[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (688)

Search Parameters:
Keywords = peptide substrate

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
27 pages, 30182 KiB  
Article
Synthetic Extracellular Matrix of Polyvinyl Alcohol Nanofibers for Three-Dimensional Cell Culture
by Thi Xuan Thuy Tran, Gyu-Min Sun, Hue Vy An Tran, Young Hun Jeong, Petr Slama, Young-Chae Chang, In-Jeong Lee and Jong-Young Kwak
J. Funct. Biomater. 2024, 15(9), 262; https://doi.org/10.3390/jfb15090262 - 10 Sep 2024
Viewed by 422
Abstract
An ideal extracellular matrix (ECM) replacement scaffold in a three-dimensional cell (3D) culture should induce in vivo-like interactions between the ECM and cultured cells. Highly hydrophilic polyvinyl alcohol (PVA) nanofibers disintegrate upon contact with water, resulting in the loss of their fibrous morphology [...] Read more.
An ideal extracellular matrix (ECM) replacement scaffold in a three-dimensional cell (3D) culture should induce in vivo-like interactions between the ECM and cultured cells. Highly hydrophilic polyvinyl alcohol (PVA) nanofibers disintegrate upon contact with water, resulting in the loss of their fibrous morphology in cell cultures. This can be resolved by using chemical crosslinkers and post-crosslinking. A crosslinked, water-stable, porous, and optically transparent PVA nanofibrous membrane (NM) supports the 3D growth of various cell types. The binding of cells attached to the porous PVA NM is low, resulting in the aggregation of cultured cells in prolonged cultures. PVA NMs containing integrin-binding peptides of fibronectin and laminin were produced to retain the blended peptides as cell-binding substrates. These peptide-blended PVA NMs promote peptide-specific cell adherence and growth. Various cells, including epithelial cells, cultured on these PVA NMs form layers instead of cell aggregates and spheroids, and their growth patterns are similar to those of the cells cultured on an ECM-coated PVA NM. The peptide-retained PVA NMs are non-stimulatory to dendritic cells cultured on the membranes. These peptide-retaining PVA NMs can be used as an ECM replacement matrix by providing in vivo-like interactions between the matrix and cultured cells. Full article
(This article belongs to the Special Issue Advanced Technologies for Processing Functional Biomaterials)
Show Figures

Figure 1

Figure 1
<p>Production of water-stable and optically transparent PVA NMs. The membranes are untreated or soaked in distilled water and dried. (<b>a</b>) Structure of nanofibers measured by SEM. (<b>b</b>) Visual and optical transparency of membrane scanned using a spectrophotometer. (<b>c</b>) The structure and diameter of the PVA NMs measured by SEM.</p>
Full article ">Figure 2
<p>FTIR spectra of peptide-blended PVA NMs. Electrospun PVA NM is untreated (DW-treated PVA NM) or treated (HCl/DW-treated PVA NM) with HCl vapor for 2 min, followed by soaking in water for 18 h. The NMs are washed three times with PBS, dried, and then analyzed by FTIR spectroscopy. (<b>a</b>) FTIR spectra of HCl-treated PVA NM, (<b>b</b>) YIGSR-blended PVA NM, and (<b>c</b>) HCl and/or DW-treated YIGSR-PVA NMs.</p>
Full article ">Figure 3
<p>The 3D adhesion of cultured cells on PVA NMs. (<b>a</b>) CellTracker Red-labeled NIH 3T3 cells on PVA NM. (<b>b</b>) Cells alone without PVA NM imaging. (<b>c</b>) Cross-sectioned view of cells on PVA NM analyzed using a confocal microscope. Images are shown using the surface function of Imaris software. (<b>d</b>,<b>e</b>) NIH 3T3 and MLE-12 cells are cultured on HCl vapor-treated PVA NM and observed using SEM.</p>
Full article ">Figure 4
<p>Adhesion of cultured cells on PVA NMs. Merged fluorescence and DIC images of cells labeled with CellTracker Red and cultured for the indicated times on the membranes.</p>
Full article ">Figure 5
<p>Effects of NaOH treatment on peptide release and cell adhesion in peptide-blended PVA NMs. (<b>a</b>) Fluorescence levels in the media measured by fluorescence spectrometry and the membranes measured by confocal microscope. (<b>b</b>) Merged fluorescence and DIC images of fluorescence-labeled cells. (<b>c</b>) Images shown using the surface function of Imaris software and SEM. * <span class="html-italic">p</span> &lt; 0.05 versus 0.05 h treatment.</p>
Full article ">Figure 6
<p>Culture of cells on peptide-retained PVA NMs. (<b>a</b>) DIC images of the cells and CCK-8 assay of viable cells in culture media and attached to NMs. (<b>b</b>) CCK-8 assay of viable cells adhered to the membranes. (<b>c</b>) DIC and fluorescence images using a confocal microscope and CCK-8 assay of viable cells. * <span class="html-italic">p</span> &lt; 0.05 versus untreated. ** <span class="html-italic">p</span> &lt; 0.05 versus PVA NM. *** <span class="html-italic">p</span> &lt; 0.05 versus none.</p>
Full article ">Figure 7
<p>Culture of cells on PVA NMs containing various types of peptides. (<b>a</b>) NIH 3T3 and MLE-12 cells, (<b>b</b>) MLE-12 cells, and (<b>c</b>) primary colon epithelial cells observed using a confocal microscope and numbers of detached cells were counted based on 9 different areas (marked from 1 to 9). * <span class="html-italic">p</span> &lt; 0.05 versus without peptides. ** <span class="html-italic">p</span> &lt; 0.05 versus with one type of peptide. *** <span class="html-italic">p</span> &lt; 0.05 versus YIGSR. **** <span class="html-italic">p</span> &lt; 0.05 versus PVA NMs.</p>
Full article ">Figure 8
<p>Pattern of cell growth on ECM protein-coated and peptide-retained PVA NMs. (<b>a</b>) Imaris view of confocal microscopic images. The scale bar is 50 μm. (<b>b</b>) NIH 3T3 and (<b>c</b>) MLE-12 cells shown in confocal microscopic images, with cell nuclei (Hoechst 33342, blue), actin microfilaments (Alexa Flour 488 Phalloidin, green), zona occludin-1 (ZO-1, red) Arrows indicate ZO-1 expression.</p>
Full article ">Figure 9
<p>Growth rate of the cells cultured on the peptide-retained PVA NMs. (<b>a,b</b>) CCK-8 assay of NIH 3T3 cells. (<b>c</b>) NIH 3T3 cells cultured in the conditioned media on the culture plate. (<b>d</b>) CCK-8 assay of HepG2 cells. * <span class="html-italic">p</span> &lt; 0.05 versus culture plate, ** <span class="html-italic">p</span> &lt; 0.05 versus PVA NM.</p>
Full article ">Figure 10
<p>Culture of BMDCs on peptide-blended PVA NMs. (<b>a</b>) Cell morphology assessed using SEM. (<b>b</b>) Expression level of CD86 measured by flow cytometry. (<b>c</b>) The concentration of TNF-α measured by ELISA. (<b>d</b>) Cell morphology and CD86 expression evaluated by confocal microscopy (cell nuclei (Hoechst 33342, blue), actin microfilaments (Alexa Flour 488 Phalloidin, green), CD86 (red)).</p>
Full article ">
23 pages, 2764 KiB  
Review
Enzymes from Fishery and Aquaculture Waste: Research Trends in the Era of Artificial Intelligence and Circular Bio-Economy
by Zied Khiari
Mar. Drugs 2024, 22(9), 411; https://doi.org/10.3390/md22090411 - 10 Sep 2024
Viewed by 628
Abstract
In the era of the blue bio-economy, which promotes the sustainable utilization and exploitation of marine resources for economic growth and development, the fisheries and aquaculture industries still face huge sustainability issues. One of the major challenges of these industries is associated with [...] Read more.
In the era of the blue bio-economy, which promotes the sustainable utilization and exploitation of marine resources for economic growth and development, the fisheries and aquaculture industries still face huge sustainability issues. One of the major challenges of these industries is associated with the generation and management of wastes, which pose a serious threat to human health and the environment if not properly treated. In the best-case scenario, fishery and aquaculture waste is processed into low-value commodities such as fishmeal and fish oil. However, this renewable organic biomass contains a number of highly valuable bioproducts, including enzymes, bioactive peptides, as well as functional proteins and polysaccharides. Marine-derived enzymes are known to have unique physical, chemical and catalytic characteristics and are reported to be superior to those from plant and animal origins. Moreover, it has been established that enzymes from marine species possess cold-adapted properties, which makes them interesting from technological, economic and sustainability points of view. Therefore, this review centers around enzymes from fishery and aquaculture waste, with a special focus on proteases, lipases, carbohydrases, chitinases and transglutaminases. Additionally, the use of fishery and aquaculture waste as a substrate for the production of industrially relevant microbial enzymes is discussed. The application of emerging technologies (i.e., artificial intelligence and machine learning) in microbial enzyme production is also presented. Full article
(This article belongs to the Special Issue Enzymes from Marine By-Products and Wastes)
Show Figures

Figure 1

Figure 1
<p>Approximate percentages of waste generated during processing of fish and shellfish: (<b>A</b>) waste generated during processing of fish, and (<b>B</b>) waste generated during processing of shellfish (i.e., shrimps and crabs).</p>
Full article ">Figure 2
<p>Structures of representative proteases: (<b>A</b>) pepsin, (<b>B</b>) trypsin, (<b>C</b>) α-chymotrypsin and (<b>D</b>) pancreatic elastase. The enzyme structures were obtained from the Research Collaboratory for Structural Bioinformatics Protein Data Bank—RCSB PDB (<a href="http://RCSB.org" target="_blank">RCSB.org</a>)—using PDB ID 1AM5 for pepsin from Atlantic cod [<a href="#B44-marinedrugs-22-00411" class="html-bibr">44</a>], PDB ID 1HJ8 for trypsin from Atlantic salmon [<a href="#B45-marinedrugs-22-00411" class="html-bibr">45</a>], PDB ID 4CHA for bovine α-chymotrypsin [<a href="#B46-marinedrugs-22-00411" class="html-bibr">46</a>] and PDB ID 1ELT for elastase from North Atlantic salmon [<a href="#B47-marinedrugs-22-00411" class="html-bibr">47</a>]. The different colors in subfigure (<b>C</b>) represent the different chains in α-chymotrypsin.</p>
Full article ">Figure 3
<p>Structure of a representative transglutaminase. The enzyme structure was obtained from the Research Collaboratory for Structural Bioinformatics Protein Data Bank—RCSB PDB (<a href="http://RCSB.org" target="_blank">RCSB.org</a>)—using PDB ID 1G0D for transglutaminase from red sea bream [<a href="#B165-marinedrugs-22-00411" class="html-bibr">165</a>].</p>
Full article ">
13 pages, 1365 KiB  
Article
Efficient Solution-Phase Dipeptide Synthesis Using Titanium Tetrachloride and Microwave Heating
by Palmira Alessia Cavallaro, Marzia De Santo, Rocco Marinaro, Emilia Lucia Belsito, Angelo Liguori and Antonella Leggio
Int. J. Mol. Sci. 2024, 25(17), 9729; https://doi.org/10.3390/ijms25179729 - 8 Sep 2024
Viewed by 356
Abstract
Microwaves have been successfully employed in the Lewis acid titanium tetrachloride-assisted synthesis of peptide systems. Dipeptide systems with their amino function differently protected with urethane protecting groups have been synthesized in short periods of time and with high yields. The formation of the [...] Read more.
Microwaves have been successfully employed in the Lewis acid titanium tetrachloride-assisted synthesis of peptide systems. Dipeptide systems with their amino function differently protected with urethane protecting groups have been synthesized in short periods of time and with high yields. The formation of the peptide bond between the two reacting amino acids was achieved in pyridine by using titanium tetrachloride as a condensing agent and heating the reaction mixture with a microwave reactor. The reaction conditions are compatible with amino acids featuring various side chains and different protecting groups on both the amino function and side chains. Additionally, the substrates retain their chiral integrity after reaction. Full article
(This article belongs to the Section Physical Chemistry and Chemical Physics)
Show Figures

Figure 1

Figure 1
<p><sup>1</sup>H NMR spectra of <span class="html-italic">N</span>-Boc dipeptides: (<b>A</b>) <span class="html-italic">N</span>-Boc-L-Phe-D-Ala-OCH<sub>3</sub> (<b>2b</b>); (<b>B</b>) <span class="html-italic">N</span>-Boc-D-Phe-D-Ala-OCH<sub>3</sub> (<b>3b</b>); (<b>C</b>) mixture consisting of 60% <b>2b</b> and 40% <b>3b</b>.</p>
Full article ">Scheme 1
<p>Synthesis of <span class="html-italic">N</span>-Fmoc-Phe-Ala-OMe (<b>1a</b>).</p>
Full article ">Scheme 2
<p>Synthesis of <span class="html-italic">N</span>-Fmoc-dipeptide methyl esters <b>2a</b>–<b>6a</b>.</p>
Full article ">Scheme 3
<p>Synthesis of <span class="html-italic">N</span>-Boc-protected dipeptide methyl esters <b>1b</b>–<b>7b</b>.</p>
Full article ">Scheme 4
<p>Solution-phase synthesis of <span class="html-italic">N</span>-Z-protected dipeptide methyl esters <b>1c</b>–<b>3c</b>.</p>
Full article ">
25 pages, 5373 KiB  
Article
SIRT1 Regulates Mitochondrial Damage in N2a Cells Treated with the Prion Protein Fragment 106–126 via PGC-1α-TFAM-Mediated Mitochondrial Biogenesis
by Mengyang Zhao, Jie Li, Zhiping Li, Dongming Yang, Dongdong Wang, Zhixin Sun, Pei Wen, Fengting Gou, Yuexin Dai, Yilan Ji, Wen Li, Deming Zhao and Lifeng Yang
Int. J. Mol. Sci. 2024, 25(17), 9707; https://doi.org/10.3390/ijms25179707 - 7 Sep 2024
Viewed by 440
Abstract
Mitochondrial damage is an early and key marker of neuronal damage in prion diseases. As a process involved in mitochondrial quality control, mitochondrial biogenesis regulates mitochondrial homeostasis in neurons and promotes neuron health by increasing the number of effective mitochondria in the cytoplasm. [...] Read more.
Mitochondrial damage is an early and key marker of neuronal damage in prion diseases. As a process involved in mitochondrial quality control, mitochondrial biogenesis regulates mitochondrial homeostasis in neurons and promotes neuron health by increasing the number of effective mitochondria in the cytoplasm. Sirtuin 1 (SIRT1) is a NAD+-dependent deacetylase that regulates neuronal mitochondrial biogenesis and quality control in neurodegenerative diseases via deacetylation of a variety of substrates. In a cellular model of prion diseases, we found that both SIRT1 protein levels and deacetylase activity decreased, and SIRT1 overexpression and activation significantly ameliorated mitochondrial morphological damage and dysfunction caused by the neurotoxic peptide PrP106–126. Moreover, we found that mitochondrial biogenesis was impaired, and SIRT1 overexpression and activation alleviated PrP106–126-induced impairment of mitochondrial biogenesis in N2a cells. Further studies in PrP106–126-treated N2a cells revealed that SIRT1 regulates mitochondrial biogenesis through the PGC-1α-TFAM pathway. Finally, we showed that resveratrol resolved PrP106–126-induced mitochondrial dysfunction and cell apoptosis by promoting mitochondrial biogenesis through activation of the SIRT1-dependent PGC-1α/TFAM signaling pathway in N2a cells. Taken together, our findings further describe SIRT1 regulation of mitochondrial biogenesis and improve our understanding of mitochondria-related pathogenesis in prion diseases. Our findings support further investigation of SIRT1 as a potential target for therapeutic intervention of prion diseases. Full article
(This article belongs to the Section Molecular Neurobiology)
Show Figures

Figure 1

Figure 1
<p>SIRT1 expression and deacetylase activity are downregulated in PrP<sup>106–126</sup>-exposed N2a cells. (<b>A</b>) Western blots of SIRT1 in N2a cells exposed to 150 μM PrP<sup>106–126</sup>. (<b>B</b>) Quantitation of SIRT1 expression levels shown in (<b>A</b>). (<b>C</b>) Relative SIRT1 deacetylase activity in PrP<sup>106–126</sup>-exposed N2a cells, measured using a SIRT1 assay kit. (<b>D</b>) Relative intracellular NAD+ levels. (<b>E</b>) Relative SIRT1 deacetylase activity measured after 25 μM NAD+ supplementation in PrP<sup>106–126</sup>-exposed N2a cells, measured using a SIRT1 assay kit. Data are expressed as the mean ± SEM. Statistical significance was analyzed via ordinary one-way ANOVA with Tukey’s multiple comparisons test for (<b>B</b>–<b>E</b>). <span class="html-italic">n</span> = at least 3 biologically independent treatments of cells for each. ns, not significant; * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 2
<p>SIRT1 regulates mitochondrial morphological damage and dysfunction in PrP<sup>106–126</sup>-exposed N2a cells. (<b>A</b>) Immunofluorescence images of DsRed-Mito-tagged mitochondria showing differing morphologies following EX527, SRT1720, and PrP<sup>106–126</sup> treatments. Scale bar: zoom out = 5 μm, scale bar: zoom in = 2 μm. (<b>B</b>) Mitochondrial lengths of N2a cells (<span class="html-italic">n</span> = 50) in (<b>A</b>). (<b>C</b>) Mitochondrial aspect ratios of N2a cells (<span class="html-italic">n</span> = 50) in (<b>A</b>). (<b>D</b>) Immunofluorescence images of DsRed-Mito-tagged mitochondria showing differing morphologies after SIRT1 knockdown, overexpression, and PrP<sup>106–126</sup> treatment. Scale bar: zoom out = 10 μm, scale bar: zoom in = 2 μm. (<b>E</b>) Mitochondrial lengths of N2a cells (<span class="html-italic">n</span> = 50) in (<b>D</b>). (<b>F</b>) Mitochondrial aspect ratios of N2a cells (<span class="html-italic">n</span> = 50) in (<b>D</b>). (<b>G</b>) JC-1 was used to quantify changes in the mitochondrial membrane potential (MMP) in EX527-, SRT1720-, and PrP<sup>106–126</sup>-treated N2a cells. Changes in MMP indicate damage to the mitochondria. (<b>H</b>) JC-1 was used to quantify changes in MMP in N2a cells after SIRT1 knockdown, overexpression, and PrP<sup>106–126</sup> treatment. (<b>I</b>) Relative intracellular ATP levels in EX527-, SRT1720-, and PrP<sup>106–126</sup>-treated N2a cells. (<b>J</b>) Relative intracellular ATP levels in EX527-, SRT1720-, and PrP<sup>106–126</sup>-treated N2a cells. Data are expressed as the mean ± SEM. Statistical significance was analyzed via ordinary one-way ANOVA with Tukey’s multiple comparisons test for (<b>B</b>,<b>C</b>,<b>E</b>,<b>F</b>,<b>G</b>,<b>J</b>). <span class="html-italic">n</span> = at least 3 biologically independent treatments/transfections of cells for each. *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 3
<p>SIRT1 is involved in the regulation of mitochondrial biogenesis impairment caused by PrP<sup>106–126</sup> treatment. (<b>A</b>) Relative mitochondrial DNA copy number (mtDNA/nDNA) in EX527- and SRT1720-treated N2a cells incubated with PrP<sup>106–126</sup>. (<b>B</b>) Relative mitochondrial DNA copy number (mtDNA/nDNA) in <span class="html-italic">SIRT1</span> siRNA-transfected and SIRT1-overexpressed N2a cells incubated with PrP<sup>106–126</sup>. (<b>C</b>) The mRNA levels of mitochondrial-encoded genes (<span class="html-italic">MTCO2</span> and <span class="html-italic">MT</span>-<span class="html-italic">Cytb</span>) in EX527- and SRT1720-treated N2a cells incubated with PrP<sup>106–126</sup>. Levels were calculated relative to the <span class="html-italic">GAPDH</span> mRNA level and were normalized to that measured for the control. (<b>D</b>) The mRNA levels of mitochondrial-encoded genes (<span class="html-italic">MTCO2</span> and <span class="html-italic">MT</span>-<span class="html-italic">Cytb</span>) in <span class="html-italic">SIRT1</span> siRNA-transfected and SIRT1-overexpressed N2a cells incubated with PrP<sup>106–126</sup>. The levels were calculated relative to the <span class="html-italic">GAPDH</span> mRNA level and were normalized to that measured for the control. (<b>E</b>) Western blots of SIRT1, mitochondrial-encoded proteins (MTCO2 and MT-Cytb), and nuclear genome-encoded subunits of mitochondrial complexes (SDHA and NDUFB8) in EX527- and SRT1720-treated N2a cells incubated with PrP<sup>106–126</sup>. (<b>F</b>) Quantitation of the results shown in (<b>E</b>). (<b>G</b>) Western blots of SIRT1, mitochondrial-encoded proteins (MTCO2 and MT-Cytb), and nuclear genome-encoded subunits of mitochondrial complexes (SDHA and NDUFB8) in <span class="html-italic">SIRT1</span> siRNA-transfected and SIRT1-overexpressed N2a cells incubated with PrP<sup>106–126</sup>. (<b>H</b>) Quantitation of the results shown in (<b>G</b>). Data are expressed as the mean ± SEM. Statistical significance was analyzed via ordinary one-way ANOVA with Tukey’s multiple comparisons test for (<b>A</b>,<b>B</b>) or two-way ANOVA with Sidak’s multiple comparisons test for (<b>C</b>,<b>D</b>,<b>F</b>,<b>H</b>). <span class="html-italic">n</span> = at least 3 biologically independent treatments/transfections of cells for each. * <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 4
<p>TFAM is required for SIRT1-mediated mitochondrial biogenesis in PrP<sup>106–126</sup>-treated N2a cells. (<b>A</b>) Relative mitochondrial DNA copy numbers (mtDNA/nDNA) in TFAM siRNA-transfected and TFAM-overexpressed N2a cells incubated with PrP<sup>106–126</sup>. (<b>B</b>) The mRNA levels of mitochondrial-encoded genes (<span class="html-italic">MTCO2</span> and <span class="html-italic">MT</span>-<span class="html-italic">Cytb</span>) in <span class="html-italic">TFAM</span> siRNA-transfected and TFAM-overexpressed N2a cells incubated with PrP<sup>106–126</sup>. (<b>C</b>) Western blots of TFAM, mitochondrial-encoded proteins (MTCO2 and MT-Cytb), and nuclear genome-encoded subunits of mitochondrial complexes (SDHA and NDUFB8) in <span class="html-italic">TFAM</span> siRNA-transfected and TFAM-overexpressed N2a cells incubated with PrP<sup>106–126</sup>. (<b>D</b>) Quantitation of the results shown in (<b>C</b>). (<b>E</b>) <span class="html-italic">TFAM</span> mRNA levels in EX527- and SRT1720-treated N2a cells incubated with PrP<sup>106–126</sup>. (<b>F</b>) Western blots showing TFAM levels in EX527- and SRT1720-treated N2a cells incubated with PrP<sup>106–126</sup>. (<b>G</b>) Quantitation of the results shown in (<b>F</b>). (<b>H</b>) <span class="html-italic">TFAM</span> mRNA levels in <span class="html-italic">SIRT1</span> siRNA-transfected and SIRT1-overexpressed N2a cells incubated with PrP<sup>106–126</sup>. (<b>I</b>) Western blots showing the TFAM levels in <span class="html-italic">SIRT1</span> siRNA-transfected and SIRT1-overexpressed N2a cells incubated with PrP<sup>106–126</sup>. (<b>J</b>) Quantitation of the results shown in (<b>I</b>). For all mRNA measurements, the levels were calculated relative to the <span class="html-italic">GAPDH</span> mRNA level and were normalized to that measured for the control. Data are expressed as the mean ± SEM. Statistical significance was analyzed via ordinary one-way ANOVA with Tukey’s multiple comparisons test for (<b>A</b>,<b>E</b>,<b>G</b>,<b>H</b>,<b>J</b>) or two-way ANOVA with Sidak’s multiple comparisons test for (<b>B</b>,<b>D</b>). <span class="html-italic">n</span> = at least 3 biologically independent treatments/transfections of cells for each. * <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 5
<p>PGC-1α participates in SIRT1-mediated mitochondrial biogenesis in PrP<sup>106–126</sup>-treated N2a cells. (<b>A</b>) Relative mitochondrial DNA copy numbers (mtDNA/nDNA) in <span class="html-italic">PGC-1α</span> siRNA-transfected and PGC-1α-overexpressed N2a cells incubated with PrP<sup>106–126</sup>. (<b>B</b>) The mRNA levels for <span class="html-italic">TFAM</span> and mitochondrial-encoded genes (<span class="html-italic">MTCO2</span> and <span class="html-italic">MT</span>-<span class="html-italic">Cytb</span>) in <span class="html-italic">PGC-1α</span> siRNA-transfected and PGC-1α-overexpressed N2a cells incubated with PrP<sup>106–126</sup>. (<b>C</b>) Western blots of PGC-1α, TFAM, mitochondrial-encoded proteins (MTCO2 and MT-Cytb), and nuclear genome-encoded subunits of mitochondrial complexes (SDHA and NDUFB8) in <span class="html-italic">PGC-1α</span> siRNA-transfected and PGC-1α-overexpressed N2a cells incubated with PrP<sup>106–126</sup>. (<b>D</b>) Quantitation of the results shown in (<b>C</b>). (<b>E</b>) Western blots of PGC-1α in EX527- and SRT1720-treated N2a cells incubated with PrP<sup>106–126</sup>. (<b>F</b>) Quantitation of the results shown in (<b>E</b>). (<b>G</b>) Western blot of PGC-1α in <span class="html-italic">SIRT1</span> siRNA-transfected and SIRT1-overexpressed N2a cells incubated with PrP<sup>106–126</sup>. (<b>H</b>) Quantitation of the results shown in (<b>G</b>). For all mRNA measurements, the levels were calculated relative to the <span class="html-italic">GAPDH</span> mRNA level and were normalized to that measured for the control. Data are expressed as the mean ± SEM. Statistical significance was analyzed via ordinary one-way ANOVA with Tukey’s multiple comparisons test for (<b>A</b>,<b>F</b>,<b>H</b>) or two-way ANOVA with Sidak’s multiple comparisons test for (<b>B</b>,<b>D</b>). <span class="html-italic">n</span> = at least 3 biologically independent treatments/transfections of cells for each. * <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 6
<p>Knockdown of PGC-1α or TFAM negated the beneficial effects of SIRT1 overexpression and activation on mitochondrial biogenesis in PrP<sup>106–126</sup>-treated N2a cells. (<b>A</b>) Relative mitochondrial copy numbers (mtDNA/nDNA). (<b>B</b>) The relative mRNA levels for <span class="html-italic">TFAM</span> and the mitochondrial-encoded genes <span class="html-italic">MTCO2</span> and <span class="html-italic">MT</span>-<span class="html-italic">Cytb</span>. (<b>C</b>) Western blots of PGC-1α, TFAM, MTCO2, MT-Cytb, SDHA, and NDUFB8 in <span class="html-italic">PGC-1α</span> siRNA-transfected and SIRT1-overexpressed or -activated N2a cells incubated with PrP<sup>106–126</sup>. (<b>D</b>) Quantitation of the results shown in (<b>C</b>). (<b>E</b>) Western blots of TFAM, MTCO2, MT-Cytb, SDHA, and NDUFB8 in <span class="html-italic">TFAM</span> siRNA-transfected and SIRT1-overexpressed or -activated N2a cells incubated with PrP<sup>106–126</sup>. (<b>F</b>) Quantitation of the results shown in (<b>E</b>). The experiments shown in (<b>A</b>,<b>B</b>) include the following treatment groups: control, PrP<sup>106–126</sup> incubation, SIRT1-overexpressed and PrP<sup>106–126</sup> incubation, SRT1720 and PrP<sup>106–126</sup> co-treatment, SIRT1-overexpressed and PrP<sup>106–126</sup> incubation after PGC-1α knockdown, SRT1720 and PrP<sup>106–126</sup> co-treatment after PGC-1α knockdown, SIRT1-overexpressed and PrP<sup>106–126</sup> incubation after TFAM knockdown, and SRT1720 and PrP<sup>106–126</sup> co-treatment after TFAM knockdown. All mRNA levels were calculated relative to the <span class="html-italic">GAPDH</span> mRNA level and were normalized to that measured for the control. Data are expressed as the mean ± SEM. Statistical significance was analyzed via ordinary one-way ANOVA with Tukey’s multiple comparisons test for (<b>A</b>) or two-way ANOVA with Sidak’s multiple comparisons test. for (<b>B</b>,<b>D</b>,<b>F</b>); <span class="html-italic">n</span> = at least 3 biologically independent treatments/transfections of cells for each. * <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 7
<p>RSV ameliorates PrP<sup>106–126</sup>-induced mitochondrial dysfunction by activating SIRT1-dependent mitochondrial biogenesis in N2a cells. (<b>A</b>) Relative SIRT1 deacetylase activity. (<b>B</b>) Relative intracellular ATP levels. (<b>C</b>) JC-1 was used to quantify changes in the mitochondrial membrane potential (MMP), which indicates damage to the mitochondria. (<b>D</b>) Immunofluorescence images of DsRed-Mito-tagged mitochondria showing differing morphologies. Scale bar: zoom out = 5 μm, scale bar: zoom in = 2 μm. (<b>E</b>) Mitochondrial length of the N2a cells (<span class="html-italic">n</span> = 50) in (<b>D</b>). (<b>F</b>) Mitochondrial aspect ratio of the N2a cells (<span class="html-italic">n</span> = 50) in (<b>D</b>). (<b>G</b>) Relative mitochondrial copy numbers (mtDNA/nDNA). (<b>H</b>) The relative mRNA levels for <span class="html-italic">TFAM</span> and the mitochondrial-encoded genes <span class="html-italic">MTCO2</span> and <span class="html-italic">MT</span>-<span class="html-italic">Cytb</span>. (<b>I</b>) Western blots of SIRT1, PGC-1α, TFAM, mitochondrial-encoded proteins (MTCO2 and MT-Cytb), and nuclear genome-encoded subunits of mitochondrial complexes (SDHA and NDUFB8). (<b>J</b>) Quantitation of the results shown in (<b>I</b>). All of the experiments shown in this figure include the following treatment groups: control, PrP<sup>106–126</sup> incubation, RSV supplement, RSV and PrP<sup>106–126</sup> co-treatment, and RSV and PrP<sup>106–126</sup> co-treatment after SIRT1 knockdown. All the mRNA levels were calculated relative to the <span class="html-italic">GAPDH</span> mRNA level and were normalized to that measured for the control. Data are expressed as the mean ± SEM. Statistical significance was analyzed via ordinary one-way ANOVA with Tukey’s multiple comparisons test for (<b>A</b>–<b>C</b>,<b>E</b>–<b>G</b>) or two-way ANOVA with Sidak’s multiple comparisons test for (<b>H</b>,<b>J</b>). <span class="html-italic">n</span> = at least 3 biologically independent treatments/transfections of cells for each. ns, not significant; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 8
<p>RSV supplementation attenuates PrP<sup>106–126</sup>-induced N2a cell apoptosis. (<b>A</b>) Cell viability was assayed using the cell counting CCK-8 kit in EX527-, SRT1720-, and PrP<sup>106–126</sup>-treated N2a cells. (<b>B</b>) Cell viability was assayed using the cell counting CCK-8 kit in N2a cells after SIRT1 knockdown, overexpression, and PrP<sup>106–126</sup> treatment. (<b>C</b>) Cell viability was assayed using the cell counting CCK-8 kit in N2a cells after PGC-1α knockdown, overexpression, and PrP<sup>106–126</sup> treatment. (<b>D</b>) Cell viability was assayed using the cell counting CCK-8 kit in N2a cells after TFAM knockdown, overexpression, and PrP<sup>106–126</sup> treatment. (<b>E</b>) Cell viability was assayed using the cell counting CCK-8 kit in control or treated N2a cells. (<b>F</b>) Confocal images of TUNEL and DAPI staining in N2a cells. Scale bar: 50 μm. (<b>G</b>) Proportion of apoptotic cells as quantified by TUNEL staining. (<b>H</b>) Western blots of caspase-9, cleaved caspase-9, and cleaved caspase-3 to identify the rates of apoptosis. (<b>I</b>–<b>K</b>) Quantitation of the results shown in (<b>H</b>). The experiments shown in (<b>E</b>–<b>K</b>) include the following treatment groups: control, PrP<sup>106–126</sup> incubation, RSV supplement, RSV and PrP<sup>106–126</sup> co-treatment, and RSV and PrP<sup>106–126</sup> co-treatment after SIRT1 knockdown. Data are expressed as the mean ± SEM. Statistical significance was analyzed via ordinary one-way ANOVA with Tukey’s multiple comparisons test for (<b>A</b>–<b>E</b>,<b>G</b>,<b>I</b>–<b>K</b>). <span class="html-italic">n</span> = at least 3 biologically independent treatments/transfections of cells for each. * <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 ">
17 pages, 9841 KiB  
Article
Elucidating the Substrate Envelope of Enterovirus 68-3C Protease: Structural Basis of Specificity and Potential Resistance
by Vincent N. Azzolino, Ala M. Shaqra, Akbar Ali, Nese Kurt Yilmaz and Celia A. Schiffer
Viruses 2024, 16(9), 1419; https://doi.org/10.3390/v16091419 - 5 Sep 2024
Viewed by 392
Abstract
Enterovirus-D68 (EV68) has emerged as a global health concern over the last decade with severe symptomatic infections resulting in long-lasting neurological deficits and death. Unfortunately, there are currently no FDA-approved antiviral drugs for EV68 or any other non-polio enterovirus. One particularly attractive class [...] Read more.
Enterovirus-D68 (EV68) has emerged as a global health concern over the last decade with severe symptomatic infections resulting in long-lasting neurological deficits and death. Unfortunately, there are currently no FDA-approved antiviral drugs for EV68 or any other non-polio enterovirus. One particularly attractive class of potential drugs are small molecules inhibitors, which can target the conserved active site of EV68-3C protease. For other viral proteases, we have demonstrated that the emergence of drug resistance can be minimized by designing inhibitors that leverage the evolutionary constraints of substrate specificity. However, the structural characterization of EV68-3C protease bound to its substrates has been lacking. Here, we have determined the substrate specificity of EV68-3C protease through molecular modeling, molecular dynamics (MD) simulations, and co-crystal structures. Molecular models enabled us to successfully characterize the conserved hydrogen-bond networks between EV68-3C protease and the peptides corresponding to the viral cleavage sites. In addition, co-crystal structures we determined have revealed substrate-induced conformational changes of the protease which involved new interactions, primarily surrounding the S1 pocket. We calculated the substrate envelope, the three-dimensional consensus volume occupied by the substrates within the active site. With the elucidation of the EV68-3C protease substrate envelope, we evaluated how 3C protease inhibitors, AG7088 and SG-85, fit within the active site to predict potential resistance mutations. Full article
Show Figures

Figure 1

Figure 1
<p>EV68-3C protease substrates and structure. (<b>A</b>) EV68 polyprotein prior to cleavage by the 3C protease into structural and non-structural viral proteins. (<b>B</b>) Crystal structure of EV68-3C protease in surface representation bound to a peptidomimetic inhibitor (SG-85, in magenta sticks) with cysteine protease’s catalytic dyad residues (H40 and C147) highlighted in yellow (PDB ID: 3ZVF). The inhibitor is labeled with corresponding P5 to P1′ moieties. (<b>C</b>) Amino acid sequences of EV68-3C protease cleavage sites in the viral polyprotein, with a conserved glutamine (Q) in the P1 position, generally followed by a glycine (G) at P1′. The sequences are highly diverse distal to the cut site (denoted by the blue vertical bar).</p>
Full article ">Figure 2
<p>Hydrogen bond network between the substrate peptides and the active site of EV68-3C protease. (<b>A</b>) The EV68-3C protease in grey surface representation, focusing on the active site pocket with the catalytic dyad highlighted in yellow, and the end point modeled 3C3D peptide (8 mer) post MD simulation shown in blue sticks. Hydrogen bonds between the protein and peptide heavy atoms are depicted with green dashed lines. (<b>B</b>) Percent frequency of the hydrogen bonds between EV68-3C protease and modeled substrate peptides, as well as the 3B3C peptide co-crystal structure (PDB ID: 9AX9), during MD simulation trajectories. The bonds present during the simulation with higher frequency interactions are denoted in darker red and lower frequency interactions are in white.</p>
Full article ">Figure 3
<p>The substrate envelope of EV68-3C protease. A heatmap representation of the calculated dynamic substrate envelope for EV68-3C protease shows the percent occupancy of viral substrates within this three-dimensional volume. The peptide backbone is most well conserved between the P4 and P2′ positions and has more variability at P5 and at P3′. The percent occupancy ranges from lowest (dark blue) to highest (red) of finding atoms at that position within a given viral peptide.</p>
Full article ">Figure 4
<p>The experimental crystal structure of EV68-3C protease with a substrate peptide bound at the active site. (<b>A</b>) The EV68-3C protease (grey surface) co-crystal structure with the 3B3C peptide (purple sticks) in the electron density, with nearby crystallographic waters (red spheres) and hydrogen bond interactions (green dashed lines) with the surrounding protease residues (dark grey sticks). The mesh depicts the electron density (2F0-FC map) for the 3B3C peptide at the active site. (<b>B</b>) The EV68-3C protease (grey surface) and 3B3C peptide (purple sticks) co-crystal structure shown with the superimposed substrate envelope denoted by the cyan volume. The peptide, derived from an experimental protein crystal structure, fits completely within the calculated envelope from molecular modeling and MD simulations.</p>
Full article ">Figure 5
<p>Comparison of peptide-bound and apo EV68-3V protease crystal structures. Surface representation of (<b>A</b>) 3B3C-peptide-bound (PDB ID: 9AX9) and (<b>B</b>) apo (PDB ID: 8FL5) protease colored according to variation of backbone, assessed by the average difference in distance between the C-alpha atoms of residues (in Å) between the two structures. (<b>C</b>) Superposition of the two structures in stick representation with red arrows pointing to the largest changes in key residues between the apo (cyan) and the substrate-bound (pink) co-crystal structures. The 3B3C peptide is depicted as purple sticks.</p>
Full article ">Figure 6
<p>(<b>A</b>) Rupintrivir (AG7088), in slate blue, covalently bound within the active site of EV68-3C protease in grey (PDB: 7L8H). Hydrogen bonds between rupintrivir and the protein are highlighted as yellow dashed lines. The substrate envelope is shown in light blue (<b>B</b>) SG-85, in orange, covalently bound within the active site of EV68-3C protease (PDB: 3ZVF).</p>
Full article ">
17 pages, 7077 KiB  
Article
Focal Cerebral Ischemia Induces Expression of Glutaminyl Cyclase along with Downstream Molecular and Cellular Inflammatory Responses
by Corinna Höfling, Luise Ulrich, Sina Burghardt, Philippa Donkersloot, Michael Opitz, Stefanie Geissler, Stephan Schilling, Holger Cynis, Dominik Michalski and Steffen Roßner
Cells 2024, 13(17), 1412; https://doi.org/10.3390/cells13171412 - 23 Aug 2024
Viewed by 484
Abstract
Glutaminyl cyclase (QC) and its isoenzyme (isoQC) catalyze the formation of N-terminal pyroglutamate (pGlu) from glutamine on a number of neuropeptides, peptide hormones and chemokines. Chemokines of the C-C ligand (CCL) motif family are known to contribute to inflammation in neurodegenerative conditions. Here, [...] Read more.
Glutaminyl cyclase (QC) and its isoenzyme (isoQC) catalyze the formation of N-terminal pyroglutamate (pGlu) from glutamine on a number of neuropeptides, peptide hormones and chemokines. Chemokines of the C-C ligand (CCL) motif family are known to contribute to inflammation in neurodegenerative conditions. Here, we used a model of transient focal cerebral ischemia to explore functional, cellular and molecular responses to ischemia in mice lacking genes for QC, isoQC and their substrate CCL2. Mice of the different genotypes were evaluated for functional consequences of stroke, infarct volume, activation of glia cells, and for QC, isoQC and CCL2 expression. The number of QC-immunoreactive, but not of isoQC-immunoreactive, neurons increased robustly in the infarct area at 24 and 72 h after ischemia. In parallel, immunohistochemical signals for the QC substrate CCL2 increased from 24 to 72 h after ischemia induction without differences between genotypes analyzed. The increase in CCL2 was accompanied by morphological activation of Iba1-immunoreactive microglia and recruitment of MHC-II-positive cells at 72 h after ischemia. Among other chemokines quantified in the brain tissue, CCL17 showed higher concentrations at 72 h compared to 24 h after ischemia. Collectively, these data suggest a critical role for QC in inflammatory processes in the stroke-affected brain. Full article
(This article belongs to the Section Cells of the Nervous System)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Characterization of focal cerebral ischemia in knock-out mouse models. Mice of different genotypes (wild-type, CCL2 KO, QC KO and isoQC KO) were subjected to transient focal cerebral ischemia for 30 min, followed by 24 h and 72 h observation periods. (<b>A</b>) Survival rate (<span class="html-italic">p</span> = 0.858), Menzies score (<span class="html-italic">p</span> = 0.91), overall physical condition (<span class="html-italic">p</span> = 0.82) and body weight (<span class="html-italic">p</span> = 0.78) did not differ significantly between genotypes. Only 45% (wild type) to 58% (QC KO) of the animals subjected to the operation procedure survived for 72 h. For all genotypes, a constant Menzies score of 3 was determined at all time points. In contrast, the overall physical condition steadily declined (higher scores indicating worse condition) during the 72 h post-surgery period, which was also characterized by a weight loss ranging between 7% in wild-type mice and 16% in isoQC KO mice. (<b>B</b>) Immunohistochemical NeuN + HuC/D images from experimental animals of all genotypes at 24 h and 72 h after onset of ischemia. The dashed lines indicate the infarct area identified by diminished NeuN + HuC/D labeling. The infarct volume was calculated from serial brain slices for all animals of all genotypes and was found to be between 18 mm<sup>3</sup> (CCL2 KO, 24 h) and 35 mm<sup>3</sup> (QC KO, 24 h). (<b>C</b>) As a complementary measure of infarct size, increased immunosignals for Neurofilament L (NFL) in the ischemic area was quantified in serial sections. Here, the smallest ischemic areas were detected for CCL2 KO mice at 24 h and largest for QC KO at 72 h after ischemia. Note the complementary loss of NeuN + HuC/D and induction of NFL immunoreactivity in the ischemic area (I).</p>
Full article ">Figure 2
<p>Expression of QC and isoQC in ischemic brain areas. (<b>A</b>) Immunohistochemical labeling for QC and isoQC in the wild-type mouse brain at 24 h and 72 h after ischemia. (<b>B</b>) There were significant increases in the numbers of QC-immunoreactive neurons in infarct areas identified by NeuN + HuC/D labeling in consecutive brain sections for all genotypes (except QC KO) at both survival time points. (<b>C</b>) In contrast, there was significantly reduced isoQC immunoreactivity in infarct areas compared to the non-ischemic hemisphere. * <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; differences statistically significant versus control.</p>
Full article ">Figure 3
<p>Quantification of CCL2 expression and co-localization to neuronal and glial markers. (<b>A</b>) Left: single immunohistochemical labeling of CCL2 in the wild-type (WT) mouse brain 24 h and 72 h after ischemia. Note the absence of CCL2 immunoreactivity in a CCL2 KO mouse brain 72 h after ischemia. Right: note the post-ischemia time-dependent increase in signal intensity and spreading of the CCL2 immunosignal beyond the infarct (I) region demarked by the dashed line related to reduced NeuN + HuC/D labeling at 72 h. (<b>B</b>) Quantification of CCL2 immunosignals demonstrating increased CCL2 in the wild-type (WT) mouse brain at 24 and 72 h. Note that the mean values for CCL2 signals in QC KO and in isoQC KO are only half of the of the WT mice at both time points. (<b>C</b>) Appearance of CCL2 immunoreactivity in the ischemic cortex as extracellular spots, in association with vessels and in a glia-like shape. (<b>D</b>) Examples of double labeling of CCL2 with microglia markers Iba1, CD68 and TMEM119, with neurons (NeuN + HuC/D), astrocytes (GFAP) and oligodendrocytes (Olig2) in the ischemic cortex. Note the presence of CCL2 immunosignals in subsets of these neuronal and glial populations (arrows) except in oligodendrocytes. ** <span class="html-italic">p</span> &lt; 0.01; differences statistically significant versus control.</p>
Full article ">Figure 4
<p>Glia cell activation/recruitment in ischemic brain regions. (<b>A</b>) Triple immunofluorescent labeling of the microglial marker Iba1 and the astrocyte marker GFAP in combination with NeuN + HuC/D to identify the cortical infarct area (I). In the high magnification images (bottom), the activation of Iba1-positive microglia in the ischemic region and the presence of GFAP-immunoreactive reactive astrocytes in the border zone is evident. (<b>B</b>) Quantification of GFAP immunosignals in brain sections of mice of different genotypes. Note the instant increase in GFAP expression at 24 h for all genotypes, which is not increase further at 72 h. (<b>C</b>,<b>D</b>) In contrast, both numbers of activated microglia (<b>C</b>) and of MHC-II cells (<b>D</b>) only increase at 72 h after ischemia for all genotypes. Note that the mean values for both cell types in QC KO mice and in isoQC KO mice are only half of that in wild-type (WT) mice. Arrows in (<b>D</b>) point towards MHC-II-positive neurons (black). (<b>E</b>,<b>F</b>) The immunoreactivity for CD68 is increased after ischemia in all genotypes (<b>E</b>), whereas TMEM119 immunosignals are reduced (<b>F</b>). * <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; differences statistically significant versus control. # <span class="html-italic">p</span> &lt; 0.05; ## <span class="html-italic">p</span> &lt; 0.01; differences statistically significant at 72 h versus 24 h.</p>
Full article ">Figure 5
<p>Quantification of chemokines in brain tissue and plasma of mice after ischemia. The concentrations of CCL2, CX3CL1, IL-1β, TNF-α and CCL17 were quantified by multiplex analyses in brain tissue of the ischemic hemisphere (<b>left</b>) and in plasma (<b>right</b>) at 24 h and 72 h after ischemia as indicated. Upregulation in the brain tissue is only selectively mirrored in the plasma, e.g., in case of CCL2. Please note the lack of CCL2 in CCL2 KO mice. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; differences statistically significant versus control. ## <span class="html-italic">p</span> &lt; 0.01; differences statistically significant at 72 h versus 24 h.</p>
Full article ">
19 pages, 6638 KiB  
Article
Extracellular Vesicles of the Probiotic Escherichia coli Nissle 1917 Reduce PepT1 Levels in IL-1β-Treated Caco-2 Cells via Upregulation of miR-193a-3p
by Yenifer Olivo-Martínez, Sergio Martínez-Ruiz, Cecilia Cordero, Josefa Badia and Laura Baldoma
Nutrients 2024, 16(16), 2719; https://doi.org/10.3390/nu16162719 - 15 Aug 2024
Viewed by 746
Abstract
PepT1, a proton-coupled oligopeptide transporter, is crucial for intestinal homeostasis. It is mainly expressed in small intestine enterocytes, facilitating the absorption of di/tri-peptides from dietary proteins. In the colon, PepT1 expression is minimal to prevent excessive responses to proinflammatory peptides from the gut [...] Read more.
PepT1, a proton-coupled oligopeptide transporter, is crucial for intestinal homeostasis. It is mainly expressed in small intestine enterocytes, facilitating the absorption of di/tri-peptides from dietary proteins. In the colon, PepT1 expression is minimal to prevent excessive responses to proinflammatory peptides from the gut microbiota. However, increased colonic PepT1 is linked to chronic inflammatory diseases and colitis-associated cancer. Despite promising results from animal studies on the benefits of extracellular vesicles (EVs) from beneficial gut commensals in treating IBD, applying probiotic EVs as a postbiotic strategy in humans requires a thorough understanding of their mechanisms. Here, we investigate the potential of EVs of the probiotic Nissle 1917 (EcN) and the commensal EcoR12 in preventing altered PepT1 expression under inflammatory conditions, using an interleukin (IL)-1-induced inflammation model in Caco-2 cells. The effects are evaluated by analyzing the expression of PepT1 (mRNA and protein) and miR-193a-3p and miR-92b, which regulate, respectively, PepT1 mRNA translation and degradation. The influence of microbiota EVs on PepT1 expression is also analyzed in the presence of bacterial peptides that are natural substrates of colonic PepT1 to clarify how the regulatory mechanisms function under both physiological and pathological conditions. The main finding is that EcN EVs significantly decreases PepT1 protein via upregulation of miR-193a-3p. Importantly, this regulatory effect is strain-specific and only activates in cells exposed to IL-1β, suggesting that EcN EVs does not control PepT1 expression under basal conditions but can play a pivotal role in response to inflammation as a stressor. By this mechanism, EcN EVs may reduce inflammation in response to microbiota in chronic intestinal disorders by limiting the uptake of bacterial proinflammatory peptides. Full article
(This article belongs to the Special Issue Probiotics and Their Metabolites in Human Health)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Regulation of PepT1 expression by EcN or EcoR12 EVs in the in vitro IL-1β-Inflammation model. Caco-2 cells were incubated with EVs (60 µg/mL) from EcN or EcoR12 for 3 h and then challenged with IL-1β (10 ng/mL) or the vehicle (PBS) for 48 h. In parallel, cells treated with EVs (60 µg/mL) from EcN or EcoR12 were incubated in the absence of IL-1β as a control. (<b>A</b>) Relative mRNA levels of PepT1 were assessed by RT-qPCR using GAPDH as the reference gene. Data are presented as mean ± SEM from 3 independent experiments. (<b>B</b>,<b>C</b>) Quantification of PepT1 protein levels by ELISA (<b>B</b>) and by immunofluorescence confocal microscopy (<b>C</b>). Representative confocal maximal projection images of cells treated as indicated are shown in Fire LUT after image processing. The calibration bar of Fire LUT intensity is shown in the left. Scale bar: 20 µm. Quantification of the PepT1 mean intensity is shown for each treatment on the right side. Data are given as mean ± SEM of arbitrary intensity units (AU) (<span class="html-italic">n</span> = 3 independent biological replicates). Statistical differences were assessed with one-way ANOVA, followed by post hoc Tukey’s. * <span class="html-italic">p</span> ≤ 0.05, ** <span class="html-italic">p</span> ≤ 0.01 vs. control untreated cells (white bars), # <span class="html-italic">p</span> ≤ 0.05, ## <span class="html-italic">p</span> ≤ 0.01, ### <span class="html-italic">p</span> ≤ 0.001 vs. IL-1β treated cells (black bars), <span>$</span> <span class="html-italic">p</span> ≤ 0.05 between cells stimulated with EcN or EcoR12 EVs.</p>
Full article ">Figure 2
<p>Regulation of miR-193a-3p (<b>A</b>) and miR-92b (<b>B</b>) by EcN or EcoR12 EVs in the IL-1β-Inflammation model. Caco-2 cells were incubated with EVs (60 µg/mL) from EcN or EcoR12 for 3 h and then challenged with IL-1β (10 ng/mL) or the vehicle (PBS) for 48 h. In parallel, cells treated with EVs (60 µg/mL) from EcN or EcoR12 were incubated in the absence of IL-1β as a control. Relative expression levels of the indicated miRNAs were quantified by RT-qPCR and normalized to the U6 reference gene. Data are expressed as mean ± SEM from three independent experiments. Differences were evaluated with one-way ANOVA, followed by post hoc Tukey’s. * <span class="html-italic">p</span> ≤ 0.05 vs. untreated control cells (white bars); ## <span class="html-italic">p</span> ≤ 0.01 vs. IL-1β treated cells (black bars), and <span>$</span><span>$</span> <span class="html-italic">p</span> ≤ 0.01 between cells stimulated with EcN or EcoR12 EVs.</p>
Full article ">Figure 3
<p>Setting up the experimental conditions for the Tri-DAP model. (<b>A</b>) Effect of the Tri-DAP concentration on cell viability assessed by the MTT assay. Caco-2 cell monolayers were exposed to Tri-DAP at the final concentration of 1 or 5 μg/mL for 24 h. (<b>B</b>) Influence of the Tri-DAP concentration on the expression of genes encoding the pro-inflammatory cytokines IL-8 and TNF-α. Caco-2 cells were exposed to Tri-DAP at the indicated concentrations for 8 and 24 h. Untreated cells were incubated in parallel as a control (white bars). The relative mRNA levels of the indicated genes were determined by RT-qPCR using GAPDH as the reference gene. Data are expressed as mean ± SEM from three independent experiments. Differences were evaluated with one-way ANOVA, followed by post hoc Tukey’s. ** <span class="html-italic">p</span> ≤ 0.01 vs. untreated control cells.</p>
Full article ">Figure 4
<p>Regulation of PepT1 expression by EcN or EcoR12 EVs in the Tri-DAP induction model. Caco-2 cells were incubated with EVs (60 µg/mL) from EcN or EcoR12 for 3 h and then treated with Tri-DAP (5 µg/mL) or the vehicle (PBS) for 8 h. In parallel, Caco-2 cells treated with EVs (60 µg/mL) from EcN or EcoR12 were incubated in the absence of Tri-DAP as a control. (<b>A</b>) Relative mRNA levels of PepT1 were measured by RT-qPCR using GAPDH as the reference gene. Data are expressed as mean ± SEM from 3 independent experiments. (<b>B</b>,<b>C</b>) Quantification of PepT1 protein levels by ELISA (<b>B</b>) and by immunofluorescence confocal microscopy (<b>C</b>). Representative confocal maximal projection images of cells treated as indicated are shown in Fire LUT after image processing. The calibration bar of Fire LUT intensity is shown in the left. Scale bar: 20 µm. Quantification of the PepT1 mean intensity is shown for each treatment on the right side. Data are shown as mean ± SEM of arbitrary intensity units (AU) (n = 3 independent biological replicates). Statistical differences were assessed with one-way ANOVA, followed by post hoc Tukey’s. * <span class="html-italic">p</span> ≤ 0.05 vs. untreated control cells (white bars); # <span class="html-italic">p</span> ≤ 0.05 vs. Tri-DAP treated cells (blue bars).</p>
Full article ">Figure 5
<p>Regulation of miR-193a-3p (<b>A</b>) and miR-92b (<b>B</b>) by EcN or EcoR12 EVs in the Tri-DAP induction model. Caco-2 cells were incubated with EVs (60 µg/mL) from EcN or EcoR12 for 3 h and then treated with Tri-DAP (5 µg/mL) or the vehicle (PBS) for 8 h. In parallel, Caco-2 cells treated with EVs (60 µg/mL) from EcN or EcoR12 were incubated in the absence of Tri-DAP as a control. Relative expression levels of the indicated miRNAs were measured by RT-qPCR and normalized to the U6 reference gene. Data are expressed as mean ± SEM from three independent experiments. Differences were evaluated with one-way ANOVA, followed by post hoc Tukey’s. * <span class="html-italic">p</span> ≤ 0.05 vs. untreated control cells (white bars); # <span class="html-italic">p</span> ≤ 0.05 vs. Tri-DAP treated cells (blue bars).</p>
Full article ">Figure 6
<p>Expression analysis of proinflammatory cytokines in cells stimulated with Tri-DAP in the absence or presence of EcN Evs or EcoR12 Evs. Tri-DAP stimulation conditions: Caco-2 cells were incubated with Evs (60 µg/mL) from EcN or EcoR12 for 3 h and then treated with Tri-DAP (5 µg/mL) for 8 h (blue and dashed bars). Control conditions: Caco-2 cells were treated with Evs (60 µg/mL) from EcN or EcoR12 and incubated for 8 h (white and gray bars). Relative mRNA levels of the indicated cytokines were measured by RT-qPCR using GAPDH as the reference gene (upper panels), and secreted levels of IL-8 and TNF-α were quantified by ELISA in the culture supernatants. In all panels, data are presented as mean ± SEM from three independent experiments. Differences were evaluated with one-way ANOVA, followed by post hoc Tukey’s. * <span class="html-italic">p</span> ≤ 0.05, ** <span class="html-italic">p</span> ≤ 0.01, and *** <span class="html-italic">p</span> ≤ 0.001 vs. control cells.</p>
Full article ">
14 pages, 2705 KiB  
Article
Cloning, Expression, and Characterization of a Metalloprotease from Thermophilic Bacterium Streptomyces thermovulgaris
by Amna Mushtaq, Sibtain Ahmed, Tahir Mehmood, Jorge Cruz-Reyes, Amer Jamil and Shafaq Nawaz
Biology 2024, 13(8), 619; https://doi.org/10.3390/biology13080619 - 15 Aug 2024
Viewed by 539
Abstract
Proteases hydrolyze proteins and reduce them to smaller peptides or amino acids. Besides many biological processes, proteases play a crucial in different industrial applications. A 792 bp protease gene (nprB) from the thermophilic bacterium Streptomyces thermovulgaris was cloned and expressed in [...] Read more.
Proteases hydrolyze proteins and reduce them to smaller peptides or amino acids. Besides many biological processes, proteases play a crucial in different industrial applications. A 792 bp protease gene (nprB) from the thermophilic bacterium Streptomyces thermovulgaris was cloned and expressed in E. coli BL21 using pET 50b (+). Optimal recombinant protease expression was observed at 1 mM IPTG, 37 °C for 4 h. The resulting protease was observed in soluble form. The molecular mass estimated by SDS-PAGE and Western blot analysis of the protease (NprB) fused with His and Nus tag is ~70 KDa. The protease protein was purified by Ammonium sulfate precipitation and immobilized metal ion affinity chromatography. The optimum pH and temperature for protease activity using casein as substrate were 7.2 and 70 °C, respectively. The mature protease was active and retained 80% of its activity in a broad spectrum of pH 6–8 after 4 h of incubation. Also, the half-life of the protease at 70 °C was 4 h. EDTA (5 mM) completely inhibited the enzyme, proving the isolated protease was a metalloprotease. NprB activity was enhanced in the presence of Zn2+, Mn2+, Fe2+ and Ca2+, while Hg2+ and Ni2+ decreased its activity. Exposure to organic solvents did not affect the protease activity. The recombinant protease was stable in the presence of 10% organic solvents and surfactants. Further characterization showed that zinc-metalloprotease is promising for the detergent, laundry, leather, and pharmaceutical industries. Full article
(This article belongs to the Section Biochemistry and Molecular Biology)
Show Figures

Figure 1

Figure 1
<p>SDS-PAGE for the protein expression in <span class="html-italic">E. coli</span> BL21 (Rosetta-gami pLysS) DE3 harboring pET-50b (+) construct (Coomassie blue staining). Lane M is a protein molecular weight (MW) ladder (Precision Plus protein, Bio-RAD), Lane 1 = uninduced (nprB) pET50b (+) construct. Lane 3–5 = IPTG-induced (nprB) pET50b (+) construct. Lane 2 = Ni-coated Hi-trap FPLC-purified protease protein (~20 kDa) fused with Nus + His-tag (~50 kDa), making a total ~70 kDa.</p>
Full article ">Figure 2
<p>FPLC chromatogram of recombinant neutral protease B (NprB) cloned in pET-50b (+) and expressed in <span class="html-italic">E. coli</span> BL21 (Rosetta–gami pLysS) DE3. Fraction elution A2 and A3 showed positive results.</p>
Full article ">Figure 3
<p>Western blot analysis for protease expression extracted from the <span class="html-italic">E. coli</span> BL21 (Rosetta–gami pLysS) DE3 harboring recombinant (nprB) pET50-b (+) at 1 mM IPTG induction and run on 12% SDS-PAGE. Lane M protein molecular weight (MW) ladder (Thermo Scientific page Ruler). IPTG Lanes 1 = Transferred protein on nitrocellulose using rabbit-anticamelysin (1:5000) and anti-rabbit IgG- alkaline phosphatase (Sigma) (protease protein (~20 kDa) fused with His-tag and Nus-tag protein (~50 kDa) making a total ~70 kDa).</p>
Full article ">Figure 4
<p>The effect of pH on the recombinant neutral protease B protein relative activity. The temperature used to investigate the pH effect was 70 °C. Error bars represent the standard deviations of three measurements.</p>
Full article ">Figure 5
<p>The effect of temperature on the activity of neutral protease B. Error bars represent the standard deviations of three measurements.</p>
Full article ">Figure 6
<p>Effect of different surfactants and inhibitors on protease activity. Error bars represent the standard deviations of three measurements.</p>
Full article ">Figure 7
<p>Effect of different inhibitors on protease activity at 1 mM and 5 mM concentrations. Error bars represent the standard deviations of three measurements.</p>
Full article ">
13 pages, 3706 KiB  
Article
Signal-On Detection of Caspase-3 with Methylene Blue-Loaded Metal-Organic Frameworks as Signal Reporters
by Yaliang Huang, Jiaqiang Wang, Yirui Xu, Jiwen Zhang and Ning Xia
Molecules 2024, 29(15), 3700; https://doi.org/10.3390/molecules29153700 - 5 Aug 2024
Viewed by 513
Abstract
In this work, we report on an electrochemical method for the signal-on detection of caspase-3 and the evaluation of apoptosis based on the biotinylation reaction and the signal amplification of methylene blue (MB)-loaded metal–organic frameworks (MOFs). Zr-based UiO-66-NH2 MOFs were used as [...] Read more.
In this work, we report on an electrochemical method for the signal-on detection of caspase-3 and the evaluation of apoptosis based on the biotinylation reaction and the signal amplification of methylene blue (MB)-loaded metal–organic frameworks (MOFs). Zr-based UiO-66-NH2 MOFs were used as the nanocarriers to load electroactive MB molecules. Recombinant hexahistidine (His6)-tagged streptavidin (rSA) was attached to the MOFs through the coordination interaction between the His6 tag in rSA and the metal ions on the surface of the MOFs. The acetylated peptide substrate Ac-GDEVDGGGPPPPC was immobilized on the gold electrode. In the presence of caspase-3, the peptide was specifically cleaved, leading to the release of the Ac-GDEVD sequence. A N-terminal amine group was generated and then biotinylated in the presence of biotin-NHS. Based on the strong interaction between rSA and biotin, rSA@MOF@MB was captured by the biotinylated peptide-modified electrode, producing a significantly amplified electrochemical signal. Caspase-3 was sensitively determined with a linear range from 0.1 to 25 pg/mL and a limit of detection down to 0.04 pg/mL. Further, the active caspase-3 in apoptosis inducer-treated HeLa cells was further quantified by this method. The proposed signal-on biosensor is compatible with the complex biological samples and shows great potential for apoptosis-related diagnosis and the screening of caspase-targeting drugs. Full article
(This article belongs to the Special Issue Metal Organic Frameworks (MOFs) for Sensing Applications)
Show Figures

Figure 1

Figure 1
<p>SEM images of (<b>A</b>) UiO-66-NH<sub>2</sub> MOF and (<b>B</b>) MOF@MB. (<b>C</b>) XRD patterns of UiO-66-NH<sub>2</sub> MOF (black curve) and MOF@MB (red curve). (<b>D</b>) UV–vis absorption spectra of MB (black curve), UiO-66-NH<sub>2</sub> MOF (red curve), and MOF@MB (blue curve). The inset in panel A shows the size distribution of the UiO-66-NH<sub>2</sub> MOF.</p>
Full article ">Figure 2
<p>BET analysis of UiO-66-NH<sub>2</sub> MOF and MOF@MB.</p>
Full article ">Figure 3
<p>DPV responses corresponding to the peptide-modified sensing electrode after treatment by biotin-NHS + rSA@MOF@MB (<b>a</b>), caspase-3 + biotin-NHS + rSA@MOF@MB (<b>b</b>), and caspase-3 + rSA@MOF@MB (<b>c</b>).</p>
Full article ">Figure 4
<p>The optimized times for peptide modification (<b>A</b>) and proteolysis (<b>B</b>).</p>
Full article ">Figure 5
<p>(<b>A</b>) DPV responses of the electrochemical biosensor at different concentrations of caspase-3 (from top to bottom: 0, 0.1, 1, 5, 10, 25, 50, and 100 pg/mL). (<b>B</b>) The relationship between peak current and caspase-3 concentration. The inset shows the linear portion of the fitting curve.</p>
Full article ">Figure 6
<p>DPV responses of the electrochemical biosensor for the detection of different samples. The concentrations of caspase-3, inhibitor, BSA, thrombin, and trypsin were 25 pg/mL, 50 pM, 10 ng/mL, 1 ng/mL, and 1 ng/mL, respectively.</p>
Full article ">Figure 7
<p>The relationship between peak current and inhibitor concentration for the assay of 25 pg/mL caspase-3.</p>
Full article ">Figure 8
<p>The relationship between peak current and cell number for the assays of caspase-3 in HeLa cells with and without treatment by STS.</p>
Full article ">Scheme 1
<p>A schematic illustration of the preparation of rSA@MOF@MB and the principle for the electrochemical detection of caspase-3.</p>
Full article ">
15 pages, 7083 KiB  
Article
Metal Ion Binding to Human Glutaminyl Cyclase: A Structural Perspective
by Giusy Tassone, Cecilia Pozzi and Stefano Mangani
Int. J. Mol. Sci. 2024, 25(15), 8279; https://doi.org/10.3390/ijms25158279 - 29 Jul 2024
Viewed by 464
Abstract
Glutaminyl-peptide cyclotransferases (QCs) convert the N-terminal glutamine or glutamate residues of protein and peptide substrates into pyroglutamate (pE) by releasing ammonia or a water molecule. The N-terminal pE modification protects peptides/proteins against proteolytic degradation by amino- or exopeptidases, increasing their stability. Mammalian QC [...] Read more.
Glutaminyl-peptide cyclotransferases (QCs) convert the N-terminal glutamine or glutamate residues of protein and peptide substrates into pyroglutamate (pE) by releasing ammonia or a water molecule. The N-terminal pE modification protects peptides/proteins against proteolytic degradation by amino- or exopeptidases, increasing their stability. Mammalian QC is abundant in the brain and a large amount of evidence indicates that pE peptides are involved in the onset of neural human pathologies such as Alzheimer’s and Huntington’s disease and synucleinopathies. Hence, human QC (hQC) has become an intensively studied target for drug development against these diseases. Soon after its characterization, hQC was identified as a Zn-dependent enzyme, but a partial restoration of the enzyme activity in the presence of the Co(II) ion was also reported, suggesting a possible role of this metal ion in catalysis. The present work aims to investigate the structure of demetallated hQC and of the reconstituted enzyme with Zn(II) and Co(II) and their behavior in the presence of known inhibitors. Furthermore, our structural determinations provide a possible explanation for the presence of the mononuclear metal binding site of hQC, despite the presence of the same conserved metal binding motifs present in distantly related dinuclear aminopeptidase enzymes. Full article
Show Figures

Figure 1

Figure 1
<p>The chemical structure of PBD-150 and SEN-177.</p>
Full article ">Figure 2
<p>(<b>A</b>) Active site view of apo-hQC-2X (in cartoon, light cyan carbons; residues in sticks). A water molecule, WatC, occupies the metal cofactor pocket, forming H-bonds with Asp159, Glu202, His330, and a further solvent molecule, Wat2. (<b>B</b>) Active site residues and water molecules surrounded by the 2<span class="html-italic">F<sub>o</sub></span> − <span class="html-italic">F<sub>c</sub></span> map (blue wire, contoured at 1.5 σ). Oxygen and nitrogen atoms are colored red and blue, respectively. Water molecules are represented as red spheres and H-bonds as blue dashed lines.</p>
Full article ">Figure 3
<p>Active site view of the superimposition between apo-hQC-2X (in cartoon, light cyan carbons; residues in sticks; water molecules in cyan) and hQC-2X in complex with Zn(II) ions (gold carton, carbons, and water molecules; Zn(II) ion in gray; PDB id 4UY9 [<a href="#B36-ijms-25-08279" class="html-bibr">36</a>]). The comparison shows the high conservation of the catalytic cavity configuration upon metal removal.</p>
Full article ">Figure 4
<p>(<b>A</b>) Active site view of the structure of hQC-2X (in cartoon, light cyan carbons; residues in ticks) in complex with Co(II) ions (gray sphere). Coordination bonds are colored gray. (<b>B</b>) Zn(II) ion and its coordinating residues are surrounded by the 2<span class="html-italic">F<sub>o</sub></span> − <span class="html-italic">F<sub>c</sub></span> map (blue wire, contoured at 1.5 σ), and the anomalous map computed at the zinc K-edge (dark gray wire, contoured at 5 σ). (<b>C</b>) Active site view of the structure of hQC-2X (in cartoon, light cyan carbons; residues in ticks) in complex with Co(II) ions (magenta sphere). Coordination and H-bonds are colored magenta and blue, respectively. (<b>D</b>) Co(II) ion and its coordinating residues are surrounded by the 2<span class="html-italic">F<sub>o</sub></span> − <span class="html-italic">F<sub>c</sub></span> map (blue wire, contoured at 1.5 σ), and the anomalous map computed at the cobalt K-edge (purple wire, contoured at 5 σ). Oxygen atoms are colored red and nitrogen blue. Water molecules are represented as red spheres.</p>
Full article ">Figure 5
<p>(<b>A</b>) Binding of PBD-150 (in sticks, orange carbon atoms) to the active site of hQC-2X (in cartoon, light cyan carbons; residues in ticks) in complex with Co(II) ions (magenta sphere). Coordination and H-bonds are colored magenta and blue, respectively. (<b>B</b>) PBD-150, Co(II) ion and its coordinating residues are surrounded by the 2<span class="html-italic">F<sub>o</sub></span> − <span class="html-italic">F<sub>c</sub></span> map (blue wire, contoured at 1.5 σ), and the anomalous map computed at the cobalt K-edge (purple wire, contoured at 5 σ). (<b>C</b>–<b>E</b>) Active site view of PBD-150 and Co(II) ion surrounded by the omit map (dark green wire, contoured at 2.5 σ) and the anomalous map computed at the cobalt K-edge (purple wire, contoured at 5 σ), respectively, in chain A (panel (<b>C</b>)), B (panel (<b>D</b>)), and C (panel (<b>E</b>)). Oxygen atoms are colored red, nitrogen blue, and sulfur yellow. Water molecules are represented as red spheres.</p>
Full article ">Figure 6
<p>(<b>A</b>) Binding of SEN-177 (in sticks, green carbon atoms) to the active site of hQC-2X (in cartoon, light cyan carbons; residues in ticks) in complex with Co(II) ions (magenta sphere). Coordination and H-bonds are colored magenta and blue, respectively. (<b>B</b>) SEN-177, Co(II) ion and its coordinating residues are surrounded by the 2<span class="html-italic">F<sub>o</sub></span> − <span class="html-italic">F<sub>c</sub></span> map (blue wire, contoured at 1.5 σ), and the anomalous map computed at the cobalt K-edge (purple wire, contoured at 5 σ). (<b>C</b>–<b>E</b>) Active site view of SEN-177 and Co(II) ion surrounded by the omit map (dark green wire, contoured at 2.5 σ) and the anomalous map computed at the cobalt K-edge (purple wire, contoured at 5 σ), respectively, in chain A (panel (<b>C</b>)), B (panel (<b>D</b>)), and C (panel (<b>E</b>)). Oxygen atoms are colored red, nitrogen blue, sulfur yellow, and halogen gray. Water molecules are represented as red spheres.</p>
Full article ">Figure 7
<p>Active site view of the structural comparison between the complexes (<b>A</b>) hQC-2X–Co(II) – PBD-150 (light cyan cartoon and carbons; Co(II) ion as magenta sphere and PBD-150 in sticks, cyan carbons) and hQC-2X–Zn(II)–PBD-150 (gold cartoon and carbons; Zn(II) ion as gray sphere and PBD-150 in sticks, gold carbons; PDB id 4YWY [<a href="#B36-ijms-25-08279" class="html-bibr">36</a>]); (<b>B</b>) hQC-2X–Co(II)–SEN-177 (light cyan cartoon and carbons; Co(II) ion as magenta sphere and SEN-177 in sticks, cyan carbons) and hQC-2X–Zn(II) –SEN-177 (gold cartoon and carbons; Zn(II) ion as gray sphere and SEN-177 in sticks, gold carbons; PDB id 6GBX [<a href="#B32-ijms-25-08279" class="html-bibr">32</a>]). The comparisons highlight the similar binding modes of the inhibitors in the complexes with Zn(II) and Co(II) metal ions.</p>
Full article ">Figure 8
<p>(<b>A</b>) Superimposition of Zn1-loaded hQC-2X (pale cyan sticks and blue labels) with ApAP (green sticks and green labels). Although distant in phylogeny and in sequence, the remarkable structural coincidence of the metal binding sites can readily be appreciated. (<b>B</b>) Surface representation (coded by electrostatic potential) of the active site cavity of ApAP with the two Zn(II) ions (green) protruding in the cavity. (<b>C</b>) Surface representation (coded by electrostatic potential) of the active site cavity of hQC-2X with the Zn(II) ion (white) protruding in the cavity. The described H-bonds between are shown as black dashed lines, while the Ser160-Zn2 distance is shown as red dashed line. The difference in the cavity volume due to the Ser160-Asp248 H-bond can be appreciated.</p>
Full article ">Scheme 1
<p>Schematic reactions catalyzed by hQC. Data regarding the optimum pH for the catalyzed reactions are from Schilling et al. [<a href="#B5-ijms-25-08279" class="html-bibr">5</a>].</p>
Full article ">
18 pages, 3320 KiB  
Review
Functional Diversity and Engineering of the Adenylation Domains in Nonribosomal Peptide Synthetases
by Mengli Zhang, Zijing Peng, Zhenkuai Huang, Jiaqi Fang, Xinhai Li and Xiaoting Qiu
Mar. Drugs 2024, 22(8), 349; https://doi.org/10.3390/md22080349 - 29 Jul 2024
Viewed by 807
Abstract
Nonribosomal peptides (NRPs) are biosynthesized by nonribosomal peptide synthetases (NRPSs) and are widely distributed in both terrestrial and marine organisms. Many NRPs and their analogs are biologically active and serve as therapeutic agents. The adenylation (A) domain is a key catalytic domain that [...] Read more.
Nonribosomal peptides (NRPs) are biosynthesized by nonribosomal peptide synthetases (NRPSs) and are widely distributed in both terrestrial and marine organisms. Many NRPs and their analogs are biologically active and serve as therapeutic agents. The adenylation (A) domain is a key catalytic domain that primarily controls the sequence of a product during the assembling of NRPs and thus plays a predominant role in the structural diversity of NRPs. Engineering of the A domain to alter substrate specificity is a potential strategy for obtaining novel NRPs for pharmaceutical studies. On the basis of introducing the catalytic mechanism and multiple functions of the A domains, this article systematically describes several representative NRPS engineering strategies targeting the A domain, including mutagenesis of substrate-specificity codes, substitution of condensation-adenylation bidomains, the entire A domain or its subdomains, domain insertion, and whole-module rearrangements. Full article
(This article belongs to the Section Synthesis and Medicinal Chemistry of Marine Natural Products)
Show Figures

Figure 1

Figure 1
<p>The typical domain arrangement of NRPS. Adenylation (A) domain, thiolation (T) domain, condensation (C) domain, and thioesterase (TE) domain are represented in orange, blue, green, and pink circles, respectively. Note: The module illustrated in the figure refers to a unit that comprises the three core domains C, A, and T (the initiation module, shown on the left, consists of A and T domains only, while the termination module, illustrated on the right, contains C, A, T and TE domains).</p>
Full article ">Figure 2
<p>Activation and loading of substrate catalyzed by the A domain. Initially, the A domain activates the acyl monomer: substrate reacts with adenosine 5′-triphosphate (ATP) to generate the acyl-adenosine monophosphate (AMP) intermediate and inorganic pyrophosphate (PPi). Subsequently, the aminoacyl-AMP undergoes nucleophilic attack by the thiol group located at the terminus of the 4′-phosphopantetheine arm of the downstream thiolation (T) domain, leading to the formation of a thioester-bound aminoacyl-S-T domain by linking to the thiol of the phosphopantetheine arm of T domain, followed by the release of AMP.</p>
Full article ">Figure 3
<p>Composition of some interrupted A domains: In KtzH, an interruption occurs between a8 and a9 codes of the A domain in module 4, where an O-methylase domain is inserted [<a href="#B55-marinedrugs-22-00349" class="html-bibr">55</a>]. In TioS, interruptions occur between a8 and a9 codes of the A domain in both module 3 and module 4, with the insertion of an N-methylase domain [<a href="#B53-marinedrugs-22-00349" class="html-bibr">53</a>]. In TioN, an interruption occurs between a2 and a3 codes of the A domain, with the insertion of an S-methylase domain [<a href="#B50-marinedrugs-22-00349" class="html-bibr">50</a>]. In MarQ, an interruption occurs between a2 and a3 codes of the A domain, with the insertion of an S-methylase domain [<a href="#B57-marinedrugs-22-00349" class="html-bibr">57</a>]. M<sub>b</sub>: main-chain methylase domain (yellow); M<sub>s</sub>: side-chain methylase domain (light green).</p>
Full article ">Figure 4
<p>Two typical cases of A domain substitution. (<b>A</b>) The A domain in <span class="html-italic">Bacillus subtilis</span> SrfA-C, which recognizes Leu, is replaced with an A domain that recognizes Cys. (<b>B</b>) The A domain of the second module of PvdD that recognizes L-Thr along with the linker region is replaced with a randomly selected one from nine different types of A domains that recognize other substrates and their corresponding linker regions. Among these, substitutions of six exchanged A domains labeled with dashed circles achieved higher yields. Abbreviations: aad: δ-L-α-aminoadipyl residue, fhOrn: N5-formyl-N5-hydroxyornithine residue.</p>
Full article ">Figure 5
<p>The cartoon representation of the overall structure of GrsA (PDB ID: 1AMU), phenylalanine-selective A domain of gramicidin synthetase 1, showing the subdomain and core domain of the A domain, which are represented in yellow and cyan, respectively.</p>
Full article ">Figure 6
<p>The second A domain of EndA that recognizes L-Thr. The FSD responsible for substrate recognition in this domain is replaced with the FSD from EndC that recognizes Ser, enabling the second A domain of EndA to recognize Ser. Abbreviation: Hpg: hydroxyphenylglycine.</p>
Full article ">Figure 7
<p>Several typical cases of sequence removal and insertion in A domains. (<b>A</b>) The wild-type A domain of KtzH is naturally interrupted. The methylase activity of this domain is abolished after its interrupted Ms domain is removed. (<b>B</b>) Replacing the interrupted M<sub>s</sub> domain in KtzH with the interrupted M domain from TioN yields an A domain exhibiting both methylation and adenylation functions; inserting the interrupted M domains from KtzH and TioS into the uninterrupted A domain of Ecm6 results in the generation of corresponding A domains with dual functions of methylation and adenylation.</p>
Full article ">Figure 8
<p>The domain arrangement and overall structure of SrfA-C. (<b>A</b>) The domain arrangement of the SrfA-C module: C domain, A domain, T domain, TE domain, and linker regions that connect these domains are represented as green, orange, blue, pink, and gray, respectively. (<b>B</b>) The surface representation of the overall structure of SrfA-C (PDB ID: 2VSQ), as well as the colors of the domains and linker regions, are identical to that in (<b>A</b>).</p>
Full article ">Figure 9
<p>Two typical cases of whole-module rearrangement. (<b>A</b>) Deleting the second module in SrfA-A and connecting modules 1 and 3 resulted in the production of a hexapeptide lacking a Leu residue. (<b>B</b>) By the insertion of a composite module composed of the T-E bidomain from module 4 and the C-A bidomain from module 5 in BpsB into modules 4 and 5 of BpsB, leading to the production of a novel octapeptide with three Hpg residues. Abbreviation: Hpg: hydroxyphenylglycine.</p>
Full article ">
19 pages, 5369 KiB  
Article
What Are the Key Factors for the Detection of Peptides Using Mass Spectrometry on Boron-Doped Diamond Surfaces?
by Juvissan Aguedo, Marian Vojs, Martin Vrška, Marek Nemcovic, Zuzana Pakanova, Katerina Aubrechtova Dragounova, Oleksandr Romanyuk, Alexander Kromka, Marian Varga, Michal Hatala, Marian Marton and Jan Tkac
Nanomaterials 2024, 14(15), 1241; https://doi.org/10.3390/nano14151241 - 24 Jul 2024
Viewed by 940
Abstract
We investigated the use of boron-doped diamond (BDD) with different surface morphologies for the enhanced detection of nine different peptides by matrix-assisted laser desorption/ionisation mass spectrometry (MALDI-MS). For the first time, we compared three different nanostructured BDD film morphologies (Continuous, Nanograss, and Nanotips) [...] Read more.
We investigated the use of boron-doped diamond (BDD) with different surface morphologies for the enhanced detection of nine different peptides by matrix-assisted laser desorption/ionisation mass spectrometry (MALDI-MS). For the first time, we compared three different nanostructured BDD film morphologies (Continuous, Nanograss, and Nanotips) with differently terminated surfaces (-H, -O, and -F) to commercially available Ground Steel plates. All these surfaces were evaluated for their effectiveness in detecting the nine different peptides by MALDI-MS. Our results demonstrated that certain nanostructured BDD surfaces exhibited superior performance for the detection of especially hydrophobic peptides (e.g., bradykinin 1–7, substance P, and the renin substrate), with a limit of detection of down to 2.3 pM. Further investigation showed that hydrophobic peptides (e.g., bradykinin 1–7, substance P, and the renin substrate) were effectively detected on hydrogen-terminated BDD surfaces. On the other hand, the highly acidic negatively charged peptide adrenocorticotropic hormone fragment 18–39 was effectively identified on oxygen-/fluorine-terminated BDD surfaces. Furthermore, BDD surfaces reduced sodium adduct contamination significantly. Full article
(This article belongs to the Special Issue Carbon-Based Nanomaterials for Biomedicine Applications)
Show Figures

Figure 1

Figure 1
<p>SEM images (top, angular, and cross-sectional views) of BDD morphologies for Continuous film and structured Nanograss or Nanotips.</p>
Full article ">Figure 2
<p>Raman spectra of the as-grown (Continuous) and structured (Nanograss and Nanotip) BDD films.</p>
Full article ">Figure 3
<p>Wetting properties of as-grown (Continuous) and structured (Nanotip) BDD samples after different surface termination treatments.</p>
Full article ">Figure 4
<p>XPS survey spectra of (<b>a</b>) Continuous BDD, (<b>b</b>) Nanograss BDD, and (<b>c</b>) Nanotip BDD samples as-prepared (black) and with -H (red), -O (green), and -F (blue) terminations.</p>
Full article ">Figure 5
<p>Photograph of prepared MALDI-MS chip with BDD spots deposited.</p>
Full article ">Figure 6
<p>MALDI-MS workflow and MS spectra of the standard peptide mixture solution (0.2 pmol μL<sup>−1</sup>) performed in positive mode using the DHB matrix in all the cases: (<b>A</b>) Ground Steel, (<b>B</b>) Continuous BDD, and (<b>C</b>) Nanograss BDD.</p>
Full article ">Figure 7
<p>Classes of amino acids present in the peptides expressed in percentages (%). Brad: Bradykinin 1–7; Ang II: Angiotensin II; Ang I: Angiotensin I; Sub P: Substance P; Bomb: Bombesin; Ren S: Renin Substrate; AC 1: adrenocorticotropic hormone fragment 1–17; AC 18: adrenocorticotropic hormone fragment 18–39; Som: Somatostatin. Peptides highlighted in green showed significantly better LODs when detected at BDD interfaces in comparison to the commercially available Ground Steel chip. This figure was made based on data taken from <a href="#app1-nanomaterials-14-01241" class="html-app">Table S1</a>. The dashed green line represents the average percentages of hydrophobic amino acids present in Brad, Sub P, and Ren S.</p>
Full article ">Figure 8
<p>Three-dimensional (3D) plots showing the dependences of the LOD<sub>ratio</sub> value on the <span class="html-italic">m</span>/<span class="html-italic">z</span> and charge.</p>
Full article ">Figure 9
<p>The influence of work function (WF) on LOD<sub>ratio</sub> values for the Sub P peptide. The data point marked by the red circle was not a part of the linear plot.</p>
Full article ">Scheme 1
<p>Preparation of BDD interfaces.</p>
Full article ">
18 pages, 4998 KiB  
Review
Structural Catalytic Core of the Members of the Superfamily of Acid Proteases
by Alexander I. Denesyuk, Konstantin Denessiouk, Mark S. Johnson and Vladimir N. Uversky
Molecules 2024, 29(15), 3451; https://doi.org/10.3390/molecules29153451 - 23 Jul 2024
Viewed by 533
Abstract
The superfamily of acid proteases has two catalytic aspartates for proteolysis of their peptide substrates. Here, we show a minimal structural scaffold, the structural catalytic core (SCC), which is conserved within each family of acid proteases, but varies between families, and thus can [...] Read more.
The superfamily of acid proteases has two catalytic aspartates for proteolysis of their peptide substrates. Here, we show a minimal structural scaffold, the structural catalytic core (SCC), which is conserved within each family of acid proteases, but varies between families, and thus can serve as a structural marker of four individual protease families. The SCC is a dimer of several structural blocks, such as the DD-link, D-loop, and G-loop, around two catalytic aspartates in each protease subunit or an individual chain. A dimer made of two (D-loop + DD-link) structural elements makes a DD-zone, and the D-loop + G-loop combination makes a psi-loop. These structural markers are useful for protein comparison, structure identification, protein family separation, and protein engineering. Full article
(This article belongs to the Special Issue Protein Structure, Function and Interaction)
Show Figures

Figure 1

Figure 1
<p>Three building blocks of the structural catalytic core (SCC) in propepsin (PDB ID: 3PSG), as a representative member of the pepsin-like family of the acid protease superfamily. (<b>A</b>) DD-zone, (<b>B</b>) psi-loop<sub>N</sub>, and (<b>C</b>) psi-loop<sub>C</sub>. The dashed lines show long-range hydrogen bonds between the bordering amino acids of fragments of the primary structure of the protein: D-loops, DD-link, mediator, and G-loops, thus determining the cyclic nature and composition of the residues of each block separately. A dimer of dipeptides, Asp<sub>32</sub>-Thr<sub>33</sub> and Asp<sub>215</sub>-Thr<sub>216</sub>, from two D-loops, form the fireman’s grip in the DD-zone, which is characterized by four long-range hydrogen bonds, while tetrapeptides, Asp<sub>32</sub>-...-Ser<sub>35</sub> and Asp<sub>215</sub>-...-Thr<sub>218,</sub> from two D-loops, form the Asx-motif in psi-loop<sub>N</sub> and psi-loop<sub>C</sub>, which is characterized by two short-range hydrogen bonds. Structural differences in two long-range hydrogen bonds located within psi-loop<sub>N</sub> (O/Asp<sub>32</sub>-N/Leu<sub>123</sub> and (O/Ser<sub>35</sub>-N/Ala<sub>124</sub>) and psi-loop<sub>C</sub> (O/Thr<sub>218</sub>-N/Asp<sub>303</sub> and O/Ser<sub>219</sub>-N/Val<sub>304</sub>) influence the functional differences between the catalytic aspartates.</p>
Full article ">Figure 2
<p>Interface organization of interactions between the SCC of pepsin and the ligand saquinavir. (<b>A</b>) A smooth coil representation is shown that passes through the CA atom positions of the pepsin’s SCC. The dashed lines show the complete set of long-range hydrogen bonds between the bordering residues of the six amino-acid sequence fragments. (<b>B</b>) The potential hydrogen bonding interactions between the D-loops of the DD-zone and saquinavir are shown with dashed lines.</p>
Full article ">Figure 3
<p>The 3D structure of the active site in pepsin-like family aspartic proteases. The three boxes show the location of the structural catalytic core (SCC) in propepsin (PDB ID: 3PSG_A). It consists of a DD-zone (a central rectangle constructed using dotted lines) and two psi-loops (solid lines). The discussed structural elements (loops and links) are highlighted and labeled.</p>
Full article ">Figure 4
<p>The building blocks of the SCC in the HIV-1 and XMRV homodimer proteases (PDB IDs: 3IXO and 3NR6, correspondingly), as the representative members of the retroviral protease (retropepsin) family of the acid protease superfamily. (<b>A</b>) DD-zone of HIV-1 protease, (<b>B</b>) DD-zone of XMRV protease, and (<b>C</b>) psi-loop of HIV-1 protease. (<b>D</b>) The potential hydrogen bonding interactions (dashed lines) between two identical D-loops of the DD-zone and the ligand in the HIV-1 protease with inhibitor KNI-1657 complex (PDB ID: 5YOK).</p>
Full article ">Figure 5
<p>SCC of (<b>A</b>) HIV-1 and (<b>B</b>) XMRV proteases. A smooth coil representation is used in the figures, which passes through the CA atom of SCC positions of the corresponding retroviral proteases. The SCC of the XMRV protease differs from the SCC of the HIV-1 protease by the inclusion of the mediator residue Arg<sub>95</sub> from the G-loop in each monomer.</p>
Full article ">Figure 6
<p>The 3D structure of the active site in retroviral protease (retropepsin) family aspartic proteases. The three boxes show the location of the structural catalytic core (SCC) in HIV-1 protease (PDB ID: 3IXO_A, B). It consists of a DD-zone (a central rectangle constructed using dotted lines) and two psi-loops (solid lines). The discussed structural elements (loops and links) are highlighted and labeled.</p>
Full article ">Figure 7
<p>The building blocks of the SCC in the Ddi1 protease, Lpg0085 protein, and ApRick protease (PDB IDs: 4Z2Z, 2PMA and 5C9F, correspondingly), as the representative members of the dimeric aspartyl protease and LPG0085-like families of the acid protease superfamily. (<b>A</b>) DD-zone of Ddi1 protease, (<b>B</b>) DD-zone of protein Lpg0085, and (<b>C</b>) psi-loop of ApRick protease.</p>
Full article ">Figure 8
<p>SCC of (<b>A</b>) Ddi1 protease and (<b>B</b>) protein Lpg0085. The main differences between the SCCs of the two proteins are the amino acid composition of the DD-links and the use of a mediator-dipeptide in the structural formation of the DD-zone in the protein Lpg0085.</p>
Full article ">
16 pages, 1735 KiB  
Article
Low-Molecular-Weight Peptides Prepared from Hypsizygus marmoreus Exhibit Strong Antioxidant and Antibacterial Activities
by Shaoxiong Zhou, Zheng Xiao, Junzheng Sun, Longxiang Li, Yingying Wei, Mengjie Yang, Yanrong Yang, Junchen Chen and Pufu Lai
Molecules 2024, 29(14), 3393; https://doi.org/10.3390/molecules29143393 - 19 Jul 2024
Viewed by 664
Abstract
Hypsizygus marmoreus has abundant proteins and is a potential source for the development of bioactive peptides. However, currently, the research on the bioactive components of H. marmoreus mainly focuses on polysaccharides, and there is no relevant research on the preparation of bioactive peptides. [...] Read more.
Hypsizygus marmoreus has abundant proteins and is a potential source for the development of bioactive peptides. However, currently, the research on the bioactive components of H. marmoreus mainly focuses on polysaccharides, and there is no relevant research on the preparation of bioactive peptides. In this article, an ultrasound-assisted extraction method was used to extract proteins from H. marmoreus, and then, four peptides with different molecular weight ranges were prepared through protease hydrolysis and molecular classification. The antioxidant and antibacterial activities were also studied. Under the optimal conditions, the extraction rate of H. marmoreus proteins was 53.6%. Trypsin exhibited the highest hydrolysis rate of H. marmoreus proteins. The optimal parameters for enzymatic hydrolysis were a substrate concentration of 3.7%, enzyme addition of 5700 U/g, pH value of 7, extraction temperature of 55 °C, and time of 3.3 h. Under these conditions, the peptide yield was 59.7%. The four types of H. marmoreus peptides were prepared by molecular weight grading. Among them, peptides with low molecular weight (<1 kDa) had stronger antioxidant and antibacterial activities. This study provides a theoretical basis for the efficient preparation of H. marmoreus peptides and the development of antioxidant and antibacterial peptide products. Full article
Show Figures

Figure 1

Figure 1
<p>The optimization of parameters of <span class="html-italic">H. marmoreus</span> protein hydrolysis: (<b>A</b>) Proteases, (<b>B</b>) enzymatic hydrolysis time, (<b>C</b>) substrate concentration, (<b>D</b>) pH value, (<b>E</b>) extraction temperature, and (<b>F</b>) enzyme dosage. Different letters in the same indicators indicate significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>The effect of the interaction between hydrolysis time and enzyme dosage on hydrolysis degree.</p>
Full article ">Figure 3
<p>Scavenging rates <span class="html-italic">H. marmoreus</span> peptides on DPPH (<b>A</b>), •OH (<b>B</b>), ·O<sub>2</sub><sup>−</sup> (<b>C</b>) radicals. Different letters in the same indicators indicate significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
20 pages, 9899 KiB  
Article
Anti-Type II Diabetic Effects of Coix Seed Prolamin Hydrolysates: Physiological and Transcriptomic Analyses
by Guifang Zhang, Zhiming Li, Shu Zhang, Lu Bai, Hangqing Zhou and Dongjie Zhang
Foods 2024, 13(14), 2203; https://doi.org/10.3390/foods13142203 - 12 Jul 2024
Viewed by 782
Abstract
Previous studies have demonstrated that enzymatically prepared coix seed prolamin hydrolysates (CHPs) contain several bioactive peptides that efficiently inhibit the activity of target enzymes (α-glucosidase and dipeptidyl kinase-IV) in type 2 diabetes mellitus (T2DM). However, the anti-T2DM effects and potential mechanisms of CHPs [...] Read more.
Previous studies have demonstrated that enzymatically prepared coix seed prolamin hydrolysates (CHPs) contain several bioactive peptides that efficiently inhibit the activity of target enzymes (α-glucosidase and dipeptidyl kinase-IV) in type 2 diabetes mellitus (T2DM). However, the anti-T2DM effects and potential mechanisms of CHPs as a whole in vivo have not yet been systematically explored. Therefore, we evaluated the preventive, therapeutic, and modifying effects of CHPs on T2DM by combining physiological and liver transcriptomics with a T2DM mouse model. The results showed that sustained high-fructose intake led to prediabetic symptoms in mice, with abnormal fluctuations in blood glucose and blood lipid levels. Intervention with CPHs effectively prevented weight loss; regulated abnormal changes in blood glucose; improved impaired glucose tolerance; inhibited the abnormal expression of total cholesterol, triglycerides, and low-density lipoproteins; alleviated insulin resistance; and restored pancreatic islet tissue function in mice fed a high-fructose diet. In addition, we found that CHPs also play a palliative role in the loss of liver function and protect various organ tissues (including the liver, kidneys, pancreas, and heart), and are effective in preventing damage to the liver and pancreatic islet cells. We also found that the intake of CHPs reversed the abnormally altered hepatic gene profile in model mice and identified 381 differentially expressed genes that could serve as key genes for preventing the development of T2DM, which are highly correlated with multiple glycolipid metabolic pathways. We demonstrated that CHPs play a positive role in the normal functioning of the insulin signalling pathway dominated by the IRS-1/PI3K/AKT (insulin receptor substrates-1/phosphoinositide 3-kinase/protein kinase B) pathway. In summary, CHPs can be used as effective food-borne glucose-modifying components of healthy foods. Full article
Show Figures

Figure 1

Figure 1
<p>Effect of CHPs on body weight and organ index in T2DM mice (<span class="html-italic">n</span> = 12). (<b>A</b>) The experimental design of the study. (<b>B</b>) Body weight. (<b>C</b>) Body weight gain. (<b>D</b>) Liver index. (<b>E</b>) Pancreas index. (<b>F</b>) Renal index. (<b>G</b>) Cardiac index. The data are presented as means ± S.D, which were analysed by ANOVA test, followed by Tukey’s test between multiple groups. * indicates significant difference compared with the DC groups, * <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. ns: not significant.</p>
Full article ">Figure 2
<p>The effects of CHPs on serum glucose related-parameters in T2DM mice. (<b>A</b>) Fasting blood glucose. (<b>B</b>) OGTT. (<b>C</b>) The area under the curve of OGTT. (<b>D</b>) Insulin. The data are presented as means ± S.D, which were analysed by ANOVA test, followed by Tukey’s test between multiple groups. * indicates significant difference compared with the DC groups, ** <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 3
<p>The effects of CHPs on lipid levels and liver function index of serum in T2DM mice (<span class="html-italic">n</span> = 12). (<b>A</b>) TC, (<b>B</b>) TG, (<b>C</b>) HDL-c, (<b>D</b>) LDL-c, (<b>E</b>) ALT, (<b>F</b>) AST, (<b>G</b>) AIP, (<b>H</b>) CRI-I, (<b>I</b>) CRI-II. The data are presented as means ± S.D, which were analysed by ANOVA test, followed by Tukey’s test between multiple groups. * indicates significant difference compared with the DC groups, * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001. ns: not significant.</p>
Full article ">Figure 4
<p>The histopathological images and analyses of the liver and pancreas by HE staining. The image of the liver (<b>A</b>) and pancreas (<b>B</b>) at 200× magnification. The site indicated by the arrow in the figure represents ballooning from fat accumulation in hepatocytes.</p>
Full article ">Figure 5
<p>The effect of CHPs on transcriptomics profiles of T2DM mice liver (<span class="html-italic">n</span> = 5). (<b>A</b>) PCA. (<b>B</b>) The correlated heatmap in three groups. (<b>C</b>) The number of differentially expressed genes and heatmap of these differentially expressed genes in DC group versus NC group. (<b>D</b>) The number of differentially expressed genes and heatmap of these differentially expressed genes in CHP group versus DC group. (<b>E</b>) Venn diagrams. (<b>F</b>) Hierarchical clustering heatmap of differentially expressed genes in three groups. In the volcano plot analysis, the coloured dots correspond to upregulated (red) and downregulated (green) genes. In the clustering analysis, up- and downregulated genes are coloured red and blue, respectively.</p>
Full article ">Figure 6
<p>GO (<b>A</b>,<b>B</b>) analyses and KEGG (<b>C</b>,<b>D</b>) enrichment of the differentially expressed genes reversed by CHPs in liver tissues. (<b>A</b>,<b>C</b>) are the results of GO and KEGG enriched by upregulated genes, respectively; (<b>B</b>,<b>D</b>) are the results of GO and KEGG enriched by downregulated genes, respectively.</p>
Full article ">
Back to TopTop