www.fgks.org   »   [go: up one dir, main page]

You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (4,849)

Search Parameters:
Keywords = recombinant protein

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
27 pages, 7337 KiB  
Article
Entamoeba histolytica: EhADH, an Alix Protein, Participates in Several Virulence Events through Its Different Domains
by Dxinegueela Zanatta, Abigail Betanzos, Elisa Azuara-Liceaga, Sarita Montaño and Esther Orozco
Int. J. Mol. Sci. 2024, 25(14), 7609; https://doi.org/10.3390/ijms25147609 (registering DOI) - 11 Jul 2024
Viewed by 115
Abstract
Entamoeba histolytica is the protozoan causative of human amoebiasis. The EhADH adhesin (687 aa) is a protein involved in tissue invasion, phagocytosis and host-cell lysis. EhADH adheres to the prey and follows its arrival to the multivesicular bodies. It is an accessory protein [...] Read more.
Entamoeba histolytica is the protozoan causative of human amoebiasis. The EhADH adhesin (687 aa) is a protein involved in tissue invasion, phagocytosis and host-cell lysis. EhADH adheres to the prey and follows its arrival to the multivesicular bodies. It is an accessory protein of the endosomal sorting complexes required for transport (ESCRT) machinery. Here, to study the role of different parts of EhADH during virulence events, we produced trophozoites overexpressing the three domains of EhADH, Bro1 (1–400 aa), Linker (246–446 aa) and Adh (444–687 aa) to evaluate their role in virulence. The TrophozBro11–400 slightly increased adherence and phagocytosis, but these trophozoites showed a higher ability to destroy cell monolayers, augment the permeability of cultured epithelial cells and mouse colon, and produce more damage to hamster livers. The TrophozLinker226–446 also increased the virulence properties, but with lower effect than the TrophozBro11–400. In addition, this fragment participates in cholesterol transport and GTPase binding. Interestingly, the TrophozAdh444–687 produced the highest effect on adherence and phagocytosis, but it poorly influenced the monolayers destruction; nevertheless, they augmented the colon and liver damage. To identify the protein partners of each domain, we used recombinant peptides. Pull-down assays and mass spectrometry showed that Bro1 domain interplays with EhADH, Gal/GalNAc lectin, EhCPs, ESCRT machinery components and cytoskeleton proteins. While EhADH, ubiquitin, EhRabB, EhNPC1 and EhHSP70 were associated to the Linker domain, and EhADH, EhHSP70, EhPrx and metabolic enzymes interacted to the Adh domain. The diverse protein association confirms that EhADH is a versatile molecule with multiple functions probably given by its capacity to form distinct molecular complexes. Full article
(This article belongs to the Special Issue Molecular Dynamics of Membrane Proteins)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Scheme of the EhADH functions. (<b>I</b>) Adherence to target cells [<a href="#B7-ijms-25-07609" class="html-bibr">7</a>,<a href="#B14-ijms-25-07609" class="html-bibr">14</a>]. (<b>II</b>) Phagocytosis [<a href="#B7-ijms-25-07609" class="html-bibr">7</a>,<a href="#B14-ijms-25-07609" class="html-bibr">14</a>]. (<b>III</b>,<b>IV</b>) Vesicular trafficking and MVBs formation (it participates as an accessory protein of the ESCRT machinery) [<a href="#B14-ijms-25-07609" class="html-bibr">14</a>]. (<b>V</b>) Epithelial barrier impairing [<a href="#B7-ijms-25-07609" class="html-bibr">7</a>]. (<b>VI</b>) Potential vaccine candidate against amoebiasis [<a href="#B12-ijms-25-07609" class="html-bibr">12</a>]. In the center: 3D model of EhADH.</p>
Full article ">Figure 2
<p>Experimental approaches to study the EhADH domains. (<b>A</b>) EhADH 3D structure obtained from the I-TASSER server, showing in distinct colors the Bro1 (blue) and the V (red) domains as well as the Adh domain (yellow). N: amino terminal. C: carboxy terminal. (<b>B</b>) Upper panel: Generation of trophozoites overexpressing Bro1 (TrophozBro1<sub>1–400</sub>), Linker (TrophozLinker<sub>246–446</sub>) and Adh (TrophozAdh<sub>444–687</sub>) domains. Trophozoites transfected with the empty vector (<span class="html-italic">pTet</span>) are used as control (TrophozControl). Lower panel: Production of recombinant proteins with different tags containing the whole EhADH proteins or embracing different domains. (<b>C</b>) Upper panel: Experiments carried out to analyze the expression and localization of transfected trophozoites. Middle panel: Trophozoites overexpressing different domains were used for in vitro and in vivo virulence assays. Lower panel: Identification of binding-partners of each EhAdh domain by pull-down assays and mass spectrometry analysis.</p>
Full article ">Figure 3
<p>EhADH interactions with other <span class="html-italic">E. histolytica</span> proteins. (<b>A</b>) EhADH interactome generated by STRING. (<b>B</b>) Pull-down assays using the GST-EhADH recombinant protein and <span class="html-italic">E. histolytica</span> lysates. Randomly selected interacting proteins were analyzed by 10% SDS-PAGE followed by western blot assays, employing the α-EhCP112, α-EhVps23, α-EhVps32 and α-actin antibodies. Input: GST-EhADH. Numbers at the left indicate molecular weight standards in kDa. Arrows signal the immunodetected proteins.</p>
Full article ">Figure 4
<p>Molecular docking of EhADH and EhCP112. (<b>A</b>–<b>D</b>) Molecular docking of EhADH with EhCP112 at 200 (<b>A</b>), 300 (<b>B</b>), 400 (<b>C</b>) and 500 (<b>D</b>) ns, were obtained using the ClusPro server. EhADH: red. EhCP112: green. ΔG: binding energy. (<b>E</b>–<b>H</b>) Residues involved in the association of EhADH with EhCP112. Black amino acids belong to EhADH. Green amino acids belong to EhCP112. Axes: X in red, Y in green, Z in blue.</p>
Full article ">Figure 5
<p>Generation of specific antibodies against the EhADH domains. (<b>A</b>) Schematic representation of the primary structure and 3D model of the Bro1 (blue), Linker (red) and Adh (yellow) domains. Sequences displayed in green in the red squares indicate the peptides used for the antibodies’ generation in different animal models. N: amino terminal. C: carboxy terminal. (<b>B</b>–<b>D</b>) Western blots assays of trophozoite lysates, using the α-Bro1, α-Linker and α-Adh antibodies. PS: Pre-immune sera. Numbers at the left indicate standard molecular weight in kDa. Arrows signal the immunodetected proteins.</p>
Full article ">Figure 6
<p>Expression and cellular localization of the EhADH domains in <span class="html-italic">E. histolytica</span> trophozoites. TrophozControl, TrophozBro1<sub>1–400,</sub> TrophozLinker<sub>246–446</sub> or TrophozAdh<sub>444–687</sub> were generated by transfection of <span class="html-italic">pTet</span>, <span class="html-italic">pTet/Bro1</span>, <span class="html-italic">pTet/Linker</span> or <span class="html-italic">pTet/Adh</span> plasmids, respectively. (<b>A</b>–<b>C</b>) Western blot assays of trophozoite lysates, using the α-Bro1, α-Linker and α-Adh antibodies. The α-actin antibody was used as a loading control. Numbers at the left indicate standard molecular weight in kDa. Arrows signal the immunodetected proteins. (<b>D</b>–<b>F</b>) Immunofluorescence experiments of trophozoites employing α-Bro1, α-Linker and α-Adh as primary antibodies and FITC (green) as the secondary antibody. Nuclei (blue) were DAPI stained. Asterisks: cytoplasmic dots that could correspond to vesicular structures; arrowheads: plasma membrane; arrows: dots at cytoplasm and membrane. Bar = 20 μm.</p>
Full article ">Figure 7
<p>Adhesion to and phagocytosis of erythrocytes by trophozoites overexpressing the EhADH domains. TrophozControl, TrophozBro1<sub>1–400</sub>, TrophozLinker<sub>226–446</sub> and TrophozAdh<sub>444–687</sub> populations were incubated at different times with erythrocytes at 4 °C or 37 °C for adhesion and erythrophagocytosis assays, respectively. (<b>A</b>) Novikoff staining of trophozoites after 30 min of adhesion assays. (<b>B</b>) Number of erythrocytes adhered to trophozoites. Data represent the mean and standard error of the erythrocytes number counted on 100 randomly selected trophozoites in three independent experiments. (<b>C</b>) Novikoff staining of trophozoites that ingested erythrocytes for 30 min. T: trophozoite; e: erythrocyte. (<b>D</b>) Rate of erythrophagocytosis measured by the hemoglobin concentration inside trophozoites. **** <span class="html-italic">p</span> &lt; 0.0001, *** <span class="html-italic">p</span> &lt; 0.001 ** <span class="html-italic">p</span> &lt; 0.01, * <span class="html-italic">p</span> &lt; 0.1.</p>
Full article ">Figure 8
<p>Cytopathic and cytotoxic effect on epithelial cells produced by trophozoites overexpressing the EhADH domains. Destruction of MDCK cell monolayers incubated with live trophozoites (<b>A</b>) or extracts from trophozoites (50, 100, and 200 × 10<sup>3</sup>) (<b>B</b>). Monolayer damage was spectrophotometrically determined by methylene blue absorbed by the remaining monolayers. Thus, the cellular destruction was deducted from comparing harmed monolayers with confluent cells not incubated with trophozoites (DMEM), which were preserved intact. Representative images of stained MDCK monolayers with 200 × 10<sup>3</sup> trophozoites are shown at the right. Values represent the mean and standard error of three independent experiments. **** <span class="html-italic">p</span> &lt; 0.0001, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 9
<p>In vitro and in vivo effect of trophozoites overexpressing the EhADH domains on epithelial permeability. TrophozControl, TrophozBro1<sub>1–400</sub>, TrophozLinker<sub>246–446</sub> and TrophozAdh<sub>444–687</sub> were incubated with MDCK cells or inoculated in the mice colon. (<b>A</b>) MDCK monolayers were incubated with trophozoites for 120 min, and TEER was monitored. TEER was normalized according to the initial value for each Transwell (~1000 Ω·cm<sup>2</sup>). (<b>B</b>) FITC-dextran was apically added to MDCK cells grown in Transwells and incubated for 120 min with trophozoites. FITC-dextran obtained from the basal side was measured by fluorescence spectroscopy and normalized according to the epithelial cells treated with 5 mM EDTA used as positive control. DMEM: MDCK cells incubated with culture medium, without parasites. (<b>C</b>) C57/BL6 mice were rectally inoculated with trophozoites (10<sup>5</sup>) and after 30 min, the damage in the colonic epithelium was evaluated by Evan’s blue staining absorption. Representative images of distal and proximal portions of the colon after staining are shown. PBS: colon treated with PBS, without parasites. (<b>D</b>) Evan’s blue was eluted and spectrophotometrically measured at OD<sub>610nm</sub> to evaluate the epithelial permeability. Means and standard error are represented for each time point of three independent assays performed in triplicate. **** <span class="html-italic">p</span> &lt; 0.0001, *** <span class="html-italic">p</span> &lt; 0.001, ** <span class="html-italic">p</span> &lt; 0.01, * <span class="html-italic">p</span> &lt; 0.1.</p>
Full article ">Figure 10
<p>ALA in hamsters inoculated with trophozoites overexpressing the EhADH domains. Hamsters were intraportally inoculated with TrophozControl, TrophozBro1<sub>1–400</sub>, TrophozLinker<sub>246–446</sub> or TrophozAdh<sub>444–687</sub> trophozoites (2 × 10<sup>6</sup>). Eight days later, animals were anesthetized, and livers were extracted to examine the size and damage produced. (<b>A</b>) Livers´ weights. (<b>B</b>) Representative dorsal and ventral views of livers. (<b>C</b>) Hepatic damage evaluated as the weight of the abscesses formed divided by the weight of the whole liver before the injured areas were removed. Values represent the mean and standard error of five independent experiments. **** <span class="html-italic">p</span> &lt; 0.0001, ** <span class="html-italic">p</span> &lt; 0.01, * <span class="html-italic">p</span> &lt; 0.1.</p>
Full article ">Figure 11
<p>Generation of the recombinant EhADH domains. (<b>A</b>) Western blot experiments of recombinant proteins generated from the nucleotide encoding sequence for each domain that was cloned into expression vectors, generating the <span class="html-italic">pCold/Bro1</span>, <span class="html-italic">pGEX/Linker</span> and <span class="html-italic">pGEX/Adh</span> constructs. The expression of recombinant proteins was IPTG-induced. Numbers at the left indicate standard molecular weight in kDa. (<b>B</b>–<b>D</b>) Western blot assays of interacting proteins obtained from pull-down assays using the recombinant proteins ((<b>B</b>): His-Bro1; (<b>C</b>): GST-Linker; and (<b>D</b>): GST-Adh) and <span class="html-italic">E. histolytica</span> lysates. Numbers at the right show the molecular weight of immunodetected proteins.</p>
Full article ">Figure 12
<p>Working model of possible functions for each EhADH domain and recruited molecules involved in virulence events. (a) The Adh domain participates in target cells adhesion, interacting with other parasite molecules including the Gal/GalNAc lectin. (b) It also participates in phagocytosis, together with the ESCRT machinery. (c) Bro1 domain is involved in the IJs disruption, probably by its binging to occludin and claudins, and promotes the complexes formation with EhCPs and ESCRT machinery members. (d) The trophozoites can reach the liver by the portal vein, generating ALA with the main participation of Adh domain, which interplays with EhCPs and metabolic enzymes.</p>
Full article ">
21 pages, 2525 KiB  
Article
A Novel Cytotoxic Mechanism for Triple-Negative Breast Cancer Cells Induced by the Type II Heat-Labile Enterotoxin LT-IIc through Ganglioside Ligation
by Natalie D. King-Lyons, Aryana S. Bhati, John C. Hu, Lorrie M. Mandell, Gautam N. Shenoy, Hugh J. Willison and Terry D. Connell
Toxins 2024, 16(7), 311; https://doi.org/10.3390/toxins16070311 - 11 Jul 2024
Viewed by 151
Abstract
Triple-negative breast cancer (TNBC), which constitutes 10–20 percent of all breast cancers, is aggressive, has high metastatic potential, and carries a poor prognosis due to limited treatment options. LT-IIc, a member of the type II subfamily of ADP-ribosylating—heat-labile enterotoxins that bind to a [...] Read more.
Triple-negative breast cancer (TNBC), which constitutes 10–20 percent of all breast cancers, is aggressive, has high metastatic potential, and carries a poor prognosis due to limited treatment options. LT-IIc, a member of the type II subfamily of ADP-ribosylating—heat-labile enterotoxins that bind to a distinctive set of cell-surface ganglioside receptors—is cytotoxic toward TNBC cell lines, but has no cytotoxic activity for non-transformed breast epithelial cells. Here, primary TNBC cells, isolated from resected human tumors, showed an enhanced cytotoxic response specifically toward LT-IIc, in contrast to other enterotoxins that were tested. MDA-MB-231 cells, a model for TNBC, were used to evaluate potential mechanisms of cytotoxicity by LT-IIc, which induced elevated intracellular cAMP and stimulated the cAMP response element-binding protein (CREB) signaling pathway. To dissect the role of ADP-ribosylation, cAMP induction, and ganglioside ligation in the cytotoxic response, MDA-MB-231 cells were exposed to wild-type LT-IIc, the recombinant B-pentamer of LT-IIc that lacks the ADP-ribosylating A polypeptide, or mutants of LT-IIc with an enzymatically inactivated A1-domain. These experiments revealed that the ADP-ribosyltransferase activity of LT-IIc was nonessential for inducing the lethality of MDA-MB-231 cells. In contrast, a mutant LT-IIc with an altered ganglioside binding activity failed to trigger a cytotoxic response in MDA-MB-231 cells. Furthermore, the pharmacological inhibition of ganglioside expression protected MDA-MB-231 cells from the cytotoxic effects of LT-IIc. These data establish that ganglioside ligation, but not the induction of cAMP production nor ADP-ribosyltransferase activity, is essential to initiating the LT-IIc-dependent cell death of MDA-MB-231 cells. These experiments unveiled previously unknown properties of LT-IIc and gangliosides in signal transduction, offering the potential for the targeted treatment of TNBC, an option that is desperately needed. Full article
(This article belongs to the Section Bacterial Toxins)
Show Figures

Figure 1

Figure 1
<p>LT-IIc elicits a strong cytotoxic response in primary cancer cells isolated from TNBC tumors. Primary cells, isolated from four human, resected tumors were each treated (in replicates of 6) with 62 nM of type I (CT and LT-I) and type II (LT-IIa, LT-IIb, and LT-IIc) HLTs for 48 h, analyzed for viability by MTT assay, and normalized to the mean absorbance at 570 nM of untreated cells (NT) from the same tumor. Statistical significance was measured by one-way ANOVA followed by Tukey’s multiple comparison test. *** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 2
<p>The effects of LT-IIc versus CT A1:B5 holotoxins on cAMP accumulation and viability in MDA-MB-231 cells. MDA-MB-231 cells were treated with 62 nM LT-IIc or CT for 6, 12, 24, or 48 h and were analyzed for (<b>A</b>) intracellular cAMP accumulation measured by ELISA (n = 3), with the concentration of cAMP measured in untreated cells indicated by a dashed line; or (<b>B</b>) viability compared to untreated cells measured by MTT assay (n = 6). Data points represent mean +/− SEM. Statistical significance was measured by two-way ANOVA. ** <span class="html-italic">p</span> &lt;0.01, *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 3
<p>Effects of LT-IIc versus CT A1:B5 holotoxins on CREB phosphorylation. MDA-MB-231 cells were untreated (NT) or treated, in triplicate, with 62 nM LT-IIc or CT for 1, 3, and 6 h. P-CREB and CREB were measured in whole-cell lysates of MDA-MB-231 by immunoblotting. Blots were quantified using ImageJ software (<a href="https://imagej.net/ij/" target="_blank">https://imagej.net/ij/</a>). A representative blot is shown below graph of relative quantitation of P-CREB/CREB. Data points represent mean +/− SD. Statistical significance was measured by comparison to untreated cells using two-way ANOVA. * <span class="html-italic">p</span> &lt; 0.5; *** <span class="html-italic">p</span> &lt; 0.001; ns = not significant.</p>
Full article ">Figure 4
<p>The effects of ADP-ribosylation and the ganglioside binding activity of LT-IIc holotoxin variants versus B5 pentamer on cAMP accumulation and viability in MDA-MB-231 cells. (<b>A</b>) MDA-MB-231 cells were untreated (NT) or treated with 62 nM wt LT-IIc, LT-IIc(dblA), LT-IIc-B5, LT-IIc(T13I), or CT for 6 h, and were analyzed for intracellular cAMP accumulation measured by ELISA (n = 3). Statistical significance was measured by one-way ANOVA. *** <span class="html-italic">p</span> &lt; 0.001; ns = not significant. (<b>B</b>) MDA-MB-231 cells were treated with 0, 15.5, 31, 62, 124, and 248 nM wt LT-IIc, LT-IIc(dblA), LT-IIc-B5, LT-IIc(T13I), or CT for 48 h and viability was measured by MTT assay (n = 4). Data points represent mean +/− SD.</p>
Full article ">Figure 5
<p>Effects of the A1-domain of LT-IIc on cytotoxicity and binding affinity. (<b>A</b>) Cartoon representation of the A1-domain (light blue), A2-domain (yellow), and B-pentamer (blue) of LT-IIc holotoxin and LT-IIc-A2-Flg chimeric protein in which the A1-domain is replaced by a Flag epitope tag (orange). (<b>B</b>) MDA-MB-231 cells were treated with 62 nM LT-IIc, LT-IIc-A2-Flg, or LT-IIc-B5 for 48 h and analyzed for viability compared to untreated cells measured by MTT assay (n = 6). Data points represent mean +/− SEM. Statistical significance was measured by one-way ANOVA. *** <span class="html-italic">p</span> &lt; 0.0001. (<b>C</b>) Example of one of the four flow cytometric analyses measuring the effect of unlabeled LT-IIc (light blue) and unlabeled LT-IIc-A2-Flg (orange) on the mean fluorescent intensity of MDA-MB-231 cells stained with Alexa Fluor 488-labeled LT-IIc (red).</p>
Full article ">Figure 6
<p>Effects of T13I substitution on LT-IIc binding affinity toward MDA-MB-231. MDA-MB-231 cells were fixed to a 96-well plate and treated, in duplicate, with two-fold dilutions of a 1 μg/mL solution of wt LT-IIc or LT-IIc(T13I). HLTs bound to MDA-MB-231 were detected by ELISA. Data points represent mean +/− SD.</p>
Full article ">Figure 7
<p>Effects of eliglustat on ganglioside GD1a expression- and LT-IIc-dependent cytotoxicity on MDA-MB-231 cells. MDA-MB-231 cells were cultured in the absence or presence of 50 nM or 500 nM eliglustat for 4 days prior to (<b>A</b>) analysis of MOG35 immunostained cells measured by flow cytometry and (<b>B</b>) treatment with 31 nM LT-IIc for 24 h prior to analysis of viability measured by MTT assay (n = 6). Viability is calculated as % of untreated cells cultured under same eliglustat concentration. Data points represent mean +/− SEM. Statistical significance was measured by comparison to LT-IIc-treated cells not cultured with eliglustat using one-way ANOVA. * <span class="html-italic">p</span> &lt; 0.5; ns = no statistical significance.</p>
Full article ">Figure 7 Cont.
<p>Effects of eliglustat on ganglioside GD1a expression- and LT-IIc-dependent cytotoxicity on MDA-MB-231 cells. MDA-MB-231 cells were cultured in the absence or presence of 50 nM or 500 nM eliglustat for 4 days prior to (<b>A</b>) analysis of MOG35 immunostained cells measured by flow cytometry and (<b>B</b>) treatment with 31 nM LT-IIc for 24 h prior to analysis of viability measured by MTT assay (n = 6). Viability is calculated as % of untreated cells cultured under same eliglustat concentration. Data points represent mean +/− SEM. Statistical significance was measured by comparison to LT-IIc-treated cells not cultured with eliglustat using one-way ANOVA. * <span class="html-italic">p</span> &lt; 0.5; ns = no statistical significance.</p>
Full article ">
14 pages, 7964 KiB  
Article
Development and Clinical Application of a Molecular Assay for Four Common Porcine Enteroviruses
by Zhonghao Xin, Shiheng Li, Xiao Lu, Liping Liu, Yuehua Gao, Feng Hu, Kexiang Yu, Xiuli Ma, Yufeng Li, Bing Huang, Jiaqiang Wu and Xiaozhen Guo
Vet. Sci. 2024, 11(7), 305; https://doi.org/10.3390/vetsci11070305 - 9 Jul 2024
Viewed by 274
Abstract
Porcine epidemic diarrhea virus (PEDV), porcine transmissible gastroenteritis virus (TGEV), porcine deltacoronavirus (PDCoV), and porcine rotavirus-A (PoRVA) are the four main pathogens that cause viral diarrhea in pigs, and they often occur in mixed infections, which are difficult to distinguish only according to [...] Read more.
Porcine epidemic diarrhea virus (PEDV), porcine transmissible gastroenteritis virus (TGEV), porcine deltacoronavirus (PDCoV), and porcine rotavirus-A (PoRVA) are the four main pathogens that cause viral diarrhea in pigs, and they often occur in mixed infections, which are difficult to distinguish only according to clinical symptoms. Here, we developed a multiplex TaqMan-probe-based real-time RT-PCR method for the simultaneous detection of PEDV, TGEV, PDCoV, and PoRVA for the first time. The specific primers and probes were designed for the M protein gene of PEDV, N protein gene of TGEV, N protein gene of PDCoV, and VP7 protein gene of PoRVA, and corresponding recombinant plasmids were constructed. The method showed extreme specificity, high sensitivity, and excellent repeatability; the limit of detection (LOD) can reach as low as 2.18 × 102 copies/μL in multiplex real-time RT-PCR assay. A total of 97 clinical samples were used to compare the results of the conventional reverse transcription PCR (RT-PCR) and this multiplex real-time RT-PCR for PEDV, TGEV, PDCoV, and PoRVA detection, and the results were 100% consistent. Subsequently, five randomly selected clinical samples that tested positive were sent for DNA sequencing verification, and the sequencing results showed consistency with the detection results of the conventional RT-PCR and our developed method in this study. In summary, this study developed a multiplex real-time RT-PCR method for simultaneous detection of PEDV, TGEV, PDCoV, and PoRVA, and the results of this study can provide technical means for the differential diagnosis and epidemiological investigation of these four porcine viral diarrheic diseases. Full article
Show Figures

Figure 1

Figure 1
<p>Establishment of amplification curves and standard curves. (<b>a</b>–<b>d</b>) Amplification curves of PEDV, TGEV, PDCoV, and PoRVA. (<b>e</b>–<b>h</b>) Standard curves of PEDV, TGEV, PDCoV, and PoRVA.</p>
Full article ">Figure 1 Cont.
<p>Establishment of amplification curves and standard curves. (<b>a</b>–<b>d</b>) Amplification curves of PEDV, TGEV, PDCoV, and PoRVA. (<b>e</b>–<b>h</b>) Standard curves of PEDV, TGEV, PDCoV, and PoRVA.</p>
Full article ">Figure 2
<p>Probe and primer combinations at different concentrations. (<b>a</b>–<b>d</b>) Amplification curves of PEDV, TGEV, PDCoV, and PoRVA with different probe and primer concentrations. The pink lines are the optimal amplification curves, respectively.</p>
Full article ">Figure 2 Cont.
<p>Probe and primer combinations at different concentrations. (<b>a</b>–<b>d</b>) Amplification curves of PEDV, TGEV, PDCoV, and PoRVA with different probe and primer concentrations. The pink lines are the optimal amplification curves, respectively.</p>
Full article ">Figure 3
<p>Specificity assays of the multiplex real-time RT-PCR. The four amplification curves represent the four positive controls for PEDV, TGEV, PDCoV, and PoRVA. No fluorescent signal was observed for other swine pathogen samples and the negative control (NC).</p>
Full article ">Figure 4
<p>Sensitivity assay of the multiplex real−time RT−PCR. (<b>a</b>) Amplification curves are FAM−PEDV, Hex−TGEV, Texas Red−PDCoV, and Cy5−PoRVA, respectively. (<b>b</b>) Amplification curves were created by using the standard plasmid (2.18 × 10<sup>2</sup> copies/μL) of PEDV, TGEV, PDCoV, and PoRVA.</p>
Full article ">Figure 5
<p>Results of multiplex real-time RT-PCR for clinical samples. This Venn diagram, plotted via the (<a href="http://bioinformatics.com.cn" target="_blank">bioinformatics.com.cn</a>) website, shows individual infections and co-infections in the samples detected.</p>
Full article ">
20 pages, 3134 KiB  
Article
A Flagellin-Adjuvanted Trivalent Mucosal Vaccine Targeting Key Periodontopathic Bacteria
by Vandara Loeurng, Sao Puth, Seol Hee Hong, Yun Suhk Lee, Kamalakannan Radhakrishnan, Jeong Tae Koh, Joong-Ki Kook, Joon Haeng Rhee and Shee Eun Lee
Vaccines 2024, 12(7), 754; https://doi.org/10.3390/vaccines12070754 - 8 Jul 2024
Viewed by 401
Abstract
Periodontal disease (PD) is caused by microbial dysbiosis and accompanying adverse inflammatory responses. Due to its high incidence and association with various systemic diseases, disease-modifying treatments that modulate dysbiosis serve as promising therapeutic approaches. In this study, to simulate the pathophysiological situation, we [...] Read more.
Periodontal disease (PD) is caused by microbial dysbiosis and accompanying adverse inflammatory responses. Due to its high incidence and association with various systemic diseases, disease-modifying treatments that modulate dysbiosis serve as promising therapeutic approaches. In this study, to simulate the pathophysiological situation, we established a “temporary ligature plus oral infection model” that incorporates a temporary silk ligature and oral infection with a cocktail of live Tannerella forsythia (Tf), Pophyromonas gingivalis (Pg), and Fusobacterium nucleatum (Fn) in mice and tested the efficacy of a new trivalent mucosal vaccine. It has been reported that Tf, a red complex pathogen, amplifies periodontitis severity by interacting with periodontopathic bacteria such as Pg and Fn. Here, we developed a recombinant mucosal vaccine targeting a surface-associated protein, BspA, of Tf by genetically combining truncated BspA with built-in adjuvant flagellin (FlaB). To simultaneously induce Tf-, Pg-, and Fn-specific immune responses, it was formulated as a trivalent mucosal vaccine containing Tf-FlaB-tBspA (BtB), Pg-Hgp44-FlaB (HB), and Fn-FlaB-tFomA (BtA). Intranasal immunization with the trivalent mucosal vaccine (BtB + HB + BtA) prevented alveolar bone loss and gingival proinflammatory cytokine production. Vaccinated mice exhibited significant induction of Tf-tBspA-, Pg-Hgp44-, and Fn-tFomA-specific IgG and IgA responses in the serum and saliva, respectively. The anti-sera and anti-saliva efficiently inhibited epithelial cell invasion by Tf and Pg and interfered with biofilm formation by Fn. The flagellin-adjuvanted trivalent mucosal vaccine offers a novel method for modulating dysbiotic bacteria associated with periodontitis. This approach leverages the adjuvant properties of flagellin to enhance the immune response, aiming to restore a balanced microbial environment and improve periodontal health. Full article
18 pages, 4068 KiB  
Article
Type I Interferon Activates PD-1 Expression through Activation of the STAT1-IRF2 Pathway in Myeloid Cells
by Liyan Liang, Yingcui Yang, Kaidi Deng, Yanmin Wu, Yan Li, Liya Bai, Yinsong Wang and Chunwan Lu
Cells 2024, 13(13), 1163; https://doi.org/10.3390/cells13131163 - 8 Jul 2024
Viewed by 298
Abstract
PD-1 (Programmed cell death protein 1) regulates the metabolic reprogramming of myeloid-derived suppressor cells and myeloid cell differentiation, as well as the type I interferon (IFN-I) signaling pathway in myeloid cells in the tumor microenvironment. PD-1, therefore, is a key inhibitory receptor in [...] Read more.
PD-1 (Programmed cell death protein 1) regulates the metabolic reprogramming of myeloid-derived suppressor cells and myeloid cell differentiation, as well as the type I interferon (IFN-I) signaling pathway in myeloid cells in the tumor microenvironment. PD-1, therefore, is a key inhibitory receptor in myeloid cells. However, the regulation of PD-1 expression in myeloid cells is unknown. We report that the expression level of PDCD1, the gene that encodes the PD-1 protein, is positively correlated with the levels of IFNB1 and IFNAR1 in myeloid cells in human colorectal cancer. Treatment of mouse myeloid cell lines with recombinant IFNβ protein elevated PD-1 expression in myeloid cells in vitro. Knocking out IFNAR1, the gene that encodes the IFN-I-specific receptor, diminished the inductive effect of IFNβ on PD-1 expression in myeloid cells in vitro. Treatment of tumor-bearing mice with a lipid nanoparticle-encapsulated IFNβ-encoding plasmid (IFNBCOL01) increased IFNβ expression, resulting in elevated PD-1 expression in tumor-infiltrating myeloid cells. At the molecular level, we determined that IFNβ activates STAT1 (signal transducer and activator of transcription 1) and IRFs (interferon regulatory factors) in myeloid cells. Analysis of the cd279 promoter identified IRF2-binding consensus sequence elements. ChIP (chromatin immunoprecipitation) analysis determined that the pSTAT1 directly binds to the irf2 promoter and that IRF2 directly binds to the cd279 promoter in myeloid cells in vitro and in vivo. In colon cancer patients, the expression levels of STAT1, IRF2 and PDCD1 are positively correlated in tumor-infiltrating myeloid cells. Our findings determine that IFNβ activates PD-1 expression at least in part by an autocrine mechanism via the stimulation of the pSTAT1-IRF2 axis in myeloid cells. Full article
Show Figures

Figure 1

Figure 1
<p><b><span class="html-italic">PDCD1</span> and <span class="html-italic">IFNΒ1</span> expression profiles in human colorectal cancer patients.</b> (<b>A</b>) Shown are the UMAP of major cell subpopulations (<b>right panel</b>) and the <span class="html-italic">PDCD1</span> expression level (<b>left panel</b>) in the indicated cell subpopulations in human colorectal cancer. The red arrows show the <span class="html-italic">PDCD1</span> expression in myeloid cells. (<b>B</b>) Shown are the <span class="html-italic">IFNB1</span> expression level (<b>left panel</b>) in the indicated cell subpopulations and the UMAP of major cell subpopulations (<b>right panel</b>) in human colorectal cancer. The violin plot of <span class="html-italic">IFNB1</span> expression levels in the indicated major cell subpopulations is shown at the bottom (<b>left panel</b>). The red arrows show the <span class="html-italic">IFNB1</span> expression in myeloid cells. (<b>C</b>) Shown is the correlation between <span class="html-italic">IFNB1</span> and <span class="html-italic">PDCD1</span> in the cell subpopulations (<b>top panel</b>), as in A, and in the myeloid cell subpopulations (<b>bottom panel</b>). (<b>D</b>) Shown are correlations between <span class="html-italic">PDCD1</span> and <span class="html-italic">IFNAR1</span> in the myeloid cell subpopulations as shown in B.</p>
Full article ">Figure 2
<p><b>Myeloid cell intrinsic IFN-I controls PD-1 expression in vitro.</b> (<b>A</b>) Total RNA was prepared from RAW264.7 and bone marrow-derived MDSCs, and analyzed for basal IFNβ mRNA expression by RT-PCR with β-actin as an internal control. (<b>B</b>) RAW264.7 scramble and RAW264.7 IFNAR1 KO cells were cultured in vitro for 24 h. Cells were stained with anti-IFNAR1 mAb and analyzed by flow cytometry with IgG as a negative control. The MFI (mean fluorescence intensity) of IFNAR1 was quantified and presented in the right panel. Column: mean; Bar: SEM. (<b>C</b>) RAW264.7 scramble and RAW264.7 IFNAR1 KO cells were treated with IFNβ (100 ng/mL) for 24 h. Cells were stained with anti-PD-1 mAb and analyzed using flow cytometry. Data were analyzed with Dunnett’s test. (<b>D</b>) Total RNA was extracted from RAW264.7 scramble and RAW264.7 IFNAR1 KO cells treated with IFNβ for 4 h and 24 h, respectively, and analyzed using qPCR for IRF1-9 expression. Shown is one representative result of three independent experiments, and the error bar is the mean of the triplicates for one experiment. Data were analyzed with the Student <span class="html-italic">t</span> test. ns: no significance.</p>
Full article ">Figure 3
<p><b>Exogenous IFNβ activates STATs and IRFs to induce PD-1 expression in myeloid cells in vitro.</b> (<b>A</b>) RAW264.7 cells were treated with recombinant IFNα (100 ng/mL), IFNβ (100 ng/mL) and IFNα and IFNβ, respectively, for 24 h. LPS (2 ug/mL) is used as the positive control. Cells were stained with anti-PD-1 mAb and analyzed by flow cytometry. (<b>B</b>) The indicated RAW264.7 cells were treated with IFNβ alone for 24 h and then stained with anti-PD-1 mAb to be analyzed by flow cytometry. (<b>C</b>) Total RNA was isolated from RAW264.7 cells treated with IFNβ for 4 h and 24 h, respectively, and analyzed for the expression of PD-1 by qPCR. (<b>D</b>) RAW264.7 cells were treated with IFNβ for 4 h and 24 h, respectively, and lysed for total protein, then analyzed by Western blotting for the indicated STATs. β-actin is used as normalization control. (<b>E</b>) Total RNA was prepared from RAW264.7 cells treated with IFNβ for 4 h and 24 h, respectively, and analyzed for STAT1-6 expression using qPCR with β-actin as an internal normalization control. ns: no significance. (<b>F</b>,<b>G</b>) Total RNA was prepared from RAW264.7 cells treated with IFNβ for 4 h and 24 h, respectively, and analyzed for IRF1-9 expression by RT-PCR (<b>F</b>) and qPCR (<b>G</b>) with β-actin as internal normalization control. For RT-PCR, shown is one representative result of each of two independent experiments. For qPCR, shown is one representative result of each of three independent experiments, and the error bar is the mean of the triplicates for one experiment.</p>
Full article ">Figure 4
<p>Overexpression of IFNβ in the tumor microenvironment activates STATs and IRFs to stimulate PD-1 expression in myeloid cells in vivo. (<b>A</b>) CT26 cells (2.5 × 10<sup>5</sup> cells/mouse) were subcutaneously injected in mice to establish in vivo tumor models. The tumor-bearing mice were then intravenously treated with LNP (n = 5) and IFNBCOL01 (n = 5) 10 days after tumor inoculation, once every 3 days, totaling 4 times. Shown are the tumor images (top panel). The tumor volume and weight were quantified and shown in the bottom panels with LNP as a control. (<b>B</b>–<b>E</b>) Total RNA was prepared from the total tumor tissues, as shown in A, and analyzed for the expression of the indicated genes using qPCR with β-actin as an internal control. (<b>F</b>) Tumor tissues, as shown in A, were digested into single cells and used CD11b positive selection beads to isolate CD11b<sup>+</sup> myeloid cells. Tumor tissues before and after isolation were stained by anti-CD11b mAb and analyzed by flow cytometry. Shown is the percentage of CD11b<sup>+</sup> myeloid cells before isolation (<b>left panel</b>) and CD11b<sup>+</sup> myeloid cells in supernatant after isolation (<b>right panel</b>). (<b>G</b>–<b>I</b>) Total RNA was prepared from CD11b<sup>+</sup> myeloid cells as shown in E and analyzed for the mRNA level of the indicated genes using qPCR with β-actin as an internal normalization control. For qPCR, shown is one representative result of two independent experiments, and the error bar is the mean of the triplicates for one experiment. CT26 cells (2.5 × 10<sup>5</sup> cells/mouse) were harvested and subcutaneously injected into the right flank of BALB/c mice. Ten days after tumor cell inoculation, tumor-bearing mice were randomly separated into two groups. DOTAP-Cholesterol (4 mM LNP, 100 μL/mouse) or IFNBCOL01 (25 μg DNA in 4 mM LNP, 100 μL/mouse) were i.v. injected into mice in the two groups, respectively, once every 3 days for a total of 4 times.</p>
Full article ">Figure 5
<p><b>IRF2 binds to <span class="html-italic">cd279</span> promoter region in myeloid cells.</b> (<b>A</b>) The mouse <span class="html-italic">cd279</span> promoter DNA sequence (−2000 to +2000 relative to the <span class="html-italic">cd279</span> transcription start site) was exported from the mouse genomic database and analyzed for potential IRF-binding elements using the <span class="html-italic">PROMO</span> database. Shown is the mouse <span class="html-italic">cd279</span> promoter structure with predicted binding locations of IRF1 and IRF2 to the <span class="html-italic">cd279</span> promoter region. (<b>B</b>) Four pairs of primers spanning from −2000 to +2000 relative to the mouse <span class="html-italic">cd279</span> promoter region are shown. The red arrow represents the transcription start site. (<b>C</b>) RAW264.7 cells were treated with IFNβ for 24 h. Chromatin was then prepared and analyzed using ChIP with an IRF2-specific antibody. The immunoprecipitated chromatin fragments were then analyzed by qPCR with primers shown in (<b>B</b>). (<b>D</b>) Chromatin was prepared from IFNΒCOL01-treated CD11b<sup>+</sup> tumor-infiltrating myeloid cells shown in <a href="#cells-13-01163-f004" class="html-fig">Figure 4</a>F and analyzed using ChIP with a IRF2-specific antibody as in (<b>C</b>). The immunoprecipitated chromatin fragments were then analyzed with qPCR as in (<b>C</b>). For ChIP analysis, shown is one representative result of two independent experiments, and the error bar is the mean of the triplicates for one experiment. ns: no significance.</p>
Full article ">Figure 6
<p><b>pSTAT1 binds to the <span class="html-italic">irf2</span> promoter region in myeloid cells.</b> (<b>A</b>) The mouse <span class="html-italic">irf2</span> promoter DNA sequence (=−2000 to +2000 relative to <span class="html-italic">irf2</span> transcription start site) was exported from the mouse genomic database and analyzed for potential STAT-binding elements using the <span class="html-italic">PROMO</span> database. Shown is the <span class="html-italic">irf2</span> structure with predicted binding locations of STAT1β to the <span class="html-italic">irf2</span> promoter region. (<b>B</b>) Four pairs of primers spanning from −2000 to +2000 relative to the <span class="html-italic">irf2</span> promoter region are shown. The red arrow represents the transcription start site. (<b>C</b>) Chromatin was prepared from RAW264.7 cells treated with IFNβ for 24 h and analyzed using ChIP with a pSTAT1-specific antibody. The immunoprecipitated chromatin fragments were then analyzed using qPCR with primers shown in (<b>B</b>). (<b>D</b>) Chromatin was prepared from IFNΒCOL01-treated CD11b<sup>+</sup> tumor-infiltrating myeloid cells shown in <a href="#cells-13-01163-f004" class="html-fig">Figure 4</a>F and analyzed using ChIP with a pSTAT1-specific antibody, as in (<b>C</b>). The immunoprecipitated chromatin fragments were then analyzed using qPCR, as in (<b>C</b>). For ChIP analysis, shown is one representative result of two independent experiments, and the error bar is the mean of the triplicates for one experiment. ns: no significance.</p>
Full article ">Figure 7
<p><b>IFNβ-STAT1-IRF2 axis upregulates PD-1 expression in bone marrow-derived MDSCs.</b> (<b>A</b>) Bone marrow cells were prepared from BALB/c WT mice and treated with GM-CSF (2 ng/mL) for 7 days. Cells were then stained with anti-CD11b mAb and analyzed using flow cytometry. The percentage of CD11b<sup>+</sup> myeloid cells after GM-CSF treatment (right pane) is shown. (<b>B</b>) Bone marrow-derived MDSCs were cultured for 24 h and then stained with anti-IFNAR1 mAb and analyzed by flow cytometry. (<b>C</b>) Bone marrow-derived MDSCs were treated with recombinant IFNα, IFNβ and IFNα+IFNβ, respectively, for 24 h. LPS was used as the positive control. Cells were then stained with anti-PD-1 mAb and analyzed by flow cytometry. ns: no significance. (<b>D</b>) The bone marrow-derived MDSCs were treated with IFNβ alone for 24 h, and then stained with anti-PD-1 mAb and analyzed by flow cytometry. (<b>E</b>) Total RNA was isolated from bone marrow-derived MDSCs treated with IFNβ for 4 h and 24 h, respectively, and analyzed for the mRNA level of PD-1 using qPCR with β-actin as the internal control. (<b>F</b>) Bone marrow-derived MDSCs were treated with IFNβ for 4 h and 24 h, respectively, and lysed for total protein, and then analyzed by Western blotting for STAT1 and pSTAT1. β-actin is used as the normalization control. (<b>G</b>) Total RNA was prepared from bone marrow-derived MDSCs treated with IFNβ for 4 h and 24 h, respectively, and analyzed for the expression of IRF2 using qPCR. (<b>H</b>,<b>I</b>) Chromatin was isolated from bone marrow-derived MDSCs treated with IFNβ for 24 h and analyzed by ChIP with an IRF2-specific antibody, (<b>H</b>) as in <a href="#cells-13-01163-f005" class="html-fig">Figure 5</a>C, and a pSTAT1-specific antibody, (<b>I</b>) as in <a href="#cells-13-01163-f006" class="html-fig">Figure 6</a>C. The immunoprecipitated chromatin fragments were then analyzed using qPCR with primers shown in <a href="#cells-13-01163-f005" class="html-fig">Figure 5</a>B and <a href="#cells-13-01163-f006" class="html-fig">Figure 6</a>B, respectively. For ChIP analysis, shown is one representative result of two independent experiments, and the error bar is the mean of the triplicates for one experiment.</p>
Full article ">
22 pages, 1472 KiB  
Article
Effects of Recombinant α1-Microglobulin on Early Proteomic Response in Risk Organs after Exposure to 177Lu-Octreotate
by Charlotte Ytterbrink, Emman Shubbar, Toshima Z. Parris, Britta Langen, Malin Druid, Emil Schüler, Sven-Erik Strand, Bo Åkerström, Magnus Gram, Khalil Helou and Eva Forssell-Aronsson
Int. J. Mol. Sci. 2024, 25(13), 7480; https://doi.org/10.3390/ijms25137480 - 8 Jul 2024
Viewed by 304
Abstract
Recombinant α1-microglobulin (A1M) is proposed as a protector during 177Lu-octreotate treatment of neuroendocrine tumors, which is currently limited by bone marrow and renal toxicity. Co-administration of 177Lu-octreotate and A1M could result in a more effective treatment by protecting healthy [...] Read more.
Recombinant α1-microglobulin (A1M) is proposed as a protector during 177Lu-octreotate treatment of neuroendocrine tumors, which is currently limited by bone marrow and renal toxicity. Co-administration of 177Lu-octreotate and A1M could result in a more effective treatment by protecting healthy tissue, but the radioprotective action of A1M is not fully understood. The aim of this study was to examine the proteomic response of kidneys and bone marrow early after 177Lu-octreotate and/or A1M administration. Mice were injected with 177Lu-octreotate and/or A1M, while control mice received saline or A1M vehicle solution. Bone marrow, kidney medulla, and kidney cortex were sampled after 24 h or 7 d. The differential protein expression was analyzed with tandem mass spectrometry. The dosimetric estimation was based on 177Lu activity in the kidney. PHLDA3 was the most prominent radiation-responsive protein in kidney tissue. In general, no statistically significant difference in the expression of radiation-related proteins was observed between the irradiated groups. Most canonical pathways were identified in bone marrow from the 177Lu-octreotate+A1M group. Altogether, a tissue-dependent proteomic response followed exposure to 177Lu-octreotate alone or together with A1M. Combining 177Lu-octreotate with A1M did not inhibit the radiation-induced protein expression early after exposure, and late effects should be further studied. Full article
(This article belongs to the Section Molecular Informatics)
Show Figures

Figure 1

Figure 1
<p>The total number of differentially regulated proteins in mouse tissues after exposure to <sup>177</sup>Lu-octreoate, <sup>177</sup>Lu-octreoate + A1M, or A1M only. Venn diagrams show unique and commonly expressed proteins with magnification showing the number of upregulated (↑) and downregulated (↓) proteins in (<b>A</b>) kidney cortex at 24 h, (<b>B</b>) kidney cortex at 7 d, (<b>C</b>) kidney medulla at 24 h, (<b>D</b>) kidney medulla at 7 d, (<b>E</b>) bone marrow at 24 h, and (<b>F</b>) bone marrow at 7 d.</p>
Full article ">Figure 2
<p>Proteins with significant regulation compared with control (|FC| ≥ 1.5) together with statistically significant differences in regulation between any of the groups (ANOVA, 5% FDR) at both time points in (<b>A</b>) kidney cortex, (<b>B</b>) kidney medulla, and (<b>C</b>) bone marrow. Error bars show the standard deviation, and brackets show statistically significant differences (<span class="html-italic">p</span> &lt; 0.05). * indicates the biological function annotations for each protein, given by the Proteome Discoverer.</p>
Full article ">
13 pages, 4237 KiB  
Article
Plant-Produced Chimeric Hepatitis E Virus-like Particles as Carriers for Antigen Presentation
by Eugenia S. Mardanova, Egor A. Vasyagin, Kira G. Kotova, Gergana G. Zahmanova and Nikolai V. Ravin
Viruses 2024, 16(7), 1093; https://doi.org/10.3390/v16071093 - 8 Jul 2024
Viewed by 628
Abstract
A wide range of virus-like particles (VLPs) is extensively employed as carriers to display various antigens for vaccine development to fight against different infections. The plant-produced truncated variant of the hepatitis E virus (HEV) coat protein is capable of forming VLPs. In this [...] Read more.
A wide range of virus-like particles (VLPs) is extensively employed as carriers to display various antigens for vaccine development to fight against different infections. The plant-produced truncated variant of the hepatitis E virus (HEV) coat protein is capable of forming VLPs. In this study, we demonstrated that recombinant fusion proteins comprising truncated HEV coat protein with green fluorescent protein (GFP) or four tandem copies of the extracellular domain of matrix protein 2 (M2e) of influenza A virus inserted at the Tyr485 position could be efficiently expressed in Nicotiana benthamiana plants using self-replicating vector based on the potato virus X genome. The plant-produced fusion proteins in vivo formed VLPs displaying GFP and 4M2e. Therefore, HEV coat protein can be used as a VLP carrier platform for the presentation of relatively large antigens comprising dozens to hundreds of amino acids. Furthermore, plant-produced HEV particles could be useful research tools for the development of recombinant vaccines against influenza. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Scheme of the expression vectors. RDRP, RNA-dependent RNA polymerase gene; Sgp1, the first promoter of subgenomic RNA of PVX; AMV, translational enhancer from alfalfa mosaic virus; 35S, promoter of the cauliflower mosaic virus RNA; Nos-T, terminator of the <span class="html-italic">A. tumefaciens</span> nopaline synthase gene; P24, suppressor of silencing from grapevine leafroll-associated virus-2; SP, signal peptide; HDEL, ER retention signal; RB and LB, the right and left borders of T-DNA region, respectively. The insertion position of the target gene I is indicated by the red arrow. (<b>b</b>) Scheme of the recombinant proteins. 4M2e, the sequence of four tandem copies of M2e peptide; GFP, green fluorescent protein; HEV, hepatitis E virus ORF2. (<b>c</b>) 3D modeling of virus-like particles formed by HEV ORF2 capsid, HEV/4M2e, and HEV/GFP proteins by SWISS MODEL.</p>
Full article ">Figure 2
<p>Expression of HEV/GFP protein in <span class="html-italic">N. benthamiana</span> plants. Coomassie brilliant blue-stained gel (<b>a</b>,<b>b</b>) and Western blotting with antibodies against GFP (<b>c</b>) and hexahistidine tag (<b>d</b>) of proteins isolated from plants. M, molecular weight marker (kDa); Lanes: 1, total proteins isolated from non-infiltrated leaf; 2, total proteins isolated from leaf infiltrated with pEff-HEV/GFP; 3–5, purified HEV/GFP protein. Position of HEV/GFP protein is shown by an arrow. Visualization GFP fluorescence in <span class="html-italic">N. benthamiana</span> leaves infiltrated with pEff_HEV/GFP and pEff-GFP in UV (<b>e</b>).</p>
Full article ">Figure 3
<p>Expression of HEV and HEV/4M2e proteins in <span class="html-italic">N. benthamiana</span> plants. Coomassie brilliant blue-stained gel (<b>a</b>,<b>b</b>) and Western blotting with antibodies against M2e (<b>c</b>) of proteins isolated from plants. M, molecular weight marker (kD); EL, total proteins isolated from the non-infiltrated leaf; Lanes: 1, total proteins isolated from leaf infiltrated with pAeff-HEV; 2, total proteins isolated from leaf infiltrated with pAeff-HEV/4M2e; 3, total proteins isolated from leaf infiltrated with pEff-HEV/4M2e; 4, total proteins isolated from leaf infiltrated with pEff-HEV; 5, purified HEV protein (expression using pAeff vector); 6, purified HEV/4M2e protein (expression using pAeff vector); 7, purified HEV/4M2e protein (expression using pEff vector); 8, purified HEV protein (expression using pEff vector). Visualization of proteins synthesized in <span class="html-italic">N. benthamiana</span> leaves infiltrated with pAeff-HEV, pAeff-HEV/4M2e, pEff-HEV/4M2e and pEff-HEV (<b>d</b>). The positions of HEV and HEV/4M2e proteins in panel (<b>a</b>) are marked with an asterisk (*).</p>
Full article ">Figure 4
<p>Analysis of virus-like particles formed by HEV/GFP and HEV/4M2e proteins by transmission electron microscopy (EM) and atomic force microscopy (AFM). The color scale shows particle height.</p>
Full article ">Figure 5
<p>Antigenicity of the recombinant proteins. Two-fold dilutions of recombinant HEV, HEV/4M2e, and HEV/GFP proteins (starting from 10 μg/mL) were used to coat ELISA plates which were then probed with antibodies specific for GFP (<b>a</b>) or M2e (<b>b</b>).</p>
Full article ">
18 pages, 3364 KiB  
Article
The Bursaphelenchus xylophilus Effector BxNMP1 Targets PtTLP-L2 to Mediate PtGLU Promoting Parasitism and Virulence in Pinus thunbergii
by Dan Yang, Lin Rui, Yi-Jun Qiu, Tong-Yue Wen, Jian-Ren Ye and Xiao-Qin Wu
Int. J. Mol. Sci. 2024, 25(13), 7452; https://doi.org/10.3390/ijms25137452 - 7 Jul 2024
Viewed by 312
Abstract
Pinus is an important economic tree species, but pine wilt disease (PWD) seriously threatens the survival of pine trees. PWD caused by Bursaphelenchus xylophilus is a major quarantine disease worldwide that causes significant economic losses. However, more information about its molecular pathogenesis is [...] Read more.
Pinus is an important economic tree species, but pine wilt disease (PWD) seriously threatens the survival of pine trees. PWD caused by Bursaphelenchus xylophilus is a major quarantine disease worldwide that causes significant economic losses. However, more information about its molecular pathogenesis is needed, resulting in a lack of effective prevention and treatment measures. In recent years, effectors have become a hot topic in exploring the molecular pathogenic mechanism of pathogens. Here, we identified a specific effector, BxNMP1, from B. xylophilus. In situ hybridization experiments revealed that BxNMP1 was specifically expressed in dorsal gland cells and intestinal cells, and RT–qPCR experiments revealed that BxNMP1 was upregulated in the early stage of infection. The sequence of BxNMP1 was different in the avirulent strain, and when BxNMP1-silenced B. xylophilus was inoculated into P. thunbergii seedlings, the disease severity significantly decreased. We demonstrated that BxNMP1 interacted with the thaumatin-like protein PtTLP-L2 in P. thunbergii. Additionally, we found that the β-1,3-glucanase PtGLU interacted with PtTLP-L2. Therefore, we hypothesized that BxNMP1 might indirectly interact with PtGLU through PtTLP-L2 as an intermediate mediator. Both targets can respond to infection, and PtTLP-L2 can enhance the resistance of pine trees. Moreover, we detected increased salicylic acid contents in P. thunbergii seedlings inoculated with B. xylophilus when BxNMP1 was silenced or when the PtTLP-L2 recombinant protein was added. In summary, we identified a key virulence effector of PWNs, BxNMP1. It positively regulates the pathogenicity of B. xylophilus and interacts directly with PtTLP-L2 and indirectly with PtGLU. It also inhibits the expression of two targets and the host salicylic acid pathway. This study provides theoretical guidance and a practical basis for controlling PWD and breeding for disease resistance. Full article
(This article belongs to the Section Molecular Toxicology)
Show Figures

Figure 1

Figure 1
<p>Expression pattern of <span class="html-italic">BxNMP1</span>. (<b>a</b>) The relative expression level of <span class="html-italic">BxNMP1</span> at nine time points after <span class="html-italic">Bursaphelenchus</span> xylophilus infection was determined using reverse transcription-quantitative PCR (RT–qPCR) analysis. Nematodes were collected from entire <span class="html-italic">Pinus thunbergii</span> seedlings inoculated with <span class="html-italic">B. xylophilus</span> (isolating AMA3) at various time points. The relative expression level of <span class="html-italic">BxNMP1</span> was calculated using the comparative threshold method. The RT–qPCR values were normalized to the transcript level of <span class="html-italic">Actin</span>. The values represent the means ± standard deviations of three independent biological samples. Different letters indicate statistically significant differences according to Duncan’s multiple range test (<span class="html-italic">p</span> &lt; 0.05). (<b>b</b>) Localization of BxNMP1 in the dorsal glands (DGs) and intestine via in situ hybridization. DIG: digoxigenin; M: median bulb. Scale bars = 50 µm.</p>
Full article ">Figure 2
<p>BxNMP1 contributes to the virulence of <span class="html-italic">B. xylophilus</span>. (<b>a</b>) Inoculation assay of <span class="html-italic">Pinus thunbergii</span> seedlings. The degree of morbidity of the <span class="html-italic">P. thunbergii</span> seedlings varied according to the color of the needle leaves. At 16 days postinoculation (dpi), four <span class="html-italic">P. thunbergii</span> seedlings inoculated with <span class="html-italic">siGFP</span>-treated <span class="html-italic">B. xylophilus</span> turned yellow, but none of the <span class="html-italic">P. thunbergii</span> seedlings inoculated with <span class="html-italic">siBxNMP1</span>-treated <span class="html-italic">B. xylophilus</span> turned yellow. At 20 dpi, all the <span class="html-italic">P. thunbergii</span> seedlings inoculated with siGFP-treated <span class="html-italic">B. xylophilus</span> turned yellow or brown, and one <span class="html-italic">P. thunbergii</span> seedling inoculated with <span class="html-italic">siBxNMP1</span>-treated <span class="html-italic">B. xylophilus</span> turned yellow. (<b>b</b>) The infection rates of <span class="html-italic">P. thunbergii</span> seedlings under different treatments. (<b>c</b>) The disease severity index of <span class="html-italic">P. thunbergii</span> seedlings under various treatments. (<b>d</b>) The relative transcript levels of pathogenesis-related genes in <span class="html-italic">P. thunbergii</span>-infected <span class="html-italic">siBxNMP1</span>-treated nematodes were upregulated compared with those in control nematodes. Stems approximately 1 cm in length were selected for RNA extraction at 12 h postinoculation. The data are presented as the means ± standard deviations (SD) from three biological replicates. Different letters indicate significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>BxNMP1 interacts with PtTLP-L2 in <span class="html-italic">Pinus thunbergii</span>. (<b>a</b>) BxNMP1 interacts with PtTLP-L2 in yeast. The Y2H Gold yeast strain cotransformed with BD-BxNMP1 and AD-PtTLP-L2 was cultured on SD/−Trp/−Leu media and then on SD/−Trp/−Leu/−His/−Ade+X−α−Gal+AbA selective media. The images were captured using a four-fold magnification of microscope. (<b>b</b>) BxNMP1 interacts with PtTLP-L2 in vivo. Coimmunoprecipitation (Co-IP) was performed on extracts of <span class="html-italic">Nicotiana benthamiana</span> leaves expressing both BxNMP1-RFP and PtTLP-L2-GFP. Green fluorescent protein (GFP) was detected via Western blot using anti-GFP antibodies. Red fluorescent protein (RFP) was detected via Western blot using anti-RFP antibodies. The immune complexes were pulled down using anti-GFP agarose beads.</p>
Full article ">Figure 4
<p>PtGLU interacts with <span class="html-italic">Pinus thunbergii</span> TLP-L2 proteins. (<b>a</b>) PtGLU interacts with Pt TLP-L2 in yeast. The Y2H Gold yeast strain cocarrying BD-PtGlu and AD-Pt-TLP-L2 was grown on SD/−Trp/−Leu and the selective medium SD/−Trp/−Leu/−His/−Ade+X−α−Gal+AbA. The images were captured using a four-fold magnification of microscope. (<b>b</b>) PtGLU interacts with PtTLP-L2 in vivo. Coimmunoprecipitation (Co-IP) was performed on extracts of <span class="html-italic">Nicotiana benthamiana</span> leaves expressing both PtGLU-RFP and PtTLP-L2-GFP. Green fluorescent protein (GFP) was detected via Western blot using anti-GFP antibodies. Red fluorescent protein (RFP) was detected via Western blot using anti-RFP antibodies. The immune complexes were pulled down using anti-GFP agarose beads.</p>
Full article ">Figure 5
<p>PtTLP-L2 colocalizes with BxNMP1 and PtGLU in the nucleus and cytoplasm. Proteins were expressed in <span class="html-italic">Nicotiana benthamiana</span> via agroinfiltration. (<b>a</b>) Confocal microscopy imaging of <span class="html-italic">N. benthamiana</span> leaves transiently expressing green fluorescent protein (GFP)-tagged PtTLP-L2, red fluorescent protein (RFP)-tagged BxNMP1 and PtGLU, showing that PtTLP-L2 colocalizes with BxNMP1 and PtGLU in both the nucleus and the cytoplasm. Pictures taken 48 h postinfiltration show cells cotransformed with PtTLP-L2 (green channel; <b>middle left panel</b>) and BxNMP1 and PtGLU (red channel; <b>left panel</b>). Bright field images (<b>middle right panel</b>) and the overlay (<b>right panel</b>) are also shown. Scale bars, 10 μm. (<b>b</b>) Expression of GFP-PtTLP-L2 together with RFP-BxNMP1 and RFP-PtGLU was confirmed via Western blotting using anti-GFP and anti-RFP antibodies. Protein loading is indicated with Ponceau S staining of RuBisCO.</p>
Full article ">Figure 6
<p>Expression pattern of <span class="html-italic">PtTLP-L2</span> and <span class="html-italic">PtGLU</span> in infected <span class="html-italic">Pinus thunbergii,</span> and PtTLP-L2 can enhance the resistance of <span class="html-italic">P. thunbergii</span>. (<b>a</b>) The expression of <span class="html-italic">PtTLP-L2</span> in infected <span class="html-italic">Pinus thunbergii</span> at different time points. (<b>b</b>) The expression of <span class="html-italic">PtGLU</span> in infected <span class="html-italic">P. thunbergii</span> at different time points. In total, 2000 PWNs were inoculated into 3-year-old <span class="html-italic">P. thunbergii</span> seedlings. Stems approximately 1 cm in length were collected to analyze the relative expression of <span class="html-italic">PtTLP-L2</span> and <span class="html-italic">PtGLU</span>. (<b>c</b>) The symptoms of <span class="html-italic">P. thunbergii</span> at 12 and 18 days postinoculation (dpi) with <span class="html-italic">Bursaphelenchus xylophilus</span> and two different purified recombinant proteins (EVrec: pET32a and PtTLP-L2rec). (<b>d</b>,<b>e</b>) The infection rate and disease severity index of <span class="html-italic">P. thunbergii</span> seedlings were calculated at 12 and 18 dpi. Three independent experiments were performed, and at least 3 individual <span class="html-italic">P. thunbergii</span> seedlings were used for each treatment. The data are presented as the means ± standard deviations (SD) from three experiments. Different letters indicate significant differences (<span class="html-italic">p &lt;</span> 0.05).</p>
Full article ">Figure 7
<p>Changes in the activity of the salicylic acid pathway in <span class="html-italic">Pinus thunbergii</span> under different treatments. In total, 2000 <span class="html-italic">siBxNMP1</span>-treated nematodes, <span class="html-italic">siGFP</span>-treated nematodes, a total of 200 µg of purified PtTLP-L2 recombinant protein (PtTLP-L2rec) or EV recombinant protein (EVrec), and untreated PWNs were inoculated into 3-year-old <span class="html-italic">P. thunbergii</span> seedlings. Stems approximately 1 cm in length were selected for measurement of enzyme activity or for extraction of RNA at 12 h postinoculation. (<b>a</b>) Relative expression of <span class="html-italic">PtTLP-L2</span> and <span class="html-italic">PtGlu</span> in <span class="html-italic">P. thunbergii</span> seedlings under four treatments. (<b>b</b>–<b>d</b>) Contents of salicylic acid, PAL, and GOGAT activity in <span class="html-italic">P. thunbergii</span> seedlings under four treatments. The values represent the means ± SDs of three independent biological samples. Different letters on top of the bars indicate statistically significant differences (<span class="html-italic">p</span> &lt; 0.05, <span class="html-italic">t</span> test), as measured with Duncan’s multiple range test.</p>
Full article ">
15 pages, 2025 KiB  
Article
Metabolic and Transcriptional Analysis Reveals Flavonoid Involvement in the Drought Stress Response of Mulberry Leaves
by Guo Chen, Dong Li, Pei Yao, Fengyao Chen, Jianglian Yuan, Bi Ma, Zhen Yang, Biyue Ding and Ningjia He
Int. J. Mol. Sci. 2024, 25(13), 7417; https://doi.org/10.3390/ijms25137417 - 6 Jul 2024
Viewed by 260
Abstract
Abiotic stress, especially drought stress, poses a significant threat to terrestrial plant growth, development, and productivity. Although mulberry has great genetic diversity and extensive stress-tolerant traits in agroforestry systems, only a few reports offer preliminary insight into the biochemical responses of mulberry leaves [...] Read more.
Abiotic stress, especially drought stress, poses a significant threat to terrestrial plant growth, development, and productivity. Although mulberry has great genetic diversity and extensive stress-tolerant traits in agroforestry systems, only a few reports offer preliminary insight into the biochemical responses of mulberry leaves under drought conditions. In this study, we performed a comparative metabolomic and transcriptomic analysis on the “drooping mulberry” (Morus alba var. pendula Dippel) under PEG-6000-simulated drought stress. Our research revealed that drought stress significantly enhanced flavonoid accumulation and upregulated the expression of phenylpropanoid biosynthetic genes. Furthermore, the activities of superoxide dismutase (SOD), catalase (CAT) and malondialdehyde (MDA) content were elevated. In vitro enzyme assays and fermentation tests indicated the involvement of flavonol synthase/flavanone 3-hydroxylase (XM_010098126.2) and anthocyanidin 3-O-glucosyltransferase 5 (XM_010101521.2) in the biosynthesis of flavonol aglycones and glycosides, respectively. The recombinant MaF3GT5 protein was found to recognize kaempferol, quercetin, and UDP-glucose as substrates but not 3-/7-O-glucosylated flavonols and UDP-rhamnose. MaF3GT5 is capable of forming 3-O- and 7-O-monoglucoside, but not di-O-glucosides, from kaempferol. This implies its role as a flavonol 3, 7-O-glucosyltransferase. The findings from this study provided insights into the biosynthesis of flavonoids and could have substantial implications for the future diversified utilization of mulberry. Full article
(This article belongs to the Special Issue Drought Stress Tolerance in Plants in 2024)
Show Figures

Figure 1

Figure 1
<p>Variation in catalase (CAT), superoxide dismutase (SOD), and malondialdehyde (MDA) levels at different hours under drought stress in <span class="html-italic">Morus alba</span> var. <span class="html-italic">pendula</span> Dippel leaves. Data represent the mean of three replicates with standard deviation (±SD).</p>
Full article ">Figure 2
<p>Metabolic profiling and flavonoid variation in <span class="html-italic">M. alba</span> var. <span class="html-italic">pendula</span> leaves under drought stress. (<b>A</b>) LC/MS profiling of flavonoids at 0, 24, 48, and 72 h in mulberry leaves under drought treatment, each with three biological replicates. The heat-map shows values displayed on log2 of the relative peak area. (<b>B</b>) Fold changes of flavonoids at 24, 48, and 72 h. Increased flavonoids are indicated in cyan color, while decreased flavonoids are indicated in blue.</p>
Full article ">Figure 3
<p>The proposed flavonoid pathway and heat-map of key genes involved in the drought response of mulberry leaves. Colored blocks indicated different relative expression levels (from blue to red). C1 to C3 represents three replicates of the control, while D1 to D3 represents the drought-treated group. Abbreviations: <span class="html-italic">PAL</span>, phenylalanine ammonia lyase; <span class="html-italic">C4H</span>, cinnamate 4-hydroxylase; <span class="html-italic">4CL</span>, 4-coumarate-CoA ligase; <span class="html-italic">CHS</span>, chalcone synthase; <span class="html-italic">CHI</span>, chalcone isomerase; <span class="html-italic">FNS</span>, flavone synthase; <span class="html-italic">F3H</span>, flavanone 3-hydroxylase; <span class="html-italic">F3</span>′<span class="html-italic">H</span>, flavanone 3′-hydroxylase; <span class="html-italic">FLS</span>, flavonol synthase; <span class="html-italic">F3GT</span>, flavonoid 3-O-glycosyltransferase; <span class="html-italic">F3G6”RT</span>, flavonol 3-<span class="html-italic">O</span>-glucoside: 6″-<span class="html-italic">O</span>-rhamnosyltransferase; K3G, kaempferol 3-<span class="html-italic">O</span>-glucoside; Q3G, quercetin 3-<span class="html-italic">O</span>-glucoside; K3R, kaempferol 3-<span class="html-italic">O</span>-rutinoside.</p>
Full article ">Figure 4
<p>Transcript levels of phenylpropanoid biosynthetic genes and candidate genes under PEG-6000-induced drought stress. Data represent the mean of three replicates with standard deviation (±SD).</p>
Full article ">Figure 5
<p>UHPLC-MS/MS analyses of the reaction of recombinant MaFLS1 protein and fermented products. (<b>A</b>) Elution profile of the reaction products of His tag protein (empty vector), MaFLS1 protein (+MaFLS1), and authentic standards. The chromatograms of fermentation extracts present a similar pattern to the in vitro enzyme assay. (<b>B</b>) Extracted fragment mass chromatograms of reaction products (N1 and N2). Abbreviations: DHK, dihydrokaempferol; DHQ, dihydroquercetin; K, kaempferol; Q, quercetin.</p>
Full article ">Figure 6
<p>HPLC-MS/MS analyses of the reaction of recombinant MaF3GT5 protein. The elution profile of the reaction products involving His tag protein (empty vector), MaF3GT5 protein (+MaF3GT5), and chromatograms of standards (K, Q, K3G, Q3G, K7G, and Q7G) are shown on the left. Extracted fragment mass chromatograms of reaction products (F1, F2, and F3) are shown on the right. Abbreviations: K, kaempferol; Q, quercetin; K3G, kaempferol 3-<span class="html-italic">O</span>-glucoside; Q3G, quercetin 3-<span class="html-italic">O</span>-glucoside; K7G, kaempferol 7-<span class="html-italic">O</span>-glucoside; Q7G, quercetin 7-<span class="html-italic">O</span>-glucoside; -glu, natural loss of glucoside.</p>
Full article ">
15 pages, 945 KiB  
Review
Effects of Lipoproteins on Metabolic Health
by Obaida Albitar, Crystal M. D’Souza and Ernest A. Adeghate
Nutrients 2024, 16(13), 2156; https://doi.org/10.3390/nu16132156 - 6 Jul 2024
Viewed by 512
Abstract
Lipids are primarily transported in the bloodstream by lipoproteins, which are macromolecules of lipids and conjugated proteins also known as apolipoproteins. The processes of lipoprotein assembly, secretion, transportation, modification, and clearance are crucial components of maintaining a healthy lipid metabolism. Disruption in any [...] Read more.
Lipids are primarily transported in the bloodstream by lipoproteins, which are macromolecules of lipids and conjugated proteins also known as apolipoproteins. The processes of lipoprotein assembly, secretion, transportation, modification, and clearance are crucial components of maintaining a healthy lipid metabolism. Disruption in any of these steps results in pathophysiological abnormalities such as dyslipidemia, obesity, insulin resistance, inflammation, atherosclerosis, peripheral artery disease, and cardiovascular diseases. By studying these genetic mutations, researchers can gain valuable insights into the underlying mechanisms that govern the relationship between protein structure and its physiological role. These lipoproteins, including HDL, LDL, lipoprotein(a), and VLDL, mainly serve the purpose of transporting lipids between tissues and organs. However, studies have provided evidence that apo(a) also possesses protective properties against pathogens. In the future, the field of study will be significantly influenced by the integration of recombinant DNA technology and human site-specific mutagenesis for treating hereditary disorders. Several medications are available for the treatment of dyslipoproteinemia. These include statins, fibrates, ezetimibe, niacin, PCSK9 inhibitors, evinacumab, DPP 4 inhibitors, glucagon-like peptide-1 receptor agonists GLP1RAs, GLP-1, and GIP dual receptor agonists, in addition to SGLT2 inhibitors. This current review article exhibits, for the first time, a comprehensive reflection of the available body of publications concerning the impact of lipoproteins on metabolic well-being across various pathological states. Full article
(This article belongs to the Section Nutrition and Public Health)
Show Figures

Figure 1

Figure 1
<p>Illustrative representation of the chemical structure of lipoproteins.</p>
Full article ">Figure 2
<p>Graphical representation of the effect of impaired lipoprotein metabolism of cardiovascular diseases.</p>
Full article ">
16 pages, 8530 KiB  
Article
Expression and Functional Analysis of the Metallothionein and Metal-Responsive Transcription Factor 1 in Phascolosoma esculenta under Zn Stress
by Shenwei Gu, Jingqian Wang, Xinming Gao, Xuebin Zheng, Yang Liu, Yiner Chen, Lianlian Sun and Junquan Zhu
Int. J. Mol. Sci. 2024, 25(13), 7368; https://doi.org/10.3390/ijms25137368 - 5 Jul 2024
Viewed by 298
Abstract
Metallothioneins (MTs) are non-enzymatic metal-binding proteins widely found in animals, plants, and microorganisms and are regulated by metal-responsive transcription factor 1 (MTF1). MT and MTF1 play crucial roles in detoxification, antioxidation, and anti-apoptosis. Therefore, they are key factors allowing organisms to endure the [...] Read more.
Metallothioneins (MTs) are non-enzymatic metal-binding proteins widely found in animals, plants, and microorganisms and are regulated by metal-responsive transcription factor 1 (MTF1). MT and MTF1 play crucial roles in detoxification, antioxidation, and anti-apoptosis. Therefore, they are key factors allowing organisms to endure the toxicity of heavy metal pollution. Phascolosoma esculenta is a marine invertebrate that inhabits intertidal zones and has a high tolerance to heavy metal stress. In this study, we cloned and identified MT and MTF1 genes from P. esculenta (designated as PeMT and PeMTF1). PeMT and PeMTF1 were widely expressed in all tissues and highly expressed in the intestine. When exposed to 16.8, 33.6, and 84 mg/L of zinc ions, the expression levels of PeMT and PeMTF1 in the intestine increased first and then decreased, peaking at 12 and 6 h, respectively, indicating that both PeMT and PeMTF1 rapidly responded to Zn stress. The recombinant pGEX-6p-1-MT protein enhanced the Zn tolerance of Escherichia coli and showed a dose-dependent ABTS free radical scavenging ability. After RNA interference (RNAi) with PeMT and 24 h of Zn stress, the oxidative stress indices (MDA content, SOD activity, and GSH content) and the apoptosis indices (Caspase 3, Caspase 8, and Caspase 9 activities) were significantly increased, implying that PeMT plays an important role in Zn detoxification, antioxidation, and anti-apoptosis. Moreover, the expression level of PeMT in the intestine was significantly decreased after RNAi with PeMTF1 and 24 h of Zn stress, which preliminarily proved that PeMTF1 has a regulatory effect on PeMT. Our data suggest that PeMT and PeMTF1 play important roles in the resistance of P. esculenta to Zn stress and are the key factors allowing P. esculenta to endure the toxicity of Zn. Full article
(This article belongs to the Section Molecular Toxicology)
Show Figures

Figure 1

Figure 1
<p>Full-length cloning and bioinformatic analysis of <span class="html-italic">Pe</span>MT. (<b>A</b>) Completed cDNA and deduced amino acid sequence of <span class="html-italic">Pe</span>MT. The red font is the start codon and the end codon. The red underline is a putative polyadenylation signal. The blue shadow is the characteristic sequence of invertebrate MTs. The yellow shadow is similar to the characteristic sequence of Mollusca MTs. * represents the termination codon. (<b>B</b>) N-terminus and C-terminus of <span class="html-italic">Pe</span>MT. (<b>C</b>) Yellow parts show the conserved Cys residues, and gray spheres indicate Zn<sup>2+</sup>. (<b>D</b>) Multiple sequence alignment of <span class="html-italic">Pe</span>MT. The similarities between <span class="html-italic">Pe</span>MT and its homologs in <span class="html-italic">A. plicatula</span>, <span class="html-italic">C. gigas</span>, <span class="html-italic">S. cumingii</span>, <span class="html-italic">C. edule</span>, <span class="html-italic">P. viridis</span>, <span class="html-italic">D. rerio</span>, <span class="html-italic">O. latipes</span>, <span class="html-italic">M. gallus</span>, and <span class="html-italic">H. sapiens</span> were 37.1%, 36.8%, 33.3%, 35.2%, 32.4%, 25.7%, 20.0%, 17.1%, and 16.2%, respectively. The same amino acid residues are shaded in yellow, blue regions indicate amino acid residues with a similarity greater than 50%, and green regions represent lower similarity. (<b>E</b>) Phylogenetic analysis of MT homologous proteins. <span class="html-italic">P. esculenta</span> is shown in bold font, and <span class="html-italic">Pe</span>MT belongs to the invertebrate branch.</p>
Full article ">Figure 2
<p>Full-length cloning and bioinformatic analysis of <span class="html-italic">Pe</span>MTF1. (<b>A</b>) Completed cDNA and deduced amino acid sequence of <span class="html-italic">Pe</span>MTF1. The red shadow is the start codon and the end codon. The yellow shadows are the C<sub>2</sub>H<sub>2</sub>-type zinc finger domains. * represents the termination codon. (<b>B</b>) N-terminus and C-terminus of the protein sequence. (<b>C</b>) Yellow parts are the conserved C<sub>2</sub>H<sub>2</sub>-type zinc finger domains. (<b>D</b>) Multiple sequence alignment of <span class="html-italic">Pe</span>MTF1. The similarities between <span class="html-italic">Pe</span>MTF1 and its homologs in <span class="html-italic">C. pelagica</span>, <span class="html-italic">G. gallus</span>, <span class="html-italic">H. sapiens</span>, <span class="html-italic">M. musculus</span>, <span class="html-italic">R. norvegicus</span>, <span class="html-italic">C. tigris</span>, <span class="html-italic">L. agilis</span>, and <span class="html-italic">D. rerio</span> were 36.4%, 37.0%, 34.4%, 37.0%, 38.7%, 39.9%, 36.5%, and 43.0%, respectively, and the similarities of the zinc finger domains were 84.0%, 84.0%, 84.6%, 84.6%, 84.6%, 84.6%, 84.6%, and 85.7%, respectively. The red boxes are six conserved C<sub>2</sub>H<sub>2</sub>-type zinc finger domains. The same amino acid residues are shaded in yellow, blue regions indicate amino acid residues with a similarity greater than 50%, and green regions represent lower similarity. (<b>E</b>) Phylogenetic analysis of MTF1 homologous proteins. <span class="html-italic">P. esculenta</span> is shown in bold font, and <span class="html-italic">Pe</span>MTF1 belongs to the invertebrate branch.</p>
Full article ">Figure 3
<p>Tissue-specific expression of <span class="html-italic">PeMT</span> and <span class="html-italic">PeMTF1</span> mRNA. (<b>A</b>) RT-PCR detection of <span class="html-italic">PeMT</span> mRNA. (<b>B</b>) Gray value of RT-PCR analyzed using Image J. (<b>C</b>) RT-PCR detection of <span class="html-italic">PeMTF1</span> mRNA. (<b>D</b>) Gray value of RT-PCR analyzed using Image J. <span class="html-italic">β-actin</span> gene was used as an internal reference, and all data are presented as the mean ± SD (<span class="html-italic">n</span> = 3). Different letters indicate significant differences among the tissues (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Expression changes of (<b>A</b>) <span class="html-italic">PeMT</span> and (<b>B</b>) <span class="html-italic">PeMTF1</span> mRNA in the intestine of <span class="html-italic">P. esculenta</span> under Zn stress. <span class="html-italic">β-actin</span> gene was used as an internal reference, and all data are presented as the mean ± SD (<span class="html-italic">n</span> = 3). Different letters indicate significant differences among the groups (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Expression and purification of the recombinant pGEX-6p-1-MT. Line M: marker; Line 1 shows the cell lysate of pGEX-6p-1-MT without induction; Line 2 shows the cell lysate of pGEX-6p-1-MT induced by 1 mM IPTG (isopropyl-β-D-thiogalactoside); Line 3 shows the supernatant of cell lysate; Line 4 shows the precipitation of cell lysate; Lines 5–7 show the purified protein; black arrow: the expressed and purified protein.</p>
Full article ">Figure 6
<p>The growth status of the bacteria that transformed pGEX-6p-1-MT and pGEX-6p-1: (<b>A</b>) the control group; (<b>B</b>) the 0.3 mM Zn stress group. All data are presented as the mean ± SD (<span class="html-italic">n</span> = 3). “*”: <span class="html-italic">p</span> &lt; 0.05; “**”: <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 7
<p>The ABTS free radical scavenging ability of the recombinant <span class="html-italic">Pe</span>MT. Different concentrations of pGEX-6p-1 and GSH were set as the control groups. All data are presented as the mean ± SD (<span class="html-italic">n</span> = 3); different letters indicate significant differences among the groups (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 8
<p>Changes in the relative expression of <span class="html-italic">PeMT</span> mRNA after RNAi with <span class="html-italic">PeMT</span> and 24 h of Zn stress. <span class="html-italic">β-actin</span> gene was used as an internal reference, and all data are presented as the mean ± SD (<span class="html-italic">n</span> = 4), “*”: <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 9
<p>Changes in oxidative stress and apoptosis indices after RNAi with <span class="html-italic">PeMT</span> and 24 h of Zn stress. (<b>A</b>) MDA content, (<b>B</b>) SOD activity, (<b>C</b>) GSH content, (<b>D</b>) Caspase 3 activity, (<b>E</b>) Caspase 8 activity, and (<b>F</b>) Caspase 9 activity. All data are presented as the mean ± SD (<span class="html-italic">n</span> = 4), “*”: <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 10
<p>The expression changes of (<b>A</b>) <span class="html-italic">PeMTF1</span> and (<b>B</b>) <span class="html-italic">PeMT</span> mRNA in the intestine of <span class="html-italic">P. esculenta</span> after RNAi with <span class="html-italic">PeMTF1</span> and 24 h of Zn stress. <span class="html-italic">β-actin</span> gene was used as an internal reference, and all data are presented as the mean ± SD (<span class="html-italic">n</span> = 4), “*”: <span class="html-italic">p</span> &lt; 0.05, “**”: <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">
14 pages, 2569 KiB  
Article
Anti-Inflammatory Effect of Atorvastatin and Rosuvastatin on Monosodium Urate-Induced Inflammation through IL-37/Smad3-Complex Activation in an In Vitro Study Using THP-1 Macrophages
by Seong-Kyu Kim, Jung-Yoon Choe, Ji-Won Kim, Ki-Yeun Park and Boyoung Kim
Pharmaceuticals 2024, 17(7), 883; https://doi.org/10.3390/ph17070883 - 3 Jul 2024
Viewed by 300
Abstract
Objective: The pleiotropic effect of hydroxy-3-methylglutaryl coenzyme A (HMG-CoA) reductase inhibitors (statins) is responsible for potent defense against inflammatory response. This study evaluated the inhibitory effects of HMG-CoA reductase inhibitors on the monosodium urate (MSU)-induced inflammatory response through the regulation of interleukin-37 [...] Read more.
Objective: The pleiotropic effect of hydroxy-3-methylglutaryl coenzyme A (HMG-CoA) reductase inhibitors (statins) is responsible for potent defense against inflammatory response. This study evaluated the inhibitory effects of HMG-CoA reductase inhibitors on the monosodium urate (MSU)-induced inflammatory response through the regulation of interleukin-37 (IL-37) expression. Methods: Serum was collected from patients with gout (n = 40) and from healthy controls (n = 30). The mRNA and protein expression of the target molecules IL-1β, IL-37, caspase-1, and Smad3 were measured in THP-1 macrophages stimulated with MSU, atorvastatin, or rosuvastatin using a real-time quantitative polymerase chain reaction and Western blot assay. Transfection with IL-1β or Smad3 siRNA in THP-1 macrophages was used to verify the pharmaceutical effect of statins in uric-acid-induced inflammation. Results: Serum IL-37 levels in gout patients were significantly higher than in controls (p < 0.001) and was associated with the serum uric acid level (r = 0.382, p = 0.008). THP-1 cells stimulated with MSU markedly induced IL-37 mRNA expression and the transition of IL-37 from the cytoplasm to the nucleus. Recombinant IL-37 treatment dose-dependently inhibited activation of caspase-1 and IL-1β in MSU-induced inflammation. Atorvastatin and rosuvastatin attenuated caspase-1 activation and mature IL-1β expression but augmented translocation of IL-37 from the cytoplasm to the nucleus. Atorvastatin and rosuvastatin induced phosphorylation of Smad3 in THP-1 cells treated with MSU crystals. Statins potently attenuated translocation of IL-37 from the cytoplasm to the nucleus in THP-1 macrophages transfected with Smad3 siRNA compared to cells with negative control siRNA. Conclusions: This study revealed that statins inhibit the MSU-induced inflammatory response through phosphorylated Smad3-mediated IL-37 expression in THP-1 macrophages. Full article
(This article belongs to the Section Pharmacology)
Show Figures

Figure 1

Figure 1
<p>IL-37 expression in gout and uric-acid-induced inflammation. (<b>A</b>) Comparison of serum IL-37 level between gout patients (<span class="html-italic">n</span> = 40) and controls (<span class="html-italic">n</span> = 30). (<b>B</b>) Expression of IL-37 mRNA and IL-37 protein in cytoplasm and nucleus in THP-1 macrophages treated with MSU crystals. (<b>C</b>) Correlation of IL-37 level with uric acid, ESR, and CRP in serum in gout patients. (<b>D</b>) Expression of IL-37 mRNA and IL-37 protein in cytoplasm and nucleus in THP-1 macrophages transfected with IL-1β siRNA under stimulation with MSU crystals. (<b>E</b>) Expression of active caspase-1 and IL-1β in THP-1 macrophages stimulated with recombinant IL-37. * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01. Values presented as mean ± SEM of three independent experiments. The representative images are illustrated after three independent experiments. Abbreviation: MSU, monosodium urate; IL-37, interleukin-37; IL-1β, interleukin-1β; ESR, erythrocyte sedimentation rate; CRP, C-reactive protein.</p>
Full article ">Figure 2
<p>Changes in caspase-1, IL-1β, and IL-37 mRNA and protein expression by atorvastatin or rosuvastatin in THP-1 macrophages treated with MSU crystals. (<b>A</b>) Expression of IL-37 mRNA after addition of atorvastatin or rosuvastatin in THP-1 macrophages treated with MSU crystals. (<b>B</b>) Expression of cleaved caspase-1, cleaved IL-1β, and IL-37 protein in cytoplasm and nucleus after addition of atorvastatin or rosuvastatin in THP-1 macrophages treated with MSU crystals. * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01. Values presented as mean ± SEM of three independent experiments. The representative images are illustrated after three independent experiments. Abbreviation: MSU, monosodium urate; IL-37, interleukin-37; IL-1β, interleukin-1β.</p>
Full article ">Figure 3
<p>Changes in Smad3 and IL-37 expression by atorvastatin or rosuvastatin in MSU-induced inflammation. (<b>A</b>) Expression of phosphorylation of Smad3 after addition of atorvastatin or rosuvastatin in THP-1 macrophages treated with MSU crystals. (<b>B</b>) Comparison of IL-37 mRNA expression between cells transfected with NC siRNA and Smad3 siRNA under stimulation with MSU crystals. (<b>C</b>) Comparison of IL-37 protein expression in cytoplasm and nucleus between cells transfected with NC siRNA and Smad3 siRNA under stimulation with MSU crystals. * <span class="html-italic">p</span> &lt; 0.05. Values presented as mean ± SEM of three independent experiments. The representative images are illustrated after three independent experiments. Abbreviation: MSU, monosodium urate; IL-37, interleukin-37.</p>
Full article ">Figure 4
<p>Changes in proteolytic enzymes by atorvastatin or rosuvastatin in THP-1 macrophages. (<b>A</b>,<b>B</b>) Expression of elastase and cathepsin S mRNA and protein by atorvastatin or rosuvastatin in THP-1 macrophages treated with MSU crystals. Values presented as mean ± SEM of three independent experiments. The representative images are illustrated after three independent experiments. Abbreviation: MSU, monosodium urate.</p>
Full article ">Figure 5
<p>Proposed model for protective effect of statins through interaction with Smad3 and IL-37 in uric-acid-induced inflammation. MSU crystals induce activation of caspase-1 in NLRP3 inflammasome, leading to conversion of mature IL-37 from pro-IL-37. Statins block caspase-1 activation, but not other proteases in THP-1 macrophages treated with MSU crystals. In addition, statins activate phosphorylation of Smad3. Binding with activated IL-37 and Smad3 in cytoplasm is translocated into the nucleus, then stimulates transcription of anti-inflammatory molecules.</p>
Full article ">
22 pages, 4407 KiB  
Article
Recombinant Ixodes scapularis Calreticulin Binds Complement Proteins but Does Not Protect Borrelia burgdorferi from Complement Killing
by Moiz Ashraf Ansari, Thu-Thuy Nguyen, Klaudia Izabela Kocurek, William Tae Heung Kim, Tae Kwon Kim and Albert Mulenga
Pathogens 2024, 13(7), 560; https://doi.org/10.3390/pathogens13070560 - 3 Jul 2024
Viewed by 374
Abstract
Ixodes scapularis is a blood-feeding obligate ectoparasite responsible for transmitting the Lyme disease (LD) agent, Borrelia burgdorferi. During the feeding process, I. scapularis injects B. burgdorferi into the host along with its saliva, facilitating the transmission and colonization of the LD agent. [...] Read more.
Ixodes scapularis is a blood-feeding obligate ectoparasite responsible for transmitting the Lyme disease (LD) agent, Borrelia burgdorferi. During the feeding process, I. scapularis injects B. burgdorferi into the host along with its saliva, facilitating the transmission and colonization of the LD agent. Tick calreticulin (CRT) is one of the earliest tick saliva proteins identified and is currently utilized as a biomarker for tick bites. Our recent findings revealed elevated levels of CRT in the saliva proteome of B. burgdorferi-infected I. scapularis nymphs compared to uninfected ticks. Differential precipitation of proteins (DiffPOP) and LC-MS/MS analyses were used to identify the interactions between Ixs (I. scapularis) CRT and human plasma proteins and further explore its potential role in shielding B. burgdorferi from complement killing. We observed that although yeast-expressed recombinant (r) IxsCRT binds to the C1 complex (C1q, C1r, and C1s), the activator of complement via the classical cascade, it did not inhibit the deposition of the membrane attack complex (MAC) via the classical pathway. Intriguingly, rIxsCRT binds intermediate complement proteins (C3, C5, and C9) and reduces MAC deposition through the lectin pathway. Despite the inhibition of MAC deposition in the lectin pathway, rIxsCRT did not protect a serum-sensitive B. burgdorferi strain (B314/pBBE22Luc) from complement-induced killing. As B. burgdorferi establishes a local dermal infection before disseminating to secondary organs, it is noteworthy that rIxsCRT promotes the replication of B. burgdorferi in culture. We hypothesize that rIxsCRT may contribute to the transmission and/or host colonization of B. burgdorferi by acting as a decoy activator of complement and by fostering B. burgdorferi replication at the transmission site. Full article
(This article belongs to the Special Issue Ticks and Tick-Borne Pathogens: And Now What?)
Show Figures

Figure 1

Figure 1
<p><b>Expression and affinity purification of recombinant <span class="html-italic">I. scapularis</span> calreticulin.</b> Recombinant <span class="html-italic">I. scapularis</span> calreticulin (r<span class="html-italic">Ixs</span>CRT, 47.61 kDa) with His-tag was expressed in <span class="html-italic">Pichia pastoris</span> over 3 days and affinity-purified under native conditions. Affinity purification of r<span class="html-italic">Ixs</span>CRT was validated by standard SDS-PAGE followed by silver staining (<b>A</b>) and Western blotting using the antibody for the Histidine fusion tag (<b>B</b>). Lanes 1–4 represent the 50, 75, 100, and 200 imidazole concentrations (mM) at which the r<span class="html-italic">Ixs</span>CRT was eluted, dialyzed in PBS, concentrated, and used for the assays.</p>
Full article ">Figure 2
<p><b>Differential Precipitation of Proteins (DiffPOP) reveals interactions between r<span class="html-italic">Ixs</span>CRT and complement proteins.</b> Affinity-purified r<span class="html-italic">Ixs</span>CRT (10 µg) was pre-incubated with 10% normal human serum (NHS) for 90 min at 37 °C, while NHS and r<span class="html-italic">Ixs</span>CRT served as controls. The reactions were then stabilized and subjected to differential precipitation using escalating amounts of precipitating buffer, as described in the materials and methods section (see <a href="#app1-pathogens-13-00560" class="html-app">Supplementary Table S1</a>). Different fractions obtained from the precipitation process were analyzed by standard SDS-PAGE and silver staining. (<b>A</b>) shows the protein profile of NHS only, (<b>B</b>) illustrates the profile of the NHS + r<span class="html-italic">Ixs</span>CRT mixture, and (<b>C</b>) displays the profile of r<span class="html-italic">Ixs</span>CRT only. Additionally, the NHS + r<span class="html-italic">Ixs</span>CRT mixture was subjected to Western blotting analysis using an antibody against the histidine fusion tag to track the precipitation of r<span class="html-italic">Ixs</span>CRT (<b>D</b>). The distinct protein profile between NHS and NHS + r<span class="html-italic">Ixs</span>CRT, highlighted by the red arrowhead in panels (<b>A</b>,<b>B</b>), indicated potential interactions between r<span class="html-italic">Ixs</span>CRT and specific proteins. The precipitation profile depicted in (<b>D</b>) guided the selection of fractions for subsequent LCMS/MS analysis, shedding light on the specific complement proteins interacting with r<span class="html-italic">Ixs</span>CRT. L: ladder depicting molecular weight (kDa), Lanes 1–10 = DiffPOP fractions 1–10.</p>
Full article ">Figure 3
<p><b>Volcano plot analyses revealed human serum proteins that differentially co-precipitated with r<span class="html-italic">Ixs</span>CRT.</b> Normalized abundance values of three biological replicates were used in Proteome Discoverer™ 2.4 software (Thermo Fisher Scientific, Dallas, TX, USA) to generate the volcano plot. In the volcano plot, the Y-axis shows the −log10 <span class="html-italic">p</span>-value and the X-axis shows the magnitude of change (log2 fold change). Red dots represent proteins that co-precipitated in high amounts with r<span class="html-italic">Ixs</span>CRT, while green dots represent those that were absent or co-precipitated in low amounts in the presence of r<span class="html-italic">Ixs</span>CRT. An adjusted <span class="html-italic">p</span>-value ≤ 0.05 and log2 fold change of more than 2 were used as cut-offs to select proteins that co-precipitated in high amounts with r<span class="html-italic">Ixs</span>CRT.</p>
Full article ">Figure 4
<p><b>Differential precipitation of complement proteins in the presence of r<span class="html-italic">Ixs</span>CRT.</b> Fractions obtained from the differential precipitation (DiffPOP) were pooled into group A (fractions 1–5), group B (fraction 6), and group C (fractions 7 and 8). These fractions were subjected to Liquid Chromatography-Mass Spectrometry (LCMS/MS) analysis to investigate the abundance of complement proteins interacting with r<span class="html-italic">Ixs</span>CRT. (<b>A</b>) illustrates the abundance of complement proteins interacting with r<span class="html-italic">Ixs</span>CRT in group A, while (<b>B</b>) depicts the corresponding interactions in group B. The Y-axis in (<b>A</b>,<b>B</b>) illustrates the protein abundance, indicating the levels of proteins co-precipitated with r<span class="html-italic">Ixs</span>CRT (depicted by the red graph) in comparison to NHS only (represented by the blue graph).</p>
Full article ">Figure 5
<p><b>Validation of complement protein and r<span class="html-italic">Ixs</span>CRT interactions via Western blotting analysis.</b> To confirm the interactions observed in LCMS/MS analysis (referenced in <a href="#pathogens-13-00560-f003" class="html-fig">Figure 3</a>), DiffPOP fractions were subjected to standard Western blotting using antibodies specific for complement proteins: C1q, C1r, C1s, C3, C5, and C9, as indicated. Distinctive binding patterns, highlighting differences between NHS + r<span class="html-italic">Ixs</span>CRT and the NHS control, are denoted by red arrowheads. Panels (<b>A</b>,<b>C</b>,<b>E</b>,<b>G</b>,<b>I</b>,<b>K</b>) depict immunoblots of NHS-only samples, while panels (<b>B</b>,<b>D</b>,<b>F</b>,<b>H</b>,<b>J</b>,<b>L</b>) represent immunoblots of NHS + r<span class="html-italic">Ixs</span>CRT samples.</p>
Full article ">Figure 6
<p><b>Pull-down assay validated the r<span class="html-italic">Ixs</span>CRT binding of complement proteins.</b> Affinity-purified r<span class="html-italic">Ixs</span>CRT bound to His specific magnetic beads (Dynabeads™, Thermo Fisher Scientific, Waltham, MA, USA) was used to pull down complement proteins from human complement serum (HCS). The beads were washed and the eluted protein complexes were subjected to Western blotting analysis using antibodies for the histidine tag (<b>A</b>) and complement proteins C1q (<b>B</b>), C1r (<b>C</b>), C1s (<b>D</b>), C3 (<b>E</b>), C5 (<b>F</b>), and C9 (<b>G</b>). In all the panels (<b>A</b>–<b>G</b>), HCS is the human complement serum that was eluted from the empty beads, i.e., without the bait protein (r<span class="html-italic">Ixs</span>CRT), and HCS + r<span class="html-italic">Ixs</span>CRT denotes the human complement proteins that were pulled down and eluted from the beads loaded with r<span class="html-italic">Ixs</span>CRT. Red arrowheads denote the detected complement proteins at their expected molecular weight size (kDa).</p>
Full article ">Figure 7
<p><b>ELISA analysis demonstrates the r<span class="html-italic">Ixs</span>CRT binding of complement proteins.</b> High binding ELISA plates coated with affinity-purified r<span class="html-italic">Ixs</span>CRT (250 ng) were incubated with normal human serum (NHS) followed by antibodies for C1q (<b>A</b>), C1r (<b>B</b>), C1s (<b>C</b>), C3 (<b>D</b>), activated C3 (<b>E</b>), C5 (<b>F</b>), C9 (<b>G</b>), and C5b-9 or MAC (<b>H</b>) antibodies. Y-axis denotes the absorbance measured at <span class="html-italic">A</span><sub>450nm</sub>, which reflected the intensity of the specific antibody binding to r<span class="html-italic">Ixs</span>CRT. Non-coated wells blocked with 1% BSA were incubated with NHS and used as a negative control (Neg. Cont.). Data represent mean ± SEM of 3 biological replicates. For statistical analysis, <span class="html-italic">t</span>-test was performed on GraphPad Prism 9 and ** represents <span class="html-italic">p</span> &lt; 0.01, *** represents <span class="html-italic">p</span> &lt; 0.001, **** represents <span class="html-italic">p</span> &lt; 0.0001, and ns represents not significant.</p>
Full article ">Figure 8
<p>r<span class="html-italic">Ixs</span>CRT apparently enhances membrane attack complex (MAC) deposition in the classical pathway and alternate pathway but inhibits MAC deposition in the lectin pathway. The Wieslab<sup>®</sup> Complement System Kit (Svar Life Science AB, Malmo, Sweden) was used to detect the effects of r<span class="html-italic">Ixs</span>CRT on MAC deposition via the classical pathway, alternative pathway, and lectin pathway, as described in the materials and methods section. In brief, r<span class="html-italic">Ixs</span>CRT (4 μM) was incubated with NHS (provided with the kit) at 37 °C for 30 min and then added to wells pre-coated with the antibody for MAC. Diluent and kit-provided reagent served as negative and positive controls, respectively. After washing, the conjugate and substrate were added according to the manufacturer’s instructions for each kit. NC denotes negative control and NHS denotes normal human serum, used as the positive control. % MAC deposition (Y-axis) was calculated as mentioned in the methodology section and the MAC deposition in the positive control was represented as 100% (denoted by the black dotted line). Data are presented as deposited MAC ± SEM calculated from 3 biological replicates. Statistical significance was determined by Student’s <span class="html-italic">t</span>-test in GraphPad Prism 9. * Represents <span class="html-italic">p</span> ≤ 0.05, ** represents <span class="html-italic">p</span> ≤ 0.01, and ns represents not significant.</p>
Full article ">Figure 9
<p><b>r<span class="html-italic">Ixs</span>CRT does not protect <span class="html-italic">B. burgdorferi</span> from complement killing.</b> Normal human serum (NHS) was pre-incubated with serial dilutions of r<span class="html-italic">Ixs</span>CRT (1, 2, and 4 µM) or phosphate-buffered saline (PBS) at 37 °C for 30 min prior to the addition of 85 µL of 10<sup>6</sup> cells/mL of <span class="html-italic">B. burgdorferi</span> B314/pBBE22luc (complement-sensitive strain) and incubated in a bio-shaker at 32 °C and 100 rpm. NHS incubated with <span class="html-italic">B. burgdorferi</span> B314/pPCD100 (complement-resistant strain) was used as a positive control. Survival rates of <span class="html-italic">B. burgdorferi</span> were assessed at 3 h post-incubation. Data represent mean ± SEM of 3 biological replicates. Statistical significance was evaluated using <span class="html-italic">t</span>-test in GraphPad Prism 9 (ns: no significance, * represents <span class="html-italic">p</span>-value ≤ 0.05).</p>
Full article ">Figure 10
<p><b>r<span class="html-italic">Ixs</span>CRT promotes the growth of <span class="html-italic">B. burgdorferi</span> in culture.</b> <span class="html-italic">B. burgdorferi</span> (strain MSK5) cultured in the presence of r<span class="html-italic">Ixs</span>CRT at 2.6 mM (r<span class="html-italic">Ixs</span>CRT 125) and 5.2 mM (r<span class="html-italic">Ixs</span>CRT 250) were sampled at days 2, 5, and 7. (<b>A</b>) In triplicate, cells were quantified by manual counts on a Petroff-Hausser chamber and quantified using the following formula: number of cells/mL = Average of cells counted in all chambers × Dilution factor x 50,000. (<b>B</b>) <span class="html-italic">B. burgdorferi</span> was quantified by qPCR using the genomic DNA of the <span class="html-italic">B. burgdorferi</span> and <span class="html-italic">fLa</span>B primers. For ELISA, Student’s <span class="html-italic">t</span>-test was used for statistical analysis on GraphPad Prism 9 and <span class="html-italic">p</span>-value ≤ 0.05 (denoted by *) was considered significant for 3 biological replicates.</p>
Full article ">Figure 11
<p>One-time exposure of rabbits to <span class="html-italic">B. burgdorferi</span>-infected <span class="html-italic">I. scapularis</span> nymphs triggers high IgG antibody levels for r<span class="html-italic">Ixs</span>CRT. (<b>A</b>) Affinity-purified r<span class="html-italic">Ixs</span>CRT (250 ng) was subjected to standard ELISA using serially diluted purified IgG of rabbits that were fed upon for a single time by uninfected (blue line graph) and <span class="html-italic">B. burgdorferi</span>-infected (red line graph) <span class="html-italic">I. scapularis</span> nymph ticks. The black line graph represents pre-immune IgG binding, which was used for a negative control. (<b>B</b>–<b>D</b>) Various quantities of affinity-purified r<span class="html-italic">Ixs</span>CRT (100, 300, and 500 ng) were subjected to Western blotting analysis using purified IgG from rabbits that were fed upon for a single time by uninfected (<b>B</b>) and <span class="html-italic">B. burgdorferi</span>-infected ticks (<b>C</b>) as well as the antibody for the histidine tag, which served as a positive control (<b>D</b>).</p>
Full article ">
15 pages, 5634 KiB  
Article
Homogeneous Ensemble Feature Selection for Mass Spectrometry Data Prediction in Cancer Studies
by Yulan Liang, Amin Gharipour, Erik Kelemen and Arpad Kelemen
Mathematics 2024, 12(13), 2085; https://doi.org/10.3390/math12132085 - 3 Jul 2024
Viewed by 293
Abstract
The identification of important proteins is critical for the medical diagnosis and prognosis of common diseases. Diverse sets of computational tools have been developed for omics data reduction and protein selection. However, standard statistical models with single-feature selection involve the multi-testing burden of [...] Read more.
The identification of important proteins is critical for the medical diagnosis and prognosis of common diseases. Diverse sets of computational tools have been developed for omics data reduction and protein selection. However, standard statistical models with single-feature selection involve the multi-testing burden of low power with limited available samples. Furthermore, high correlations among proteins with high redundancy and moderate effects often lead to unstable selections and cause reproducibility issues. Ensemble feature selection in machine learning (ML) may identify a stable set of disease biomarkers that could improve the prediction performance of subsequent classification models and thereby simplify their interpretability. In this study, we developed a three-stage homogeneous ensemble feature selection (HEFS) approach for both identifying proteins and improving prediction accuracy. This approach was implemented and applied to ovarian cancer proteogenomics datasets comprising (1) binary putative homologous recombination deficiency (HRD)- positive or -negative samples; (2) multiple mRNA classes (differentiated, proliferative, immunoreactive, mesenchymal, and unknown samples). We conducted and compared various ML methods with HEFS including random forest (RF), support vector machine (SVM), and neural network (NN) for predicting both binary and multiple-class outcomes. The results indicated that the prediction accuracies varied for both binary and multiple-class classifications using various ML approaches with the proposed HEFS method. RF and NN provided better prediction accuracies than simple Naive Bayes or logistic models. For binary outcomes, with a sample size of 122 and nine selected prediction proteins using our proposed three-stage HEFS approach, the best ensemble ML (Treebag) achieved 83% accuracy, 85% sensitivity, and 81% specificity. For multiple (five)-class outcomes, the proposed HEFS-selected proteins combined with Principal Component Analysis (PCA) in NN resulted in prediction accuracies for multiple-class classifications ranging from 75% to 96% for each of the five classes. Despite the different prediction accuracies of the various models, HEFS identified consistent sets of proteins linked to the binary and multiple-class outcomes. Full article
(This article belongs to the Special Issue Current Research in Biostatistics)
Show Figures

Figure 1

Figure 1
<p>Flowchart of the ensemble process: training the models, selecting the best ones, handling redundancy, forming an ensemble, and finally generating variable-importance scores.</p>
Full article ">Figure 2
<p>The three stages of HEFS include training the models, selecting important features, handling redundancy, forming an ensemble, and generating variable-importance scores.</p>
Full article ">Figure 3
<p>Machine learning solutions with ensemble feature selections for mass spectrometry data prediction.</p>
Full article ">Figure 4
<p><b>Left</b>: Missing pattern examinations indicate that between 15% and 33% of expression missing occurred. <b>Right</b>: Multivariate scatter analysis shows the high correlations among the measured proteins.</p>
Full article ">Figure 5
<p>Neural Network with MLP (three layers, 10 hidden nodes) using 20 selected proteins for HRD two-class prediction: blue and red represent HRD-positive and -negative samples, respectively. <b>Left</b>: Prediction accuracy with ROC curves from training (<b>top</b>) and testing (<b>bottom</b>) sets; <b>right</b>: lift plot from training (<b>top</b>) and testing (<b>bottom</b>) sets.</p>
Full article ">Figure 5 Cont.
<p>Neural Network with MLP (three layers, 10 hidden nodes) using 20 selected proteins for HRD two-class prediction: blue and red represent HRD-positive and -negative samples, respectively. <b>Left</b>: Prediction accuracy with ROC curves from training (<b>top</b>) and testing (<b>bottom</b>) sets; <b>right</b>: lift plot from training (<b>top</b>) and testing (<b>bottom</b>) sets.</p>
Full article ">Figure 6
<p>ROC curve with the selected important proteins from HEFS for binary HRD class predictions with PCA: blue represents HRD-positive ovarian cancer, and red represents HRD-negative ovarian cancer.</p>
Full article ">Figure 7
<p><b>Top</b>: Neural Network with MLP (three layers, 10 hidden nodes) and HEFS-selected important proteins with combined principal components for mRNA multiple-class prediction. <b>Bottom left</b>: prediction accuracies with ROC curves and AUC for five classes from testing set; <b>bottom right</b>: lift plot from testing set.</p>
Full article ">
18 pages, 2735 KiB  
Article
Stable Production of a Recombinant Single-Chain Eel Follicle-Stimulating Hormone Analog in CHO DG44 Cells
by Munkhzaya Byambaragchaa, Sei Hyen Park, Sang-Gwon Kim, Min Gyu Shin, Shin-Kwon Kim, Myung-Hum Park, Myung-Hwa Kang and Kwan-Sik Min
Int. J. Mol. Sci. 2024, 25(13), 7282; https://doi.org/10.3390/ijms25137282 - 2 Jul 2024
Viewed by 392
Abstract
This study aimed to produce single-chain recombinant Anguillid eel follicle-stimulating hormone (rec-eel FSH) analogs with high activity in Cricetulus griseus ovary DG44 (CHO DG44) cells. We recently reported that an O-linked glycosylated carboxyl-terminal peptide (CTP) of the equine chorionic gonadotropin (eCG) β-subunit contributes [...] Read more.
This study aimed to produce single-chain recombinant Anguillid eel follicle-stimulating hormone (rec-eel FSH) analogs with high activity in Cricetulus griseus ovary DG44 (CHO DG44) cells. We recently reported that an O-linked glycosylated carboxyl-terminal peptide (CTP) of the equine chorionic gonadotropin (eCG) β-subunit contributes to high activity and time-dependent secretion in mammalian cells. We constructed a mutant (FSH-M), in which a linker including the eCG β-subunit CTP region (amino acids 115–149) was inserted between the β-subunit and α-subunit of wild-type single-chain eel FSH (FSH-wt). Plasmids containing eel FSH-wt and eel FSH-M were transfected into CHO DG44 cells, and single cells expressing each protein were isolated from 10 and 7 clones. Secretion increased gradually during the cultivation period and peaked at 4000–5000 ng/mL on day 9. The molecular weight of eel FSH-wt was 34–40 kDa, whereas that of eel FSH-M increased substantially, with two bands at 39–46 kDa. Treatment with PNGase F to remove the N glycosylation sites decreased the molecular weight remarkably to approximately 8 kDa. The EC50 value and maximal responsiveness of eel FSH-M were approximately 1.23- and 1.06-fold higher than those of eel FSH-wt, indicating that the mutant showed slightly higher biological activity. Phosphorylated extracellular-regulated kinase (pERK1/2) activation exhibited a sharp peak at 5 min, followed by a rapid decline. These findings indicate that the new rec-eel FSH molecule with the eCG β-subunit CTP linker shows potent activity and could be produced in massive quantities using the stable CHO DG44 cell system. Full article
(This article belongs to the Special Issue New Sights into Bioinformatics of Gene Regulations and Structure)
Show Figures

Figure 1

Figure 1
<p>Shape of the colony before isolation from the 96-well plate. After 10 d of incubation, the colonies were visually examined under a microscope for monoclonal colony growth. Images of representative colonies were obtained from eel FSH-wt samples approximately 3 weeks post-plating with complete cloning medium in a non-shaking incubator. Colonies were selected and transferred into 24-well plates. Next, they were transferred into 6-well, T-25 flasks and 125 mL shaker flasks. The scale bars represent 50 μm.</p>
Full article ">Figure 2
<p>Western blotting analysis of rec-eel FSH-wt proteins produced from single cells. Supernatants from 10 colonies were collected on the day of cultivation in a shaking incubator. The supernatants were resolved by sodium dodecyl sulfate-polyacrylamide gel electrophoresis and blotted onto membranes. Proteins were detected using a monoclonal antibody (anti-eel FSH5A14) and horseradish peroxidase-conjugated goat anti-mouse IgG antibodies. (<b>A</b>) In total, 20 µL of the supernatant on day 9 was loaded in the wells. Numbers denote isolated clone counts. (<b>B</b>) Colonies with substantial secretion were selected for Western blot analyses. We choose four colonies (eel FSH-wt 3, FSH-wt 5, FSH-wt 8, and FSH-wt 9) and 20 µL of supernatant was evaluated by Western blotting on the day of culture. FSH, follicle-stimulating hormone.</p>
Full article ">Figure 3
<p>Concentrations of rec-eel FSH-wt proteins secreted from CHO-DG44 cells on the day of culture. The supernatant was collected on days 0, 1, 3, 5, 9, and 11 of culture. The expression levels of rec-eel FSH-wt protein in monoclonal cells were analyzed using a sandwich enzyme-linked immunosorbent assay. Values are expressed as the mean ± standard error of mean from at least three independent experiments. FSH, follicle-stimulating hormone.</p>
Full article ">Figure 4
<p>Western blotting analysis of rec-eel FSH-M proteins produced by monoclonal cells. Supernatants were collected from seven colonies on the day of cultivation. The samples were prepared for SDS-PAGE. Membranes were detected using specific monoclonal antibodies (anti-eel FSH5A14). (<b>A</b>) In total, 20 µL collected on day 9 was loaded in the wells. Positive controls produced from the CHO-S cells were concentrated by 20 times and 20 µg was loaded in the wells. Two specific bands were detected for all samples. (<b>B</b>) Colonies judged to secrete large amounts were selected for Western blot analyses. We choose two colonies (eel FSH-M 3 and FSH-M 8) and 20 µL of the supernatant was used for Western blotting on the day of culture. Faint bands were first detected on day 3 and band intensity increased gradually, reaching a maximum on day 9.</p>
Full article ">Figure 5
<p>Deglycosylation results for eel FSH-wt and FSH-M proteins. The proteins collected from eel FSH-wt 3, FSH-M 3, and FSH-M 8 were treated with peptide-N-glycanase F to remove <span class="html-italic">N</span>-linked oligosaccharides, followed by Western blotting. The molecular weights of rec-eel FSH-wt and FSH-M decreased significantly to approximately 8–10 kDa.</p>
Full article ">Figure 6
<p>Concentrations of rec-eel FSH-M proteins secreted from CHO-DG44 cells on the day of culture. The supernatants were collected on days 0, 1, 3, 5, 9, and 11 of culture. The expression levels of rec-eel FSH-M in monoclonal cells were analyzed using a sandwich enzyme-linked immunosorbent assay. Values are expressed as the mean ± standard error of mean from at least three independent experiments.</p>
Full article ">Figure 7
<p>Effects of rec-eel FSH-wt and FSH-M on cyclic adenine monophosphate (cAMP) production in cells expressing eel follicle-stimulating hormone receptor. CHO-K1 cells transiently transfected with eel FSHR were seeded in 384-well plates (10,000 cells per well) at 24 h post-transfection. Cells were incubated with rec-eel FSH-wt or FSH-M for 30 min at room temperature. cAMP production was detected using a homogeneous time-resolved fluorescence assay and results are represented as Delta F%. Each data point represents the mean ± standard error of mean from triplicate experiments. The mean values were fitted to an equation to obtain a single-phase exponential decay curve.</p>
Full article ">Figure 8
<p>Time course for pERK1/2 activation in rec-eel FSH-wt and FSH-M. HEK293 cells were transiently transfected with eel FSH receptor, stimulated with 400 ng/mL agonist and normalized to the basal response. (<b>A</b>) Ratios are shown as delta F%. (<b>B</b>) The folds change values are shown, with 0 time set to 1-fold.</p>
Full article ">Figure 9
<p>pERK1/2 activation stimulated by eel FSH receptor. The eel FSH receptor was transiently transfected into HEK293 cells, and the cells were starved for 4–6 h and stimulated with a 400 ng/mL agonist for the indicated times. Whole-cell lysates were analyzed for pERK1/2 and total ERK levels. Twenty micrograms of protein were used in each sample lane. (<b>A</b>) Phosphorylation of ERK1/2 by western blotting. (<b>B</b>) The pERK and total ERK bands were quantified by densitometry, and pERK was normalized to total ERK levels. Representative data are shown, and graphs represent the mean and SE values from three independent experiments. No significant differences were observed between the curves representing eel FSH-wt- and FSH-M-treated samples.</p>
Full article ">Figure 10
<p>Schematic diagram of rec-eel FSH-wt and eel FSH-M. The tethered form of eel FSH β/α-wt containing the β-subunit and common α-subunit sequences was engineered. In the eel FSH-M mutant, the eCG β-subunit carboxyl-terminal peptide linker was inserted between the β-subunit and α-subunit using polymerase chain reaction. The eCG β-subunit CTP linker contained 35 amino acids sequence corresponding to the carboxyl-terminal peptide of the eCG β-subunit with approximately 12 <span class="html-italic">O</span>-linked oligosaccharide sites. The numbers in FSH-wt and FSH-M indicate the amino acids of the mature protein, except for the signal sequences. “N” denotes <span class="html-italic">N</span>-linked glycosylation sites at the eel FSH β-subunit and FSH α-subunit. Yellow indicates FSH β, the α-subunit is shown in blue, and the light blue shows the eCG CTP linker. eCTPβ (115–149) is the amino acid sequences of the eCG β-subunit CTP linker. Red in eCTPβ (115–149) denotes potential <span class="html-italic">O</span>-linked oligosaccharide sites.</p>
Full article ">
Back to TopTop