide to limit or exclude milk and dairy products from their diets. This group includes people with lactose intolerance, people with an allergy to milk protein, vegans, and those expressing personal preferences. Lactose avoiders are often self-reported as being milk intolerant. In this review, specific amounts of lactose in different types of milk and milk products are presented. The amounts of micro- and macronutrients in them are compared with the daily requirements established by accepted sources. Foods are suggested that can play vital roles in permanently avoiding lactose-containing dairy products, for example, brussels sprouts, as a good source of vitamin B1; kale, as a source of vitamin K; and cereals at breakfast for vitamin B6. Attention is paid to mature cheeses as they are extremely beneficial for health due to their rich vitamin and elemental compositions, and they are also suitable for people with lactose intolerance due to their low lactose content. This information is rarely provided on packaging. In addition, the current state of labeling for the presence of lactose in food and pharmaceutical products is discussed. The term “hidden lactose” is introduced to include added lactose in unexpected foods, drinks, and even medicines. Full article
(This article belongs to the Section Milk and Human Health)
14 pages, 1404 KiB  
Article
Development of an Indirect Competitive ELISA Based on a Stable Epitope of β-Lactoglobulin for Its Detection in Hydrolyzed Formula Milk Powder
by Qinggang Xie, Yuhao Huang, Xianli Zhang, Xiaoxi Xu and Zhenxing Li
Foods 2024, 13(21), 3477; https://doi.org/10.3390/foods13213477 - 30 Oct 2024
Viewed by 938
Abstract
The target of traditional immunological detection methods for milk allergens is usually the whole β-lactoglobulin molecule. However, thermal processes and hydrolysis can destroy the epitope of β-lactoglobulin and interfere with its accurate detection and labeling in prepackaged foods, posing a health risk to [...] Read more.
The target of traditional immunological detection methods for milk allergens is usually the whole β-lactoglobulin molecule. However, thermal processes and hydrolysis can destroy the epitope of β-lactoglobulin and interfere with its accurate detection and labeling in prepackaged foods, posing a health risk to milk-allergic patients. There currently remains a need to excavate and locate recognition sites for β-lactoglobulin in thermally processed and hydrolyzed products. Therefore, a stable epitope of β-lactoglobulin (CAQKKIIAEKTKIPAVFKIDA) was selected as the ideal recognition site, and an indirect competitive enzyme-linked immunosorbent assay (ELISA) was developed using an antibody against this stable β-lactoglobulin epitope in order to improve the detection of β-lactoglobulin in thermally processed and hydrolyzed foods in this study. The stable epitope of β-lactoglobulin was selected using a molecular dynamics simulation, and the binding ability of anti-stable epitope antibodies was characterized using indirect ELISA and indirect competitive ELISA. The limit of detection (LOD) and limit of quantitation (LOQ) of the established ELISA were 0.25 and 1.07 mg·kg−1, respectively. Furthermore, the developed ELISA only showed cross-reactivity to goat milk among 23 common foods, therefore exhibiting high specificity to bovine β-lactoglobulin. In addition, the developed ELISA was able to effectively detect β-lactoglobulin residue in processed commercial foods and hydrolyzed formula milk powder. Our findings provide a novel strategy for accurately detecting milk allergens based on stable epitope recognition in thermally processed and hydrolyzed foods. Full article
(This article belongs to the Section Food Analytical Methods)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Molecular dynamics simulation of β-lactoglobulin. The black, red, blue, green, and purple lines represent simulation at 298.15 K (25 °C), 333.15 K (60 °C), 353.15 K (80 °C), and 373.15 K (100 °C) for 1 bar and 394.1 K (121 °C) for 1.04 bar.</p>
Full article ">Figure 2
<p>Evaluation of the binding ability of three anti-epitope antibodies using indirect enzyme-linked immunosorbent assay (ELISA) and inhibition ELISA. The binding ability of the following antibodies was validated using indirect ELISA: (<b>A</b>) anti-BLG-1; (<b>B</b>) anti-BLG-2; (<b>C</b>) anti-BLG-3. (<b>D</b>) The binding ability of the three anti-epitope antibodies was validated using inhibition ELISA.</p>
Full article ">Figure 3
<p>The fitted curve of the developed indirect competitive ELISA using a four-parameter logistic formula. <span class="html-italic">B</span> represents the measured absorbance of the serial concentrations of β-lactoglobulin at 450 nm. <span class="html-italic">B</span><sub>0</sub> represents the measured absorbance at 450 nm of the control dilution buffer.</p>
Full article ">
23 pages, 3354 KiB  
Article
Labneh: A Retail Market Analysis and Selected Product Characterization
by Raman K. Bhaskaracharya, Fatima Saeed Rashed Alnuaimi, Shaikha Rashed Juma Aldarmaki, Abeena Abdulazeez and Mutamed Ayyash
Foods 2024, 13(21), 3461; https://doi.org/10.3390/foods13213461 - 29 Oct 2024
Viewed by 1160
Abstract
Labneh is a popular fermented dairy product, which contemporarily has diversified into a varied range of styles, formulated with the inclusion of multiple additives, and is sourced across the globe. This has driven labneh’s emergence as a complex product with varying textural and [...] Read more.
Labneh is a popular fermented dairy product, which contemporarily has diversified into a varied range of styles, formulated with the inclusion of multiple additives, and is sourced across the globe. This has driven labneh’s emergence as a complex product with varying textural and rheological characteristics. The lack of scientific literature about labneh products available in the United Arab Emirates (UAE) market and their characterization has prompted this study. A detailed UAE market analysis of labneh for label, formulation, nutrition, and price variability was conducted. Surveyed labneh products were categorized as unpackaged, multinational company (MNC), small and medium enterprise (SME), and specialty products. They differed in manufacturing, such as acid ± enzyme coagulation with/without post-fermentation heat treatment, and contained various stabilizers, emulsifiers, preservatives, and processing aids. Interestingly, almost equal proportions, 64.7% and 67%, of MNC and SME labneh contained additives, respectively. All MNC labneh were post-heat-treated, in contrast to only 7% of SME labneh. Organic labneh and non-bovine milk-based labneh are not yet widely available. The second part of the study involved the physicochemical characterization of a select number of packaged labneh that were categorized in accordance with fat content as high-fat (17–18%), full-fat (7.1–8%), and lite-fat (3.5–4.5%). High-fat labneh showed a significantly higher complex viscosity, complex modulus, hardness, adhesiveness, stringiness, and fracturability, followed by lite-fat labneh compared to full-fat labneh, especially when it contained pectin. Full-fat labneh with added gums (and starch) and high-fat labneh with gums showed a significantly lower complex modulus compared to their respective control labneh. This study highlights the variety of commercial labneh products available and differences in their formulation, manufacturing, and composition, and provides specific dependencies of materials with their physicochemical characteristics. Full article
(This article belongs to the Section Dairy)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Market analysis of commercial labneh in UAE market.</p>
Full article ">Figure 2
<p>Product characteristics of SME vs. MNC labneh.</p>
Full article ">Figure 3
<p>Density plot of reported nutritional content of SME vs. MNC labneh products. (<b>A</b>) Calorific content, (<b>B</b>) % protein, (<b>C</b>) % fat, (<b>D</b>) % saturated fat, (<b>E</b>) % total carbohydrate, and (<b>F</b>) = % salt.</p>
Full article ">Figure 4
<p>Rheological differences. (<b>A</b>) Storage modulus and (<b>B</b>) loss modulus measured for high-fat labneh. H1, H2—control, and H3. (<b>A</b>) Storage modulus measured over a range of angular frequencies. (<b>B</b>) Loss modulus measured using frequency sweep. (Represented data points are Mean ± SEM).</p>
Full article ">Figure 5
<p>Rheological differences in measured storage modulus during frequency sweep for full-fat (F2/F3 control and F1 with additives) labneh. (Represented data points are Mean ± SEM).</p>
Full article ">Figure 6
<p>Rheological differences. (<b>A</b>) Storage modulus and (<b>B</b>) loss modulus measured for lite-fat labneh L1 and L2 (both labneh without additives). (<b>A</b>) Storage modulus lite-fat when subjected to oscillation stress. (<b>B</b>) Loss modulus of lite-fat labneh measured using frequency sweep. (Represented data points are Mean ± SEM).</p>
Full article ">Figure 7
<p>Texture characteristics. (<b>A</b>) Hardness, (<b>B</b>) adhesiveness, and (<b>C</b>) fracturability) of high-fat containing (H1–H3), full-fat containing (F1–F3), and lite-fat containing (L1, L2) labneh. In this figure ** means <span class="html-italic">p</span> &lt; 0.01; *** means <span class="html-italic">p</span> &lt; 0.001 and **** means <span class="html-italic">p</span> &lt; 0.0001. The <span class="html-italic">p</span> values indicate the level of significance as per ANOVA analysis.</p>
Full article ">Figure 8
<p>PCA of rheological and texture characteristics by type of labneh showing high-fat containing (H1–H3), full-fat containing (F1–F3), and lite-fat containing (L1, L2) labneh.</p>
Full article ">Figure 9
<p>Biplot of textural characteristics by type of Labneh sample of high-fat (H1–H3), full-fat containing (F1–F3), and lite-fat containing (L1, L2) labneh.</p>
Full article ">
20 pages, 3017 KiB  
Article
A Novel PCR-Free Ultrasensitive GQD-Based Label-Free Electrochemical DNA Sensor for Sensitive and Rapid Detection of Francisella tularensis 
by Sumeyra Savas and Melike Sarıçam
Micromachines 2024, 15(11), 1308; https://doi.org/10.3390/mi15111308 - 28 Oct 2024
Viewed by 832
Abstract
Biological warfare agents are infectious microorganisms or toxins capable of harming or killing humans. Francisella tularensis is a potential bioterrorism agent that is highly infectious, even at very low doses. Biosensors for biological warfare agents are simple yet reliable point-of-care analytical tools. Developing [...] Read more.
Biological warfare agents are infectious microorganisms or toxins capable of harming or killing humans. Francisella tularensis is a potential bioterrorism agent that is highly infectious, even at very low doses. Biosensors for biological warfare agents are simple yet reliable point-of-care analytical tools. Developing highly sensitive, reliable, and cost-effective label-free DNA biosensors poses significant challenges, particularly when utilizing traditional techniques such as fluorescence, electrochemical methods, and others. These challenges arise primarily due to the need for labeling, enzymes, or complex modifications, which can complicate the design and implementation of biosensors. In this study, we fabricated Graphene Quantum dot (GQD)-functionalized biosensors for highly sensitive label-free DNA detection. GQDs were immobilized on the surface of screen-printed gold electrodes via mercaptoacetic acid with a thiol group. The single-stranded DNA (ssDNA) probe was also immobilized on GQDs through strong π−π interactions. The ssDNA probe can hybridize with the ssDNA target and form double-stranded DNA, leading to a decrease in the effect of GQD but a positive shift associated with the increase in DNA concentration. The specificity of the developed system was observed with different microorganism target DNAs and up to three-base mismatches in the target DNA, effectively distinguishing the target DNA. The response time for the target DNA molecule is approximately 1010 s (17 min). Experimental steps were monitored using UV/Vis spectroscopy, Atomic Force Microscopy (AFM), and electrochemical techniques to confirm the successful fabrication of the biosensor. The detection limit can reach 0.1 nM, which is two–five orders of magnitude lower than previously reported methods. The biosensor also exhibits a good linear range from 105 to 0.01 nM and has good specificity. The biosensor’s detection limit (LOD) was evaluated as 0.1 nM from the standard calibration curve, with a correlation coefficient of R2 = 0.9712, showing a good linear range and specificity. Here, we demonstrate a cost-effective, GQD-based SPGE/F. tularensis DNA test suitable for portable electrochemical devices. This application provides good perspectives for point-of-care portable electrochemical devices that integrate sample processing and detection into a single cartridge without requiring a PCR before detection. Based on these results, it can be concluded that this is the first enzyme-free electrochemical DNA biosensor developed for the rapid and sensitive detection of F. tularensis, leveraging the nanoenzyme and catalytic properties of GQDs. Full article
(This article belongs to the Special Issue Biosensors for Pathogen Detection 2024)
Show Figures

Figure 1

Figure 1
<p>Principle of the GQD-based DNA sensor for <span class="html-italic">F. tularensis</span> detection.</p>
Full article ">Figure 2
<p>(<b>a</b>) Cyclic voltammetry measurement to determine the most ideal amperometric measurement. (<b>b</b>) Real time measurement curves obtained with four different GQDs concentrations. (<b>c</b>) Optical properties of GQDs at decreasing concentrations.</p>
Full article ">Figure 3
<p>Comparative UV/Vis spectra of GQDs, GQDs-N, GQDs-Capture DNA, GQDs-BSA and GQDs-Target DNA.</p>
Full article ">Figure 4
<p>Optimization of the ideal hybridization time between capture DNA (ssDNA) and 10<sup>5</sup> nM target DNA (gDNA) on the sensor surface.</p>
Full article ">Figure 5
<p>(<b>a</b>) The amperometric measurement curves obtained for eight concentrations of <span class="html-italic">F. tularensis</span> DNA (dsDNA) concentration using 5000 ppm GQDs, 5000 ppm GQDs, and GQDs-ssDNA attachment. (<b>b</b>) The linear calibration curve of DNA assays (10<sup>5</sup> nM–0.1 nM) with a correlation coefficient of 0.9712. (<b>c</b>) Sensogram of eight concentrations of <span class="html-italic">F. tularensis</span> DNA (dsDNA). (<b>d</b>) The amperometric measurement curves obtained only between 10 nM–0.01 nM.</p>
Full article ">Figure 5 Cont.
<p>(<b>a</b>) The amperometric measurement curves obtained for eight concentrations of <span class="html-italic">F. tularensis</span> DNA (dsDNA) concentration using 5000 ppm GQDs, 5000 ppm GQDs, and GQDs-ssDNA attachment. (<b>b</b>) The linear calibration curve of DNA assays (10<sup>5</sup> nM–0.1 nM) with a correlation coefficient of 0.9712. (<b>c</b>) Sensogram of eight concentrations of <span class="html-italic">F. tularensis</span> DNA (dsDNA). (<b>d</b>) The amperometric measurement curves obtained only between 10 nM–0.01 nM.</p>
Full article ">Figure 6
<p>AFM 3D topography images of (<b>a</b>) GQD-laminated surface, (<b>b</b>) Capture DNA immobilized and (<b>c</b>) After hybridizationTarget DNA binding.</p>
Full article ">Figure 7
<p>Current responses for the detection of non-complementary DNA targets (<span class="html-italic">Salmonella</span> spp. target <span class="html-italic">Y. pestis</span> target), three-base mismatched DNA targets, one-base mismatched DNA targets, and complementary DNA targets.</p>
Full article ">Figure 8
<p>Gel electrophoresis experiments to verify the applicability of the DNA chains designed in the article (the forward primer: GCT GTA TCA TCA TTT AAT AAA CTG CTG and reverse primer: TTG GGA AGC TTG TAT CAT GGC ACT pair was used and a tul4 gene of 428 bp size was detected).</p>
Full article ">Figure 9
<p>Comparison of the similarity ratios for the same concentrations measured in three different measurements.</p>
Full article ">
Back to TopTop