[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (839)

Search Parameters:
Keywords = transmembrane domain

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
16 pages, 2563 KiB  
Review
Research Progress into the Biological Functions of IFITM3
by Qian Xie, Liangliang Wang, Xinzhong Liao, Bi Huang, Chuming Luo, Guancheng Liao, Lifang Yuan, Xuejie Liu, Huanle Luo and Yuelong Shu
Viruses 2024, 16(10), 1543; https://doi.org/10.3390/v16101543 - 29 Sep 2024
Viewed by 364
Abstract
Interferon-induced transmembrane proteins (IFITMs) are upregulated by interferons. They are not only highly conserved in evolution but also structurally consistent and have almost identical structural domains and functional domains. They are all transmembrane proteins and have multiple heritable variations in genes. The IFITM [...] Read more.
Interferon-induced transmembrane proteins (IFITMs) are upregulated by interferons. They are not only highly conserved in evolution but also structurally consistent and have almost identical structural domains and functional domains. They are all transmembrane proteins and have multiple heritable variations in genes. The IFITM protein family is closely related to a variety of biological functions, including antiviral immunity, tumor formation, bone metabolism, cell adhesion, differentiation, and intracellular signal transduction. The progress of the research on its structure and related functions, as represented by IFITM3, is reviewed. Full article
Show Figures

Figure 1

Figure 1
<p>Alignment of the amino acid sequences of the IFITM3 protein with the red dashed lines for humans, chimpanzees, and mice. (IFITM3 are derived from avian, non-avian reptiles and amphibians). The shade of color represents the level of sequence consistency, with dark gray indicating complete sequence consistency.</p>
Full article ">Figure 2
<p>Molecular domains of the IFITM family proteins.</p>
Full article ">Figure 3
<p>Schematic representation of the possible topology of the IFITM family. (<b>A</b>) Schematic diagram of the IFITM family protein type III transmembrane protein topology. (<b>B</b>) Schematic representation of the intramembrane topology of IFITM family molecules. (<b>C</b>) Schematic representation of IFITM3 protein type II transmembrane. IMD, intramembrane domain. TMD, transmembrane domain.</p>
Full article ">Figure 4
<p>Possible antiviral mechanisms of IFITM3. (1) The amphipathic helical peptide (AH peptide) of IFITM3 may interact directly with cholesterol analogs to inhibit the formation of membrane fusion, thereby preventing viral entry. (2) IFITM3 may inhibit the fusion of virus and host cell membranes both by decreasing cell membrane fluidity and by stabilizing the cytoplasmic layer of the endosomal membrane to restrict viral entry from the intracellular compartment. (3) IFITM3 may interact with influenza virus haemagglutinin (HA) to reduce the optimal pH for membrane fusion, which in turn affects virus replication. (4) IFITM3 located in the lysosomal membrane may inhibit viral entry by disrupting transport processes in endosomes.</p>
Full article ">Figure 5
<p>Possible mechanisms of IFITMs immunomodulatory effects. The expression of IFITMs was upregulated in a variety of immune cells upon activation. Th1 cells enhanced the immune function of eosinophils, macrophages, and Th2 cells by upregulating Tbet, stat1 IL-27, etc. Th2 cells facilitated this process by downregulating Gata3, IL-4, and IL-13. In addition, IFITM3 with a 21-amino-acid deletion at the N-terminus on the surface of B cells promotes antibody production by plasma cells to enhance humoral immunity. “↑”, increase; “↓”, decrease.</p>
Full article ">Figure 6
<p>Prospects for future research directions in IFITM.</p>
Full article ">
16 pages, 3571 KiB  
Article
Characterization and Expression Analysis of the C-Type Lectin Ladderlectin in Litopenaeus vannamei Post-WSSV Infection
by Qian Xue, Bingbing Yang, Kun Luo, Sheng Luan, Jie Kong, Qiang Fu, Jiawang Cao, Baolong Chen, Ping Dai, Qun Xing, Xupeng Li and Xianhong Meng
Biology 2024, 13(10), 758; https://doi.org/10.3390/biology13100758 - 24 Sep 2024
Viewed by 374
Abstract
C-type lectins are known for agglutination activity and play crucial roles in regulating the prophenoloxidase (proPO) activation system, enhancing phagocytosis and encapsulation, synthesizing antimicrobial peptides, and mediating antiviral immune responses. This work cloned a C-type lectin, ladderlectin (LvLL), from Litopenaeus vannamei [...] Read more.
C-type lectins are known for agglutination activity and play crucial roles in regulating the prophenoloxidase (proPO) activation system, enhancing phagocytosis and encapsulation, synthesizing antimicrobial peptides, and mediating antiviral immune responses. This work cloned a C-type lectin, ladderlectin (LvLL), from Litopenaeus vannamei. LvLL comprised a 531 bp open reading frame (ORF) that encoded 176 amino acids. The predicted LvLL protein included a signal peptide and a CLECT domain. LvLL was predicted to feature a transmembrane region, suggesting it may be a transmembrane protein. LvLL was predominantly expressed in the shrimp’s hepatopancreas. After WSSV infection, LvLL expression in the hepatopancreas increased significantly by 11.35-fold after 228 h, indicating a general upregulation. Knockdown of LvLL resulted in a significant decrease in WSSV viral load and a notable increase in shrimp survival rates. Additionally, knockdown of LvLL led to a significant downregulation of apoptosis-related genes Bcl-2 and caspase 8 and a significant upregulation of p53 and proPO in WSSV-infected shrimp. This study showed that LvLL played a vital role in the interaction between L. vannamei and WSSV. Full article
(This article belongs to the Section Biochemistry and Molecular Biology)
Show Figures

Figure 1

Figure 1
<p>The cDNA sequence and deduced amino acid sequence of the <span class="html-italic">LvLL</span> gene. The start and stop codons are indicated by boxes. The predicted CLECT structural domain is underlined. The predicted transmembrane regions are indicated by parentheses. The predicted phosphorylation sites are indicated in bold font. SNP sites are marked in red font.</p>
Full article ">Figure 2
<p>Multiple sequence alignment of ladderlectin amino acid sequences. The 100% identical residues are indicated by black shading, 75% identical residues are indicated by dark gray shading, and 50% identical residues are indicated by light gray shading. The GenBank accession numbers of ladderlectin amino acid sequences are as follows: <span class="html-italic">L</span>. <span class="html-italic">vannamei</span> (XP_027219993.1), <span class="html-italic">S</span>. <span class="html-italic">salar</span> (XP_045578907.1), <span class="html-italic">S</span>. <span class="html-italic">trutta</span> (XP_029622341.1), <span class="html-italic">O</span>. <span class="html-italic">kisutch</span> (XP_031670965.1), <span class="html-italic">S</span>. <span class="html-italic">pilchardus</span> (XP_062407595.1), <span class="html-italic">P</span>. <span class="html-italic">clarkii</span> (XP_045588190.1), <span class="html-italic">Clupea harengus</span> (XP_042560807.1), <span class="html-italic">P</span>. <span class="html-italic">japonicus</span> (XP_042881645.1), <span class="html-italic">P</span>. <span class="html-italic">monodon</span> (XP_037799159.1), <span class="html-italic">Pungitius pungitius</span> (XP_037330723.2), <span class="html-italic">P</span>. <span class="html-italic">chinensis</span> (XP_047499869.1), <span class="html-italic">Poecilia reticulata</span> (XP_008419991.1), <span class="html-italic">R. philippinarum</span> (XP_060552084.1), <span class="html-italic">D</span>. <span class="html-italic">rerio</span> (XP_001337601.1), <span class="html-italic">P</span>. <span class="html-italic">trituberculatus</span> (XP_045115152.1).</p>
Full article ">Figure 3
<p>Phylogenetic tree analysis of ladderlectin. The LvLL marker of <span class="html-italic">L</span>. <span class="html-italic">vannamei</span> is “▲”.</p>
Full article ">Figure 4
<p>Expression profiles of <span class="html-italic">LvLL</span> in hepatopancreas, gill, and muscle of healthy <span class="html-italic">L</span>. <span class="html-italic">vannamei</span>. (**: <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 5
<p>Expression profiles of <span class="html-italic">LvLL</span> in hepatopancreas, gill, and muscle of <span class="html-italic">L</span>. <span class="html-italic">vannamei</span> after WSSV infection. (<b>A</b>): Hepatopancreas, (<b>B</b>): gill, (<b>C</b>): muscle (*: <span class="html-italic">p</span> &lt; 0.05, **: <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 6
<p>WSSV infection was suppressed after knocking down <span class="html-italic">LvLL</span>. (<b>A</b>): The knockdown efficiency of <span class="html-italic">LvLL</span> in the hepatopancreas at 24 and 48 h after WSSV infection. The knockdown efficiencies were 94.00% and 91.00% at 24 and 48 h after WSSV infection, respectively. (<b>B</b>): The WSSV viral load in the <span class="html-italic">dsGFP</span> + WSSV, WSSV, and <span class="html-italic">dsLvLL</span> + WSSV groups after <span class="html-italic">LvLL</span> knockdown. (<b>C</b>): The survival rate of <span class="html-italic">L</span>. <span class="html-italic">vannamei</span> after WSSV infection. *: <span class="html-italic">p</span> &lt; 0.05, **: <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 7
<p>Expression profiles of immune-related genes after <span class="html-italic">LvLL</span> knockdown. (<b>A</b>): The expression levels of <span class="html-italic">Bcl-2</span> at different post-WSSV infection time points after <span class="html-italic">LvLL</span> knockdown. (<b>B</b>): The expression levels of caspase 8 at different post-WSSV infection time points after <span class="html-italic">LvLL</span> knockdown. (<b>C</b>): The expression levels of <span class="html-italic">p53</span> at different post-WSSV infection time points after <span class="html-italic">LvLL</span> knockdown. (<b>D</b>): The expression levels of <span class="html-italic">proPO</span> at different post-WSSV infection time points after <span class="html-italic">LvLL</span> knockdown. *: <span class="html-italic">p</span> &lt; 0.05, **: <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">
19 pages, 1779 KiB  
Review
A Deep Dive into the N-Terminus of STIM Proteins: Structure–Function Analysis and Evolutionary Significance of the Functional Domains
by Sasirekha Narayanasamy, Hwei Ling Ong and Indu S. Ambudkar
Biomolecules 2024, 14(10), 1200; https://doi.org/10.3390/biom14101200 - 24 Sep 2024
Viewed by 597
Abstract
Calcium is an important second messenger that is involved in almost all cellular processes. Disruptions in the regulation of intracellular Ca2+ levels ([Ca2+]i) adversely impact normal physiological function and can contribute to various diseased conditions. STIM and Orai [...] Read more.
Calcium is an important second messenger that is involved in almost all cellular processes. Disruptions in the regulation of intracellular Ca2+ levels ([Ca2+]i) adversely impact normal physiological function and can contribute to various diseased conditions. STIM and Orai proteins play important roles in maintaining [Ca2+]i through store-operated Ca2+ entry (SOCE), with STIM being the primary regulatory protein that governs the function of Orai channels. STIM1 and STIM2 are single-pass ER-transmembrane proteins with their N- and C-termini located in the ER lumen and cytoplasm, respectively. The N-terminal EF-SAM domain of STIMs senses [Ca2+]ER changes, while the C-terminus mediates clustering in ER-PM junctions and gating of Orai1. ER-Ca2+ store depletion triggers activation of the STIM proteins, which involves their multimerization and clustering in ER-PM junctions, where they recruit and activate Orai1 channels. In this review, we will discuss the structure, organization, and function of EF-hand motifs and the SAM domain of STIM proteins in relation to those of other eukaryotic proteins. Full article
(This article belongs to the Section Biomacromolecules: Proteins)
Show Figures

Figure 1

Figure 1
<p>Schematic representation of the activation of store-operated Ca<sup>2+</sup> entry (SOCE). The cascade of molecular events leading to the ER-Ca<sup>2+</sup> store depletion and the re-filling up of stores are shown here. STIM proteins are single-pass ER-TM proteins with their N-terminal domains (dashed boxes) within the ER lumen. The NMR structures for the N-terminal EF-SAM of STIM1 (PDB: 2K60 [<a href="#B27-biomolecules-14-01200" class="html-bibr">27</a>]) and STIM2 (PDB: 2L5Y [<a href="#B28-biomolecules-14-01200" class="html-bibr">28</a>]) are shown in dashed boxes. The N-terminus contains a canonical EF-hand (cEF; magenta), a non-canonical EF-hand (nEF; yellow), and a SAM domain (blue). GPCR, G-protein coupled receptor; PLC-ß, phospholipase C-ß; PIP<sub>2</sub>, phosphatidylinositol 4,5-bisphosphate; DAG, diacylglycerol; IP<sub>3</sub>, inositol 1,4,5-trisphosphate; IP<sub>3</sub>R, IP<sub>3</sub> receptor; ER, endoplasmic reticulum; SERCA, Sarco-ER Ca<sup>2+</sup>; TG, thapsigargin.</p>
Full article ">Figure 2
<p>Multiple sequence alignment of STIM1 homologs. Multiple sequence alignment of STIM1 was generated in the MAFFT alignment tool employing BLAST search using human STIM1 (Uniprotkb: Q13586) as a query sequence. The amino acids in the cEF loop and nEF are highly conserved among STIM1 orthologs, shown as red and olive green, respectively. Any variations in the consensus sequence within the loop regions are marked by different colors (dark blue, positively charged amino acids; light green, negatively charged amino acids; light blue, polar uncharged amino acids; turquoise blue, hydrophobic amino acids). Protein sequences of STIM2 (human), STIM from Worm (<span class="html-italic">C. elegans</span>), and Fruit fly (<span class="html-italic">D. melanogaster</span>) were also included in the alignment, shown as #, *, and <span>$</span>, respectively.</p>
Full article ">Figure 3
<p>Gain of function (GoF) mutations leading to TAM/York platelet syndrome in canonical (panel (<b>a</b>)) and non-canonical (panel (<b>b</b>)) EF-hands of STIM1 (PDB: 2K60 [<a href="#B27-biomolecules-14-01200" class="html-bibr">27</a>]). TAM mutations are indicated by either dashed/solid black boxes; the dashed black box around the mutants indicates mutations in the entering/exiting helix. A solid black box indicates mutations in the loop region. The numbers within the paratheses of the mutants represent the position of these amino acids in a typical EF-hand motif; the I115F mutant, seen in both TAM and York platelet syndrome, is shown as an orange box.</p>
Full article ">
13 pages, 2192 KiB  
Article
The Role of the N-Terminal Domain of Thrombomodulin and the Potential of Recombinant Human Thrombomodulin as a Therapeutic Intervention for Shiga Toxin-Induced Hemolytic-Uremic Syndrome
by Sarah Kröller, Jana Schober, Nadine Krieg, Sophie Dennhardt, Wiebke Pirschel, Michael Kiehntopf, Edward M. Conway and Sina M. Coldewey
Toxins 2024, 16(9), 409; https://doi.org/10.3390/toxins16090409 - 20 Sep 2024
Viewed by 662
Abstract
Hemolytic-uremic syndrome (HUS) is a rare complication of an infection with Shiga toxin (Stx)-producing Escherichia coli (STEC-HUS), characterized by severe acute kidney injury, thrombocytopenia and microangiopathic hemolytic anemia, and specific therapy is still lacking. Thrombomodulin (TM) is a multi-domain transmembrane endothelial cell protein [...] Read more.
Hemolytic-uremic syndrome (HUS) is a rare complication of an infection with Shiga toxin (Stx)-producing Escherichia coli (STEC-HUS), characterized by severe acute kidney injury, thrombocytopenia and microangiopathic hemolytic anemia, and specific therapy is still lacking. Thrombomodulin (TM) is a multi-domain transmembrane endothelial cell protein and its N-terminal domain has been implicated in the pathophysiology of some cases of HUS. Indeed, the administration of recombinant human TM (rhTM) may have efficacy in HUS. We used a Stx-based murine model of HUS to characterize the role of the N-terminal domain of TM. We show that mice lacking that domain (TMLed (−/−)) are more sensitive to Stx, with enhanced HUS progression seen at 4 days and increased mortality at 7 days post-HUS induction. In spite of these changes, renal function was less affected in surviving Stx-challenged TMLed (−/−) mice compared to their wild-type counterparts TMLed (+/+) at 7 days. Contrary to few clinical case reports from Japan, the administration of rhTM (0.06 mg/kg) to wild-type mice (C57BL/6J) with HUS did not protect against disease progression. This overall promising, but also contradictory body of evidence, requires further systematic preclinical and clinical investigations to clarify the role of TM in HUS as a potential therapeutic strategy. Full article
Show Figures

Figure 1

Figure 1
<p>Disease progression of TMLed (+/+) and TMLed (−/−) mice with HUS at day 4 and day 7. Survival was followed up for (<b>A</b>) 4 days or (<b>B</b>) 7 days in sham and Stx-challenged mice at days 0, 3, and 6. Displayed are (<b>A</b>,<b>B</b>) survival; (<b>C</b>,<b>D</b>) analysis of HUS progression indicated by HUS score over the duration of experiment (ranging from 1 = no signs of illness to 5 = dead); (<b>E</b>,<b>F</b>) progression of weight loss over the duration of experiment; and (<b>G</b>,<b>H</b>) weight loss at the end of the experiment. (<b>A</b>,<b>C</b>): <span class="html-italic">n</span> = 7 for TMLed (−/−) sham; <span class="html-italic">n</span> = 8 for other groups; (<b>B</b>,<b>D</b>): <span class="html-italic">n</span> = 6 for TMLed (+/+) Stx and <span class="html-italic">n</span> = 7 for TMLed (−/−); (<b>E</b>–<b>H</b>): <span class="html-italic">n</span> = 3–8 mice per group (only surviving mice on day 4 or day 7). Data are presented with mean <span class="underline">+</span> SD. (<b>A</b>,<b>B</b>) Survival by Kaplan-Meier survival analysis + post hoc test. (<b>C</b>,<b>D</b>) Two-way-ANOVA with Tukey’s multiple comparisons test. (<b>G</b>) One-way ANOVA + Sidak’s multiple comparisons test and (<b>H</b>) Mann-Whitney test. * <span class="html-italic">p</span> &lt; 0.05 vs. corresponding sham group; # <span class="html-italic">p</span> &lt; 0.05 TMLed (+/+) Stx vs. TMLed (−/−) Stx. HUS, hemolytic-uremic syndrome; Stx, Shiga toxin; ns, not significant.</p>
Full article ">Figure 2
<p>Parameters of renal injury, hemolysis, and liver injury of TMLed (+/+) and TMLed (−/−) mice with HUS at day 4 and day 7. Determination of plasma (<b>A</b>) creatinine, (<b>B</b>) urea, (<b>C</b>) cholesterol, (<b>D</b>) LDH activity, (<b>E</b>) ALAT, (<b>F</b>) ASAT, and (<b>G</b>) albumin. (<b>A</b>–<b>C</b>,<b>E</b>) One-way ANOVA + Sidak’s multiple comparisons test. (<b>D</b>,<b>F</b>,<b>G</b>) Kruskal–Wallis test + Dunn’s multiple comparisons test. <span class="html-italic">n</span> = 3–8 per group (only surviving mice with blood withdrawal on day 4 or day 7). Data are presented with mean <span class="underline">+</span> SD. * <span class="html-italic">p</span> &lt; 0.05 vs. corresponding sham group. <span>$</span> <span class="html-italic">p</span> &lt; 0.05 Stx 4 days vs. Stx 7 days (same genotype). # <span class="html-italic">p</span> &lt; 0.05 TMLed (+/+) Stx vs. TMLed (−/−) Stx. Stx, Shiga toxin; LDH, lactate dehydrogenase; ALAT, alanine transaminase; ASAT, aspartate transaminase.</p>
Full article ">Figure 3
<p>Hematological parameters of TMLed (+/+) and TMLed (−/−) mice with HUS at day 4 and day 7. Determination of whole blood (<b>A</b>) RBC, (<b>B</b>) HGB, (<b>C</b>) HCT, (<b>D</b>) MCV, (<b>E</b>) MCH, (<b>F</b>) MCHC, (<b>G</b>) RDW-SD, (<b>H</b>) RDW-CV, (<b>I</b>) PLT, (<b>J</b>) WBC, (<b>K</b>) W-SCR, and (<b>L</b>) W-LCR. (<b>A</b>–<b>C</b>,<b>F</b>,<b>J</b>–<b>L</b>) One-way ANOVA + Sidak’s multiple comparisons test. (<b>D</b>,<b>E</b>,<b>G</b>–<b>I</b>) Kruskal–Wallis test + Dunn’s multiple comparisons test. <span class="html-italic">n</span> = 3−8 per group (only surviving mice with blood withdrawal on day 4 or day 7). Data are presented with mean <span class="underline">+</span> SD. * <span class="html-italic">p</span> &lt; 0.05 vs. corresponding sham group. <span>$</span> <span class="html-italic">p</span> &lt; 0.05 Stx 4 days vs. Stx 7 days (same genotype). # <span class="html-italic">p</span> &lt; 0.05 TMLed (+/+) Stx vs. TMLed (−/−) Stx. Stx, Shiga toxin; RBC, red blood cell; HGB, hemoglobin; HCT, hematocrit; MCV, mean corpuscular volume, MCH, mean corpuscular hemoglobin; MCHC, mean corpuscular hemoglobin concentration; RDW, red blood cell distribution width; PLT, platelet; WBC, white blood cell; W-SCR, white cell–-small cell ratio (including lymphocytes); W-LCR, white cell- large cell ratio (including neutrophils).</p>
Full article ">Figure 4
<p>Clinical presentation and parameters of kidney injury of mice with HUS treated with rhTM. HUS was followed up for 7 days in sham and Stx-challenged mice with i.v. injection of rhTM (0.06 mg/kg bodyweight) on day 0, 3, and 6. (<b>A</b>) Survival is displayed by Kaplan–Meier survival analysis + post hoc test. (<b>B</b>) HUS progression is indicated by HUS score (ranging from 1 = no signs of illness to 5 = dead) and (<b>C</b>) progression of weight loss over the duration of experiment. (<b>D</b>) Plasma NGAL was determined on humane endpoint or day 7. Quantification of (<b>E</b>) PAS reaction, (<b>F</b>) fibrin deposits and (<b>G</b>) relative CD31 expression in renal sections at the end of the experiment (day 7 or humane endpoint). (<b>B</b>) Two-way-ANOVA with Tukey’s multiple comparisons test. (<b>D</b>,<b>G</b>) One-way ANOVA + Holm–Sidak’s multiple comparison test. (<b>E</b>,<b>F</b>) Kruskal–Wallis test + Dunn’s multiple comparison test. <span class="html-italic">n</span> = 8 for sham + vehicle; <span class="html-italic">n</span> = 10 for Stx + vehicle and Stx + rhTM; <span class="html-italic">n</span> = 7 for sham + rhTM. Data are presented with mean <span class="underline">+</span> SD. * <span class="html-italic">p</span> &lt; 0.05 vs. corresponding sham group; # <span class="html-italic">p</span> &lt; 0.05 Stx + vehicle vs. Stx + rhTM. HUS, hemolytic–uremic syndrome; Stx, Shiga toxin; rhTM, recombinant human thrombomodulin; NGAL, neutrophil gelatinase-associated lipocalin; PAS, periodic acid Schiff; CD31, cluster of differentiation 31.</p>
Full article ">Figure 5
<p>Hematological parameters of mice with HUS treated with rhTM. Determination of whole blood (<b>A</b>) RBC, (<b>B</b>) HGB, (<b>C</b>) HCT, (<b>D</b>) MCV, (<b>E</b>) MCH, (<b>F</b>) MCHC, (<b>G</b>) WBC, (<b>H</b>) lymphocytes, and (<b>I</b>) granulocytes on humane endpoint or day 7. (<b>A</b>–<b>D</b>,<b>F</b>,<b>H</b>,<b>I</b>) Kruskal–Wallis test + Dunn’s multiple comparison test and (<b>E</b>,<b>G</b>) one-way ANOVA + Holm–Sidak’s multiple comparison test. <span class="html-italic">n</span> = 8 for sham + vehicle; <span class="html-italic">n</span> = 10 for Stx + vehicle and Stx + rhTM; <span class="html-italic">n</span> = 7 for sham + rhTM. Data are presented as mean <span class="underline">+</span> SD. * <span class="html-italic">p</span> &lt; 0.05. HUS, hemolytic–uremic syndrome; rhTM, recombinant human thrombomodulin; RBC, red blood cell; HGB, hemoglobin; HCT, hematocrit; MCV, mean corpuscular volume; MCH, mean corpuscular hemoglobin; MCHC, mean corpuscular hemoglobin concentration; WBC, white blood cells.</p>
Full article ">
20 pages, 906 KiB  
Article
Unveiling the Full Protein Effectorome of the Black Sigatoka Pathogen Pseudocercospora fijiensis—An In Silico Approach
by Karla Gisel Carreón-Anguiano, Jewel Nicole Anna Todd, César De los Santos-Briones, Santy Peraza-Echeverría, Ignacio Islas-Flores and Blondy Canto-Canché
Microbiol. Res. 2024, 15(3), 1880-1899; https://doi.org/10.3390/microbiolres15030126 - 14 Sep 2024
Viewed by 524
Abstract
Pseudocercospora (previously Mycosphaerella) fijiensis is a hemibiotroph fungus and the causal agent of black Sigatoka disease, one of the most significant threats to banana production worldwide. Only a few genomics reports have paid any attention to effector proteins, which are key players [...] Read more.
Pseudocercospora (previously Mycosphaerella) fijiensis is a hemibiotroph fungus and the causal agent of black Sigatoka disease, one of the most significant threats to banana production worldwide. Only a few genomics reports have paid any attention to effector proteins, which are key players in pathogenicity. These reports focus on canonical effectors: small secreted proteins, rich in cysteines, containing a signal peptide and no transmembrane domain. Thus, bias in previous reports has resulted in the non-canonical effectors being, in effect, excluded from the discussion of effectors in P. fijiensis pathogenicity. Here, using WideEffHunter and EffHunter, bioinformatic tools which identify non-canonical and canonical effectors, respectively, we predict, for the first time, the full effectorome of P. fijiensis. This complete effectorome comprises 5179 proteins: 240 canonical and 4939 non-canonical effectors. Protein families related to key functions of the hemibiotrophic lifestyle, such as Salicylate hydroxylase and Isochorismatase, are widely represented families of effectors in the P. fijiensis genome. An analysis of the gene distribution in core and dispensable scaffolds of both classes of effectors revealed a novel genomic structure of the effectorome. The majority of the effectors (canonical and non-canonical) were found to be harbored in the core scaffolds, while dispensable scaffolds harbored less than 10% of the effectors, all of which were non-canonical. Additionally, we found the motifs RXLR, YFWxC, LysM, EAR, [Li]xAR, PDI, CRN, and ToxA in the effectors of P. fijiensis. This novel genomic structure of effectors (more enriched in the core than in the dispensable genome), as well as the occurrence of effector motifs which were also observed in four other fungi, evidences that these phenomena are not unique to P. fijiensis; rather, they are widely occurring characteristics of effectors in other fungi. Full article
Show Figures

Figure 1

Figure 1
<p>Protein motifs found in the effectorome of <span class="html-italic">P. fijiensis</span>: (<b>A</b>) abundance of the protein motifs in the effectors of <span class="html-italic">P. fijiensis</span>. And (<b>B</b>) distribution of the <span class="html-italic">P. fijiensis</span> effector-motifs throughout the genomic scaffolds.</p>
Full article ">Figure 1 Cont.
<p>Protein motifs found in the effectorome of <span class="html-italic">P. fijiensis</span>: (<b>A</b>) abundance of the protein motifs in the effectors of <span class="html-italic">P. fijiensis</span>. And (<b>B</b>) distribution of the <span class="html-italic">P. fijiensis</span> effector-motifs throughout the genomic scaffolds.</p>
Full article ">
27 pages, 4900 KiB  
Review
Comprehensive Insights into the Molecular Basis of HIV Glycoproteins
by Amir Elalouf, Hanan Maoz and Amit Yaniv Rosenfeld
Appl. Sci. 2024, 14(18), 8271; https://doi.org/10.3390/app14188271 - 13 Sep 2024
Viewed by 730
Abstract
Human Immunodeficiency Virus (HIV) is a diploid, C-type enveloped retrovirus belonging to the Lentivirus genus, characterized by two positive-sense single-stranded RNA genomes, that transitioned from non-human primates to humans and has become globally widespread. In its advanced stages, HIV leads to Acquired Immune [...] Read more.
Human Immunodeficiency Virus (HIV) is a diploid, C-type enveloped retrovirus belonging to the Lentivirus genus, characterized by two positive-sense single-stranded RNA genomes, that transitioned from non-human primates to humans and has become globally widespread. In its advanced stages, HIV leads to Acquired Immune Deficiency Syndrome (AIDS), which severely weakens the immune system by depleting CD4+ helper T cells. Without treatment, HIV progressively impairs immune function, making the body susceptible to various opportunistic infections and complications, including cardiovascular, respiratory, and neurological issues, as well as secondary cancers. The envelope glycoprotein complex (Env), composed of gp120 and gp41 subunits derived from the precursor gp160, plays a central role in cycle entry. gp160, synthesized in the rough endoplasmic reticulum, undergoes glycosylation and proteolytic cleavage, forming a trimeric spike on the virion surface. These structural features, including the transmembrane domain (TMD), membrane-proximal external region (MPER), and cytoplasmic tail (CT), are critical for viral infectivity and immune evasion. Glycosylation and proteolytic processing, especially by furin, are essential for Env’s fusogenic activity and capacity to evade immune detection. The virus’s outer envelope glycoprotein, gp120, interacts with host cell CD4 receptors. This interaction, along with the involvement of coreceptors CXCR4 and CCR5, prompts the exposure of the gp41 fusogenic components, enabling the fusion of viral and host cell membranes. While this is the predominant pathway for viral entry, alternative mechanisms involving receptors such as C-type lectin and mannose receptors have been found. This review aims to provide an in-depth analysis of the structural features and functional roles of HIV entry proteins, particularly gp120 and gp41, in the viral entry process. By examining these proteins’ architecture, the review elucidates how their structural properties facilitate HIV invasion of host cells. It also explores the synthesis, trafficking, and structural characteristics of Env/gp160 proteins, highlighting the interactions between gp120, gp41, and the viral matrix. These contributions advance drug resistance management and vaccine development efforts. Full article
(This article belongs to the Section Applied Biosciences and Bioengineering)
Show Figures

Figure 1

Figure 1
<p>HIV-1 Env trafficking. A detailed mechanism. The HIV-1 envelope glycoprotein (Env) undergoes a complex trafficking pathway. Initially, Env is synthesized and glycosylated within the rough endoplasmic reticulum (RER) as a 160-kDa precursor protein, designated gp160. This precursor protein predominantly oligomerizes into trimers. Concurrently, the precursor Gag protein (Pr55 Gag) is synthesized on cytosolic ribosomes and directed to the plasma membrane (PM), where it multimerizes within lipid rafts (not depicted here) to form nascent virus particles. The oligomerized gp160 is subsequently transported to the Golgi apparatus and the trans-Golgi network (TGN). Here, gp160 undergoes proteolytic processing to generate the mature surface glycoprotein gp120 and the transmembrane glycoprotein gp41. The gp120/gp41 complexes proceed through the secretory pathway, eventually reaching the plasma membrane, where they are incorporated into virus particles as trimeric spikes. At the plasma membrane, Env can undergo endocytosis mediated by clathrin adaptor complexes, leading to its internalization into early endosomes (EE). Within the cell, internalized Env has two potential fates: it can be routed to late endosomes/multivesicular bodies (LE/MVBs) for subsequent degradation in lysosomes, or it can be recycled back to the plasma membrane via recycling endosomes. The domains of Gag and Env are detailed in the inset located at the top left of the diagram. Reprinted (adapted) with permission from [<a href="#B31-applsci-14-08271" class="html-bibr">31</a>,<a href="#B56-applsci-14-08271" class="html-bibr">56</a>] Copyright Elsevier (2024).</p>
Full article ">Figure 2
<p>HIV-1 gp160 protein structure and post-translational modifications. (<b>a</b>) Structure and processing of precursor gp160. The precursor gp160 protein of HIV-1 includes a signal peptide (SP) that is cleaved during translation. Post-translationally, gp160 is processed into the surface subunit (gp120) and the transmembrane subunit (gp41) within the Golgi complex at a specific furin cleavage site. The gp120 subunit comprises five variable domains (V1–V5) and five constant domains (C1–C5). In contrast, the gp41 subunit features extracellular constant domains (C6 and C7) that includes the fusion peptide (FP), heptad repeats (HR1 and HR2), the membrane-proximal external region (MPER), a transmembrane domain (TMD), and a cytoplasmic tail (CT). An enlarged depiction of the gp41 CT highlights several critical motifs: the internalization signal YSPL, the Kennedy sequence, the amphipathic α-helices LLP-1, LLP-2 and LLP-3, and a C-terminal dileucine motif (LL) implicated in the endocytosis and intracellular distribution of Env. The glycosylation sites on various HIV-1 gp120 variants are shown in a structure-based sequence alignment. These sites are called complex or high-mannose glycosylation sites. The domains are color-coded, with the N and C termini indicated. Disulfide bonds are represented as red lines, and the inner-domain β sandwich is boxed. Oligomannose and complex glycans are symbolized by three or two-pronged forked symbols, respectively. Reprinted (adapted) from [<a href="#B34-applsci-14-08271" class="html-bibr">34</a>,<a href="#B57-applsci-14-08271" class="html-bibr">57</a>,<a href="#B58-applsci-14-08271" class="html-bibr">58</a>] under a Creative Commons license. (<b>b</b>) Conformational changes during gp120 folding and signal-peptide cleavage. Stage I: Upon the completion of translation, gp120 remains largely unfolded. An α-helical structure around the signal peptide impedes the cleavage of the signal peptide. Stage II: As long as the signal peptide is attached, the cysteine at position 28 (C28) promotes intramolecular disulfide isomerization by interacting with downstream cysteine residues. The tethering of the N-terminus restricts conformational freedom, potentially aiding the folding of the N-terminal region. Stage III: The folding and integration of the inner-domain β sandwich disrupt the helical structure of the signal peptide, exposing the consensus cleavage site. Stage IV: The cleavage of the signal peptide stabilizes the conformation of gp120 by removing the free sulfhydryl group of C28, thus halting further disulfide isomerization. In the schematic, the inner domain is shaded in gray, the outer domain is pink, the variable loops are green, and the signal peptide is orange. Solid lines represent experimentally determined interactions, while dashed lines denote predicted interactions. Reprinted (adapted) from [<a href="#B34-applsci-14-08271" class="html-bibr">34</a>] under a Creative Commons license.</p>
Full article ">Figure 3
<p>Structural Analysis of the HIV-1 Entry Protein Env. (<b>a</b>) Crystal structure of unliganded HIV-1 BG505 SOSIP.664 Env trimer: shown in a ribbon diagram, this structure (PDB ID: 4ZMJ [<a href="#B83-applsci-14-08271" class="html-bibr">83</a>]) lacks the MPER, TMD, and CT, with gp120 in cyan and gp41 in yellow. (<b>b</b>) EM density of the unliganded HIV-1 BaL Env spike: A 3D reconstruction of the unliganded HIV-1 BaL Env spike on the virion surface by cryo-electron tomography (EMDB ID: EMD-5019 for the Env portion; EMDB ID: EMD-5022 for the membrane portion) is presented in gray. (<b>c</b>) Cryo-EM structure of detergent-solubilized Clade B HIV-1 JR-FL EnvΔCT: This structure, complexed with broadly neutralizing antibody PGT151 (PDB ID: 5FUU [<a href="#B67-applsci-14-08271" class="html-bibr">67</a>]), shows gp120 in cyan, gp41 in yellow, and PGT-151 Fab in gray, lacking the CT. (<b>d</b>) NMR structure of MPER–TMD in bicelles: Reconstituted in bicelles mimicking a lipid bilayer, the NMR structure (PDB ID: 6E8W [<a href="#B75-applsci-14-08271" class="html-bibr">75</a>]) illustrates the MPER in orange and the TMD in brown. Reprinted (adapted) with permission from [<a href="#B22-applsci-14-08271" class="html-bibr">22</a>] Copyright Elsevier (2024). (<b>e</b>) Structural model of the MPER−TMD−CT trimer: This model in q ≈ 0.55 bicelles is derived from integrated NMR data of four fragments: MPER-TMD, TMD, TMD-CT LLP2, and TMD-CT. The position of the structure relative to the bilayer region of the bicelle was determined using the Paramagnetic Probe Technique (PPT). (<b>f</b>) Top View of the Trimeric Complex: The top view from the MPER perspective reveals the inner and outer rings of the baseplate, shaded in pink and blue, respectively. Red spheres indicate the palmitoylation sites at residues 764 and 837. (<b>g</b>) Hydrophobic and polar clusters: Residues forming hydrophobic clusters are shaded in yellow, while polar clusters are shaded in blue. (<b>h</b>) MPER−TMD−CT fitting to EM map density: The fit of the MPER−TMD−CT trimer to the assigned MPER density in the EM map of the HIV-1 Env trimer (EMDB ID: EMD-21412; ~11 Å resolution) is based on a cryo-ET study of the Env on the virion surface. The SOSIP.664 crystal structure (PDB ID: 5T3Z) was used to fit the ectodomain EM density. Reprinted (adapted) with permission from [<a href="#B28-applsci-14-08271" class="html-bibr">28</a>] Copyright 2024 American Chemical Society.</p>
Full article ">Figure 4
<p>HIV-1 Env gp120 and gp41 Structures. (<b>a</b>) Ribbon diagram of gp120 core: The ribbon diagram of the gp120 core displays its α-helices (α1–α5) and β-strands (1–25). The relative positions of the variable loops (V1–V5) and the N- and C-termini are indicated. In this orientation, the viral membrane is positioned at the top and the cell membrane at the bottom. Upon binding to CD4, gp120 forms a bridging sheet composed of four β-strands, which separate the inner and outer domains of gp120 with their orientation in the trimeric complex. (<b>b</b>) Ribbon diagram of gp120 core with interaction sites: This diagram shows the gp120 core as in (<b>a</b>), highlighting the N-terminus (red) and the gp41 interaction site (blue). The inner domain is depicted in red and gray, while the outer domain is shown in orange. The bridging sheet, which includes elements from the inner and outer domains, is illustrated in gray and orange. (<b>c</b>) Trimeric gp120 bound to CD4 and antibody Fab: The trimeric gp120 is shown in the same colors as in (<b>b</b>), bound to three molecules of CD4 (yellow) and the Fab fragment from the neutralizing antibody 17b (brown), used to stabilize the gp120 structure. This complex is superimposed onto the electron density observed by cryo-electron tomography (light gray). The right side shows the same structure rotated 90°, positioning the viral membrane in the plane of the page. (<b>d</b>) Three-dimensional representation of HIV-1 Env in CD4-bound conformation: On the left, a trimeric Env spike (blue) is anchored in the lipid bilayer of the viral membrane (gray). The white arrow indicates the predicted location of gp41. On the right, a ribbon diagram of the gp120 core (red) is superimposed on the density map (blue), with the V1/V2 loop in yellow and the V3 loop in green. Reprinted (adapted) with permission from [<a href="#B31-applsci-14-08271" class="html-bibr">31</a>] Copyright Elsevier (2024).</p>
Full article ">Figure 5
<p>HIV-1 gp41 CT Topology and Structure. (<b>a</b>) gp41 CT topology. Traditional model (<b>left</b>): The conventional model of the gp41 CT posits a single membrane-spanning domain, likely found in virions. This single-pass model places the entire CT inside the virion (internal). Sites of mutation that confer resistance to amprenavir (AME), which subsequently became new HIV-1 protease (PR) cleavage sites, are marked by black arrows. The epitope known as the “KS,” recognized by neutralizing antibodies, is shown in blue. Alternative model (<b>right</b>): An alternative topology suggests a three-membrane-spanning configuration for the gp41 CT, which exposes portions of the gp41 CT, including the KS, to the extracellular space. This topology may also occur with a detectable frequency in HIV-1 Env-expressing cells alongside the traditional model. Reprinted (adapted) with permission from [<a href="#B31-applsci-14-08271" class="html-bibr">31</a>] Copyright Elsevier (2024). (<b>b</b>) Segmented EM density map of gp41 trimer. The segmented EM density map of the gp41 trimer, with gp120 removed, shows the structural arrangement: The C-terminal half of HR1 (rust) forms a central three-helix bundle. The C-terminal half of HR2 (yellow) wraps helically around the trimer base. Unassigned density (beige) likely corresponds to the intervening region between HR1 and HR2, including the disulfide loop and elements from gp120 (C1 and C5). Density parallel to HR1 (brown) likely corresponds to the N-terminal half of HR1, the fusion peptide proximal region (FPPR), and the fusion peptide (FP). (<b>c</b>) Modeled portion corresponding to EM density maps shows the modeled portion of gp41. (<b>d</b>) Three-helix bundle in PGV04-bound trimer. The EM density map illustrates the three-helix bundle formed by HR1 in the structure bound by the PGV04 antibody. (<b>e</b>) Overlay of EM densities compares the three-helix bundle EM density (orange) from the PGV04-bound structure, filtered to 9.5 Å, with the 9 Å reconstruction of a 17b-bound SOSIP gp140 trimer (gray, EMDB-5462). (<b>f</b>) Reconstruction of SOSIP trimer with deletions. An 8.2 Å reconstruction of a SOSIP trimer from which the last 14 amino acids were deleted (SOSIP.650:PGV04) is shown. The difference between the SOSIP.650 and SOSIP.664 maps corresponds to a short helical segment (red) at the end of HR2 that projects toward the adjacent protomer. Reprinted (adapted) with permission from [<a href="#B123-applsci-14-08271" class="html-bibr">123</a>] Copyright Science (2024).</p>
Full article ">
18 pages, 8330 KiB  
Article
Genomic Characterization of Phage ZP3 and Its Endolysin LysZP with Antimicrobial Potential against Xanthomonas oryzae pv. oryzae
by Muchen Zhang, Xinyan Xu, Luqiong Lv, Jinyan Luo, Temoor Ahmed, Waleed A. A. Alsakkaf, Hayssam M. Ali, Ji’an Bi, Chengqi Yan, Chunyan Gu, Linfei Shou and Bin Li
Viruses 2024, 16(9), 1450; https://doi.org/10.3390/v16091450 - 11 Sep 2024
Viewed by 600
Abstract
Xanthomonas oryzae pv. oryzae (Xoo) is a significant bacterial pathogen responsible for outbreaks of bacterial leaf blight in rice, posing a major threat to rice cultivation worldwide. Effective management of this pathogen is crucial for ensuring rice yield and food security. In this [...] Read more.
Xanthomonas oryzae pv. oryzae (Xoo) is a significant bacterial pathogen responsible for outbreaks of bacterial leaf blight in rice, posing a major threat to rice cultivation worldwide. Effective management of this pathogen is crucial for ensuring rice yield and food security. In this study, we identified and characterized a novel Xoo phage, ZP3, isolated from diseased rice leaves in Zhejiang, China, which may offer new insights into biocontrol strategies against Xoo and contribute to the development of innovative approaches to combat bacterial leaf blight. Transmission electron microscopy indicated that ZP3 had a short, non-contractile tail. Genome sequencing and bioinformatic analysis showed that ZP3 had a double-stranded DNA genome with a length of 44,713 bp, a G + C content of 52.2%, and 59 predicted genes, which was similar to other OP1-type Xoo phages belonging to the genus Xipdecavirus. ZP3’s endolysin LysZP was further studied for its bacteriolytic action, and the N-terminal transmembrane domain of LysZP is suggested to be a signal–arrest–release sequence that mediates the translocation of LysZP to the periplasm. Our study contributes to the understanding of phage–Xoo interactions and suggests that phage ZP3 and its endolysin LysZP could be developed into biocontrol agents against this phytopathogen. Full article
(This article belongs to the Special Issue Recent Advances in Phage-Plant Interactions)
Show Figures

Figure 1

Figure 1
<p>Characteristics of phage ZP3. (<b>A</b>) Phage plaques of ZP3 infecting Xoo strain G5 in double-layer agar plates. (<b>B</b>) Transmission electron microscopy image of ZP3. (<b>C</b>) Bacterial growth curve of Xoo strain G5 with and without phage ZP3. (<b>D</b>) Adsorption rate of phage ZP3. (<b>E</b>) One-step growth curve of ZP3.</p>
Full article ">Figure 2
<p>Biophysical stability of phage ZP3. (<b>A</b>) The temperature stability of ZP3. (<b>B</b>) The pH stability of ZP3. The ethanol (<b>C</b>) and isopropanol (<b>D</b>) stability of ZP3.</p>
Full article ">Figure 3
<p>(<b>A</b>) Phage ZP3’s annotated genome map. Arrows are used to denote the predicted genes, and each arrow’s orientation indicates the transcription direction. Genes in the three functional modules are represented in orange (phage structure), blue (phage DNA packaging and replication), red (host lysis), and gray (hypothetical protein). (<b>B</b>) ZP3 phage neighbor-joining phylogenetic tree analysis based on terminase large subunit amino acid sequences; 1000 replications of the bootstrap values. Xoo phages belonging to the genus <span class="html-italic">Xipdecavirus</span> are marked.</p>
Full article ">Figure 4
<p>In silico characterization of the ZP3 endolysin. (<b>A</b>) Genomic organization and schematic representation of LysZP. Amino acid position is indicated by numbers. (<b>B</b>) Three-dimensional structure prediction of LysZP. (<b>C</b>) Annotation of LysZP protein. (<b>D</b>) Sequence alignment of LysZP with various Xoo phage endolysins, including X2 (MW435566), Xop411 (ABK00175.1), OP1 (YP_453585.1), OP2 (YP_453642.1), Xp15 (YP_239293.1), XPP1 (YP_010052413.1), and XPV1 (AVO24202.1). Red triangles (E34, D43, and T52) represent the catalytic triad residues. Different colors of the letters mean the homology level (black: 100%; pink: ≥ 75%; blue: ≥ 50%).</p>
Full article ">Figure 5
<p>Effects of LysZP and HolZP on bacterial growth. (<b>A</b>) Growth curves of <span class="html-italic">E. coli</span> expressing LysZP, HolZP, LysZP+HolZP, and LysZPΔTMD after IPTG induction. (<b>B</b>) Live and dead bacterial fluorescent staining experiment (green for live cells and red for dead cells). (<b>C</b>) Flow cytometry scatter plots.</p>
Full article ">Figure 6
<p>Effects of lysZP and HolZP on cell membrane integrity. (<b>A</b>) Morphological changes in bacterial cells under TEM. (<b>B</b>,<b>C</b>) β-galactosidase activity after IPTG induction for 0.5 h, 3 h, and 6 h. Color change in plates (<b>B</b>) and the corresponding OD<sub>420</sub> values (<b>C</b>). Columns with different letters (a–c) are significantly different according to the LSD test (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 7
<p>Bacteriolysis effect of LysZP on Xoo_C2R. Xoo_C2R growth was measured with and without LysZP protein treatment. The three replicates’ standard deviations are shown by the error bars. LysZP against Xoo_C2R was tested through measuring the width of the inhibition zone.</p>
Full article ">
12 pages, 5157 KiB  
Article
Comparative Analysis of the Mitochondrial Genome of Eggplant (Solanum melongena L.) to Identify Cytoplasmic Male Sterility Candidate Genes
by Wentao Deng, Guiyun Gan, Weiliu Li, Chuying Yu, Yaqin Jiang, Die Li, Qihong Yang, Wenjia Li, Peng Wang and Yikui Wang
Int. J. Mol. Sci. 2024, 25(17), 9743; https://doi.org/10.3390/ijms25179743 - 9 Sep 2024
Viewed by 593
Abstract
Cytoplasmic male sterility (CMS) is important for commercial hybrid seed production. However, it is still not used in eggplant (Solanum melongena L.), and corresponding regulatory genes and mechanisms of action have not been reported. We report CMS line 327A, which was derived [...] Read more.
Cytoplasmic male sterility (CMS) is important for commercial hybrid seed production. However, it is still not used in eggplant (Solanum melongena L.), and corresponding regulatory genes and mechanisms of action have not been reported. We report CMS line 327A, which was derived from the hybridization between cultivated and wild eggplants. By looking at different stages of anther development under a microscope, we saw that the 327A anther’s tapetum layer vacuolized during meiosis, which caused abortion. To investigate the 327A CMS regulatory genes, the mitochondrial genomes of 327A and its maintainer line 327B were assembled de novo. It was found that 15 unique ORFs (Open Reading Frame) were identified in 327A. RT-PCR and RT-QPCAR tests confirmed that orf312a and orf172a, 327A-specific ORFs with a transmembrane domain, were strongly expressed in sterile anthers of 327A. In addition, orf312a has a chimeric structure with the ribosomal protein subunit rpl16. Therefore, orf312a and orf172a can be considered strong candidate genes for CMS. Concurrently, we analyzed the characteristics of CMS to develop a functional molecular marker, CMS312, targeting a future theoretical basis for eggplant CMS three-line molecular breeding. Full article
(This article belongs to the Section Molecular Genetics and Genomics)
Show Figures

Figure 1

Figure 1
<p>The mature flower morphology of CMS line 327A and maintainer line 327B. (The left side displays CMS mature flowers and bud morphology, while the right side shows maintainer mature flowers and bud morphology. Scale: 1 cm.)</p>
Full article ">Figure 2
<p>Analyzing the anther cytology of 327A and 327B at different periods. (<b>A</b>–<b>F</b>) are anther paraffin sections of maintainer line 327B at different periods; (<b>G</b>–<b>L</b>) are paraffin sections of anther of sterile line 327A at different periods. (<b>A</b>,<b>G</b>) are spore-forming cell stages; (<b>B</b>,<b>H</b>) are the microspore mother stage; (<b>C</b>,<b>I</b>) are the meiosis phase; (<b>D</b>,<b>J</b>) are the tetrad periods; (<b>E</b>,<b>K</b>) are the binuclear phase; (<b>F</b>,<b>L</b>) is the mature stage. E: epidermal layer; EN: endodermis; ML: middle layer; T: tapetum layer; Sp: spore-forming cell; MMC: pollen mother cell; Msp: microspore; PG: pollen grains; dT: apoptotic tapetum layer; dTM: apoptotic tapetum and microspore. Scale= 100 μm.)</p>
Full article ">Figure 3
<p>Mitochondrial assembly maps of sterile line 327A and maintainer line 327B of eggplant. (The outer circle is the location coordinates of genomic components such as genes and ncRNAs, with corresponding gene names; the inner circle is the genomic GC content; different colored blocks represent the functions of different gene products.)</p>
Full article ">Figure 4
<p>The inner circle of the mitochondrial genome structure comparison diagramis the maintainer line 327B genome, and the outer circle is the sterile line 327A genome. And this diagram plotted by SyRI (Max Planck Institute for Plant Breeding Research, Planegg-Martinsried, Germany) [<a href="#B25-ijms-25-09743" class="html-bibr">25</a>]. (Collinear: homo-linear region; Translocation: translocation region; Inversion: an inversion area; Tran+Inver: translocation and inversion region; Insertion: insertion areas of 50 bp or longer; Deletion: deletion area of length greater than or equal to 50 bp; Complex InDel: regions that do not match but correspond in location; Forward chain: The forward chain of the genome sequence, where the gene coordinates increase in the clockwise direction; Reverse chain: The reverse chain of the genome sequence, where the gene coordinates increase counterclockwise.)</p>
Full article ">Figure 5
<p>327A-specific ORF analysis and identification of CMS candidate genes. (<b>A</b>) collinearity comparison between 327A and 327B; (<b>B</b>) orf561a-rpl16-orf312a co-transcription chimera structure and corresponding transmembrane domain; (<b>C</b>) orf561a-rpl16-orf312a, <span class="html-italic">rpl16</span>, <span class="html-italic">orf172a</span> RT-PCR semi-quantitative analysis; (<b>D</b>) <span class="html-italic">rpl16</span>, <span class="html-italic">orf312a</span>, <span class="html-italic">orf172a</span> RT-QPCR relative expression analysis, **** <span class="html-italic">p</span> &lt; 0.0001, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 6
<p>CMS markers were verified in 17 eggplant inbred lines.</p>
Full article ">
13 pages, 3701 KiB  
Article
Influenza A Vaccine Candidates Based on Virus-like Particles Formed by Coat Proteins of Single-Stranded RNA Phages Beihai32 and PQ465
by Egor A. Vasyagin, Anna A. Zykova, Eugenia S. Mardanova, Nikolai A. Nikitin, Marina A. Shuklina, Olga O. Ozhereleva, Liudmila A. Stepanova, Liudmila M. Tsybalova, Elena A. Blokhina and Nikolai V. Ravin
Vaccines 2024, 12(9), 1033; https://doi.org/10.3390/vaccines12091033 - 9 Sep 2024
Viewed by 687
Abstract
Efficient control of influenza A infection can potentially be achieved through the development of broad-spectrum recombinant vaccines based on conserved antigens. The extracellular domain of the transmembrane protein M2 of influenza A virus (M2e) is highly conserved but poorly immunogenic and needs to [...] Read more.
Efficient control of influenza A infection can potentially be achieved through the development of broad-spectrum recombinant vaccines based on conserved antigens. The extracellular domain of the transmembrane protein M2 of influenza A virus (M2e) is highly conserved but poorly immunogenic and needs to be fused to an adjuvant protein or carrier virus-like particles (VLPs) to increase immunogenicity and provide protection against infection. In this study, we obtained VLPs based on capsid proteins (CPs) of single-stranded RNA phages Beihai32 and PQ465 bearing the M2e peptides. Four copies of the M2e peptide were linked to the C-terminus of the CP of phage Beihai32 and to the N and C termini of the CP of phage PQ465. The hybrid proteins, being expressed in Escherichia coli, formed spherical VLPs of about 30 nm in size. Immunogold transmission electron microscopy showed that VLPs formed by the phage PQ465 CP with a C-terminal M2e fusion present the M2e peptide on the surface. Subcutaneous immunization of mice with VLPs formed by both CPs containing four copies of the M2e peptide at the C termini induced high levels of M2e-specific IgG antibodies in serum and provided mice with protection against lethal influenza A virus challenge. In the case of an N-terminal fusion of M2e with the phage PQ465 CP, the immune response against M2e was significantly lower. CPs of phages Beihai32 and PQ465, containing four copies of the M2e peptide at their C termini, can be used to develop recombinant influenza A vaccine. Full article
(This article belongs to the Special Issue Bioengineering in Vaccine Design and Delivery)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Schematic representation of the expression vectors. PT5, bacteriophage T5 promoter for <span class="html-italic">E. coli</span> RNA polymerase, with embedded <span class="html-italic">lac</span> operator; T0, lambda T0 terminator; PT7, bacteriophage T7 promoter; T7, bacteriophage T7 terminator; 4M2e, four tandem copies of the M2e peptide; PQ465, coat protein of bacteriophage PQ465; Beihai32, coat protein of bacteriophage Beihai32. The 6his tag and 19S linker are shown by orange and yellow boxes, respectively. (<b>b</b>) The structures of single monomeric proteins were predicted using Alphafold v.2.3.1 [<a href="#B28-vaccines-12-01033" class="html-bibr">28</a>] and visualized using the SWISS MODEL server [<a href="#B29-vaccines-12-01033" class="html-bibr">29</a>].</p>
Full article ">Figure 2
<p>Expression of recombinant proteins in <span class="html-italic">E. coli</span>. The proteins isolated from <span class="html-italic">E. coli</span> were analyzed by SDS-PAGE. M, molecular weight marker (kD); proteins isolated from <span class="html-italic">E. coli</span> cells before (lane 1) and after (lane 2) induction of expression. Proteins: Beihai32 (<b>a</b>), Beihai32_4M2e (<b>b</b>), PQ465 (<b>c</b>), 4M2e_PQ465 (<b>d</b>), PQ465_4M2e (<b>e</b>). Positions of the target proteins are indicated by arrows.</p>
Full article ">Figure 3
<p>Purification of the recombinant proteins. Purified proteins were analyzed by SDS-PAGE. M, molecular weight marker (kD); lanes: 1, Beihai32; 2, Beihai32_4M2e; 3, PQ465; 4, 4M2e_PQ465; 5, PQ465_4M2e.</p>
Full article ">Figure 4
<p>Analysis of VLPs formed by the recombinant proteins by transmission electron microscopy (<b>a</b>) and immunogold transmission electron microscopy (<b>b</b>). Scale bar is 50 nm.</p>
Full article ">Figure 5
<p>Antigenicity of VLPs. Two-fold dilutions of VLPs formed by Beihai32, Beihai32_4M2e, PQ465, PQ465_4M2e, and 4M2e_PQ465 proteins were coated on ELISA plates and then probed with antibodies specific for M2e.</p>
Full article ">Figure 6
<p>Antibody response in serum. BALB/c mice (5 per group) were subcutaneously immunized with VLPs three times at a two-week interval. Mice in the control group were injected with PBS. Two weeks after the third immunization, M2e-specific IgG titers were evaluated by ELISA. Data are presented as the anti-M2e IgG titers for individual mice with geometric mean titers ± SEM (Standard Error of the Mean) determined in each group. Statistically significant differences between groups are indicated (**, <span class="html-italic">p</span> &lt; 0.01; ****, <span class="html-italic">p</span> &lt; 0.0001). (<b>a</b>) Immunization with Beihai32_4M2e and Beihai32; (<b>b</b>) immunization with PQ465_4M2e, 4M2e_PQ465, and PQ465.</p>
Full article ">Figure 7
<p>Protective efficiency of the recombinant proteins. BALB/c mice (10 per group) were subcutaneously immunized with VLPs three times at a two-week interval. Mice in the control group were injected with PBS. Two weeks after the third immunization, mice were challenged with 5LD<sub>50</sub> of influenza strain A/Aichi/2/68 (H3N2). Survival of challenged mice was monitored for 14 days post-challenge. Statistically significant differences between groups are indicated (*, 0.01 &lt; <span class="html-italic">p</span> &lt; 0.05; **, 0.001 &lt; <span class="html-italic">p</span> &lt; 0.01; ***, <span class="html-italic">p</span> &lt; 0.001). (<b>a</b>) Immunization with Beihai32_4M2e; (<b>b</b>) immunization with PQ465_4M2e and 4M2e_PQ465.</p>
Full article ">
18 pages, 3225 KiB  
Article
A Novel Rare PSEN2 Val226Ala in PSEN2 in a Korean Patient with Atypical Alzheimer’s Disease, and the Importance of PSEN2 5th Transmembrane Domain (TM5) in AD Pathogenesis
by YoungSoon Yang, Eva Bagyinszky and Seong Soo A. An
Int. J. Mol. Sci. 2024, 25(17), 9678; https://doi.org/10.3390/ijms25179678 - 6 Sep 2024
Viewed by 624
Abstract
In this manuscript, a novel presenilin-2 (PSEN2) mutation, Val226Ala, was found in a 59-year-old Korean patient who exhibited rapid progressive memory dysfunction and hallucinations six months prior to her first visit to the hospital. Her Magnetic Resonance Imaging (MRI) showed brain atrophy, and [...] Read more.
In this manuscript, a novel presenilin-2 (PSEN2) mutation, Val226Ala, was found in a 59-year-old Korean patient who exhibited rapid progressive memory dysfunction and hallucinations six months prior to her first visit to the hospital. Her Magnetic Resonance Imaging (MRI) showed brain atrophy, and both amyloid positron emission tomography (PET) and multimer detection system-oligomeric amyloid-beta (Aβ) results were positive. The patient was diagnosed with early onset Alzheimer’s disease. The whole-exome analysis revealed a new PSEN2 Val226Ala mutation with heterozygosity in the 5th transmembrane domain of the PSEN2 protein near the lumen region. Analyses of the structural prediction suggested structural changes in the helix, specifically a loss of a hydrogen bond between Val226 and Gln229, which may lead to elevated helix motion. Multiple PSEN2 mutations were reported in PSEN2 transmembrane-5 (TM5), such as Tyr231Cys, Ile235Phe, Ala237Val, Leu238Phe, Leu238Pro, and Met239Thr, highlighting the dynamic importance of the 5th transmembrane domain of PSEN2. Mutations in TM5 may alter the access tunnel of the Aβ substrate in the membrane to the gamma-secretase active site, indicating a possible influence on enzyme function that increases Aβ production. Interestingly, the current patient with the Val226Ala mutation presented with a combination of hallucinations and memory dysfunction. Although the causal mechanisms of hallucinations in AD remain unclear, it is possible that PSEN2 interacts with other disease risk factors, including Notch Receptor 3 (NOTCH3) or Glucosylceramidase Beta-1 (GBA) variants, enhancing the occurrence of hallucinations. In conclusion, the direct or indirect role of PSEN2 Val226Ala in AD onset cannot be ruled out. Full article
(This article belongs to the Special Issue Genetic Research in Neurological Diseases)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Magnetic resonance imaging of the patient: observations of Axial FLAIR (A), (B), (C) sequences of the patient with mild diffuse brain atrophy. (<b>b</b>) Amyloid PET image of the patient: abnormal amyloid deposits observed in gray matter of whole brain, especially in the left temporal lobe. (A) Coronal plane. (B) Axial plane.</p>
Full article ">Figure 2
<p>Sanger sequencing data of patient with PSEN2 Val226Ala mutation.</p>
Full article ">Figure 3
<p>ExPASy predictions for PSEN2 Val226Ala, compared to normal PSEN2 and PSEN2 Val226Ala structure in terms of polarity, Kyte-Doolittle Hydropathy Plots and bulkiness index. The X axis present the residues in PSEN2 (between residue 215 and 227), while the Y axis presents the (<b>a</b>) polarity scores (<b>b</b>) the Kyte-Doolittle Hydropathy Plots (<b>c</b>) and the bulkiness index.</p>
Full article ">Figure 4
<p>(<b>a</b>) Aligned normal and mutant PSEN2 structures. (<b>b</b>) Intramolecular interactions in case of Val226. (<b>c</b>) Intramolecular interactions in case of Ala226. (<b>d</b>) 2D diagram of the intramolecular interaction of Val226 vs. Ala226. The residues which Val226 or Ala226 bind to as covalent bonds are labeled with purple, the hydrogen bonds are labeled with blue, and the Van der Waals bonds are labeled with green.</p>
Full article ">Figure 5
<p>Three-dimensional model of structure of mutations, located in TM5 of PSEN2. (<b>a</b>) Leu225Pro, (<b>b</b>) Glu228Leu, (<b>c</b>) Tyr231Cys, (<b>d</b>) Ile235Phe, (<b>e</b>) Met237Val, (<b>f</b>) Leu238Phe, (<b>g</b>) Leu238Pro, (<b>h</b>) Met239Val, (<b>i</b>) Met239Thr, and (<b>j</b>) Met239Ile.</p>
Full article ">Figure 6
<p>Mutations, located in the 5th transmembrane domain of PSEN2. Variants, which are highlighted in red, were verified to impact amyloid metabolism in cell lines, which are highlighted in red. The variants of which the pathogenic nature remained unclear are highlighted in orange. The location of Val226 is highlighted in yellow.</p>
Full article ">
16 pages, 2202 KiB  
Article
Genome-Wide Identification, Evolution, and miRNA-22 Regulation of Kruppel-Like Factor (KLF) Gene Family in Chicken (Gallus gallus)
by Zheng Ma, Huangbin Chu, Fapei Li, Guochao Han, Yingqiu Cai, Jianing Yi, Mingrou Lu, Hai Xiang, Huimin Kang, Fei Ye, Siyu Chen and Hua Li
Animals 2024, 14(17), 2594; https://doi.org/10.3390/ani14172594 - 6 Sep 2024
Viewed by 475
Abstract
Krüppel-like factors (KLFs) are a class of fundamental transcription factors that are widely present in various eukaryotes from nematodes to humans, named after their DNA binding domain which is highly homologous to the Krüppel factor in fruit flies. To investigate the composition, organization, [...] Read more.
Krüppel-like factors (KLFs) are a class of fundamental transcription factors that are widely present in various eukaryotes from nematodes to humans, named after their DNA binding domain which is highly homologous to the Krüppel factor in fruit flies. To investigate the composition, organization, and evolutionary trajectory of KLF gene family members in chickens, in our study, we leveraged conserved sequences of KLF genes from representative classes across fish, amphibians, birds, and mammals as foundational sequences. Bioinformatic tools were employed to perform homology alignment on the chicken genome database, ultimately identifying the KLF family members present in chickens. The gene structure, phylogenetic analysis, conserved base sequences, physicochemical properties, collinearity analysis, and protein structure were then analyzed using bioinformatic tools. Additionally, the impact of miRNA-22, related to poultry lipid metabolism, on the expression of the KLF gene family in the liver, heart, and muscle of Qingyuan partridge chickens was explored. The results showed that: (1) compared to fish, the KLF family in birds is more closely related to mammals and amphibians; (2) KLFs within the same subgroups are likely to be derived from a common ancestral gene duplication; (3) KLF3/8/12 in the same subgroup may have some similar or overlapping functions; (4) the motif 4 of KLF5 was most likely lost during evolution; (5) KLF9 may perform a similar function in chickens and pigs; (6) there are collinear relationships between certain KLF genes, indicating that there are related biomolecular functions between these KLF genes; (7) all members of the KLF family in chickens are non-transmembrane proteins; and (8) interference and overexpression of miRNA-22 in Qingyuan partridge chickens can affect the expression levels of KLF genes in liver, heart, and muscle. Full article
(This article belongs to the Section Animal Genetics and Genomics)
Show Figures

Figure 1

Figure 1
<p>The distribution of chicken <span class="html-italic">KLF</span> gene family members on chromosomes.</p>
Full article ">Figure 2
<p>The unrooted phylogenetic tree of the <span class="html-italic">KLF</span> gene family in <span class="html-italic">Gallus gallus</span> (chicken), <span class="html-italic">Sus scrofa</span> (pig), <span class="html-italic">Xenopus tropicalis</span> (tropical clawed frog), and <span class="html-italic">Danio rerio</span> (zebrafish).</p>
Full article ">Figure 3
<p>The gene structure of <span class="html-italic">KLF</span> genes in chicken. The green boxes and perple boxes in the gene structure diagram represent UTR and CDS.</p>
Full article ">Figure 4
<p>Phylogenetic relationships and conserved motif of <span class="html-italic">KLF</span> genes in chicken.</p>
Full article ">Figure 5
<p>Collinearity between chicken <span class="html-italic">KLF</span> gene families. The outermost ring displays the names of predicted or validated transcripts for each gene. The gray bands in the outer ring represent chromosomes, with the red numbers or letters inside indicating chromosome numbers. The middle ring, featuring red peaks and blue bands, represents gene density on each chromosome. The innermost layer shows colored lines representing the synteny relationships between KLF genes. The top right section indicates the gene density bands.</p>
Full article ">Figure 6
<p>Prediction of KLF9 protein tertiary structure.</p>
Full article ">Figure 7
<p>Prediction of KLF9 protein transmembrane structure.</p>
Full article ">Figure 8
<p>The impact of miR-22 on <span class="html-italic">KLF</span> gene relative expression in different tissues. (<b>a</b>) Shows the expression of KLFs in the liver after lentiviral miR-22 interference (miR-22-I) or overexpression (miR-22-M); (<b>b</b>) depicts the expression of KLFs in the heart following lentiviral miR-22 interference (miR-22-I); (<b>c</b>) illustrates the expression of KLFs in the pectoral muscle after lentiviral miR-22 interference (miR-22-I). FPKM represents the normalized expression levels across different samples. Blue indicates lower expression and red indicates higher expression (<b>b</b>,<b>c</b>). * indicate significant differences between the different groups (n = 3, <span class="html-italic">p</span> &lt; 0.05, Student’s t-test).</p>
Full article ">
16 pages, 4447 KiB  
Article
Molecular Characterization, Evolution and Expression Analysis of Ammonium Transporter from Four Closely Related Bactrocera Species (Tephritidae)
by Jie Zhang, Qi Wang, Chenhao Liu, Jiaying Liu, Qian Qian, Chuanjian Ru, Leyuan Liu, Shanchun Yan, Wei Liu and Guirong Wang
Life 2024, 14(9), 1114; https://doi.org/10.3390/life14091114 - 4 Sep 2024
Viewed by 517
Abstract
Numerous insects are attracted to low levels of ammonia, utilizing it as a cue to locate food sources. The Ammonium Transporter (Amt), a highly conserved, atypical olfactory receptor, has been shown to mediate the detection of ammonia in insects. While the attraction of [...] Read more.
Numerous insects are attracted to low levels of ammonia, utilizing it as a cue to locate food sources. The Ammonium Transporter (Amt), a highly conserved, atypical olfactory receptor, has been shown to mediate the detection of ammonia in insects. While the attraction of Tephritidae to ammonia is well established, knowledge about the Amt in this family is limited. The species Bactrocera dorsalis (Hendel 1912), Bactrocera cucurbitae (Coquillett 1899), Bactrocera correcta Bezzi 1916 and Bactrocera tau (Walker 1849), which are common agricultural pests within Tephritidae, exhibit numerous ecological similarities, offering a solid foundation for studying Amt characteristics in this family. In this study, we elucidated the sequences, evolutionary relationships, and expression patterns of Amt in these four species. The results indicated that these Amts share the same open reading frame, containing 1770 bp that encode a protein of 589 amino acid residues. These Amt proteins exhibit the typical structural characteristics of Amts, including an 11-transmembrane domain with an extracellular N-terminus and an intracellular C-terminus. They also have the ability to form trimers in the membrane. Additionally, they contain three conserved amino acid residues essential for ammonia transport: A189, H195, and H352. Phylogenetic and expression pattern analyses showed that they are highly conserved in Diptera and are significantly expressed in antennae. This study is the first report characterizing the Amt gene in four Tephritidae species. These findings provide a foundation for further exploration into the roles of these genes in their particular biological contexts. Full article
Show Figures

Figure 1

Figure 1
<p>The identification and sequence characteristics of Amts from <span class="html-italic">B. dorsalis</span> (BdorAmt), <span class="html-italic">B. cucurbitae</span> (<span class="html-italic">BcucAmt</span>), <span class="html-italic">B. correcta</span> (<span class="html-italic">BcorAmt</span>), and <span class="html-italic">B. tau</span> (<span class="html-italic">BtauAmt</span>). (<b>A</b>) Electrophoresis of four <span class="html-italic">Amt</span> PCR products and prediction of the physicochemical properties of the encoded proteins. (<b>B</b>) Multiple sequence alignment of amino acid sequences of Amts from <span class="html-italic">B. tryoni</span>, <span class="html-italic">B. neohumeralis</span>, <span class="html-italic">B. latifrons</span>, <span class="html-italic">B. oleae</span>, <span class="html-italic">A. ludens</span>, <span class="html-italic">C. capitata</span>, <span class="html-italic">A. obliqua</span>, <span class="html-italic">R. pomonella</span>, <span class="html-italic">T. dalmanni</span>, <span class="html-italic">M. vetustissima</span>, <span class="html-italic">M. domestica</span>, <span class="html-italic">L. cuprina</span>, <span class="html-italic">D. melanogaster</span>, <span class="html-italic">A. aegypti</span>, <span class="html-italic">A. gambiae</span>, and <span class="html-italic">A. coluzzii</span>; the accession numbers of the sequences used are listed in <a href="#app2-life-14-01114" class="html-app">Appendix B</a>. (<b>C</b>) Transmembrane topology profile prediction of four <span class="html-italic">Amts</span>.</p>
Full article ">Figure 2
<p>Phylogenetic analysis of Diptera Amts and comparison of amino acid residue similarity; the accession numbers of the sequences used are listed in <a href="#app2-life-14-01114" class="html-app">Appendix B</a>. (<b>A</b>) Phylogenetic relationships of Diptera Amt sequences. (<b>B</b>) Percent identities of amino acid residues between the Amts from different Diptera species.</p>
Full article ">Figure 3
<p>In silico analyses of Amt protein sequence from <span class="html-italic">B. dorsalis</span> (BdorAmt), <span class="html-italic">B. cucurbitae</span> (BcucAmt), <span class="html-italic">B. correcta</span> (BcorAmt), and <span class="html-italic">B. tau</span> (BtauAmt). (<b>A</b>) Protein alignment of Amts from <span class="html-italic">E. coli</span> and <span class="html-italic">B. dorsalis</span>, <span class="html-italic">B. cucurbitae</span>, <span class="html-italic">B. correcta</span>, and <span class="html-italic">B. tau</span>. (<b>B</b>–<b>E</b>) Protein structure prediction for BdorAmt, BcorAmt, BcucAmt, and BtauAmt.</p>
Full article ">Figure 4
<p>The expression levels of <span class="html-italic">Amt</span> in different body parts of <span class="html-italic">B. dorsalis</span> (<span class="html-italic">BdorAmt</span>, <b>A</b>), <span class="html-italic">B. cucurbitae</span> (<span class="html-italic">BcucAmt</span>, <b>B</b>), <span class="html-italic">B. correcta</span> (<span class="html-italic">BcorAmt</span>, <b>C</b>), and <span class="html-italic">B. tau</span> (<span class="html-italic">BtauAmt</span>, <b>D</b>). Each image separately illustrates the results of RNA-Seq quantification and the Quantitative Real-Time PCR (qRT-PCR) verification. Results are presented as the mean ± standard error. <span class="html-italic">p</span> values were determined by a two-tailed unpaired <span class="html-italic">t</span>-test. (<span class="html-italic">** p</span> &lt; 0.01; ns indicates no significant difference).</p>
Full article ">Figure A1
<p>The hydrophilicity/hydrophobicity (<b>A</b>) and confidence scores for the predicted structures of <span class="html-italic">B. dorsalis</span> Amt (BdorAmt), <span class="html-italic">B. cucurbitae</span> Amt (BcucAmt), <span class="html-italic">B. correcta</span> Amt (BcorAmt), and <span class="html-italic">B. tau</span> Amt (BtauAmt) (<b>B</b>,<b>C</b>). In (<b>B</b>), this Ramachandran plot highlights the various regions of polypeptide chain conformations, the regions are marked as follows: A, Alpha helix region (right-handed α-helix); B, Beta sheet region (β-sheet); L, Allowed regions for other conformations; a, Right-handed alpha helix; b, Antiparallel beta sheet; l, Left-handed alpha helix; p, Parallel beta sheet; ~a, Allowed region for right-handed alpha helix; ~b, Allowed region for antiparallel beta sheet; ~l, Allowed region for left-handed alpha helix; ~p, Allowed region for parallel beta sheet.</p>
Full article ">
18 pages, 20293 KiB  
Article
A Non-Canonical p75HER2 Signaling Pathway Underlying Trastuzumab Action and Resistance in Breast Cancer
by Babak Nami and Zhixiang Wang
Cells 2024, 13(17), 1452; https://doi.org/10.3390/cells13171452 - 29 Aug 2024
Viewed by 682
Abstract
Overexpression of HER2 occurs in 25% of breast cancer. Targeting HER2 has proven to be an effective therapeutic strategy for HER2-positive breast cancer. While trastuzumab is the most commonly used HER2 targeting agent, which has significantly improved outcomes, the overall response rate is [...] Read more.
Overexpression of HER2 occurs in 25% of breast cancer. Targeting HER2 has proven to be an effective therapeutic strategy for HER2-positive breast cancer. While trastuzumab is the most commonly used HER2 targeting agent, which has significantly improved outcomes, the overall response rate is low. To develop novel therapies to boost trastuzumab efficacy, it is critical to identify the mechanisms underlying trastuzumab action and resistance. We recently showed that the inhibition of breast cancer cell growth by trastuzumab is not through the inhibition of HER2 canonical signaling. Here we report the identification of a novel non-canonical HER2 signaling pathway and its interference by trastuzumab. We showed that HER2 signaled through a non-canonical pathway, regulated intramembrane proteolysis (RIP). In this pathway, HER2 is first cleaved by metalloprotease ADAM10 to produce an extracellular domain (ECD) that is released and the p95HER2 that contains the transmembrane domain (TM) and intracellular domain (ICD). p95HER2, if further cleaved by an intramembrane protease, γ-secretase, produced a soluble ICD p75HER2 with nuclear localization signal (NLS). p75HER2 is phosphorylated and translocated to the nucleus. Nuclear p75HER2 promotes cell proliferation. Trastuzumab targets this non-canonical HER2 pathway via inhibition of the proteolytic cleavage of HER2 by both ADAM10 and γ-secretase. However, p75HER2 pathway also confers resistance to trastuzumab once aberrantly activated. Combination of trastuzumab with ADAM10 and γ-secretase inhibitors completely blocks p75HER2 production in both BT474 and SKBR3 cells. We concluded that HER2 signals through the RIP signaling pathway that promotes cell proliferation and is targeted by trastuzumab. The aberrant HER2 RIP signaling confers resistance to trastuzumab that could be overcome by the application of inhibitors to ADAM10 and γ-secretase. Full article
Show Figures

Figure 1

Figure 1
<p>Trastuzumab binding to HER2 in breast cancer cell lines. Binding of trastuzumab to HER2 in two CHO cell lines (CHO-K6 and CHO-K13) expressing HER2, breast cancer cell lines with high HER2 expression levels (SKBR3 and BT474), low HER2 expression levels (MCF7 and MDA-MB-231), and HER2-negative breast cell line (MCF10). The cells were treated with 10 μg/mL trastuzumab for 1 h. HER2 was stained by mouse monoclonal antibody 9G6, followed by FITC (green)-conjugated anti-mouse IgG. Trastuzumab was stained by TRITC (red)-conjugated anti-human IgG. Scale bar: 25 μm. The gradient bar indicates HER2 expression level.</p>
Full article ">Figure 2
<p>Trastuzumab inhibits the proliferation of HER2-positive breast cancer cell lines. SKBR3, BT474, and 293T cells were treated with 10 μg/mL trastuzumab, pertuzumab, or their combination for 5 days, and then the cell proliferation was evaluated by MTT assay (absorbance at 540-nanometer wavelengths). Ten micrograms/millilitres human IgG, 10 µM vinorelbine, and 10 μM CP-724714 were used as respectively mock, anti-proliferative, and HER2 inhibitor controls. ****: <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 3
<p>Proteolytic cleavage of HER2 and its inhibition by trastuzumab (<b>A</b>) CHO-HER2 cells were treated with EGF and trastuzumab with various concentrations (0, 0.1, 0.5, 1, 5, and 10 μg/mL), and the cell lysates were immunoblotted with anti-HER2 antibody (A2). (<b>B</b>) BT474 and SKBR3 cells were treated with trastuzumab (10 μg/mL) or normal human IgG (μg/mL), and the cell lysates were immunoblotted with anti-HER2 antibody (A2). (<b>C</b>) the quantification of the data from (<b>B</b>). ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 4
<p>Nuclear localization of p75HER2 and its inhibition by trastuzumab. (<b>A</b>) Subcellular fractionation to show the nuclear p75HER with or without trastuzumab (10 μg/mL). HER2 was detected by antibody to HER2 (A2). (<b>B</b>) Quantification of the data from A. We loaded 1/10th of the total proteins isolated from the plasma membrane (PM) and cytosolic (Cyt) fraction, but 1/4th of the proteins isolated from nucleus. We normalized this in our quantification. (<b>C</b>) Nuclear localization of HER2 C-terminus by immunofluorescence. Cells were double stained with antibodies to HER2 N-(9G6, green) and C-terminus (c-18, red) and counter stained with Dapi (blue). Size bar = 10 μm. (<b>D</b>) Quantification of the data from C. C-18 stain (red) is positive for both full length and p75HER2 and is used for quantification. The nuclear localization of HER2 was expressed as the percentage of nuclear intensity out of the total cell intensity. ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 5
<p>Phosphorylation of p75HER2 and the effects of trastuzumab on BT474 cells were treated with trastuzumab (10 μg/mL) or normal human IgG (10 μg/mL). The phosphorylation and localization of HER2 were examined by immunofluorescence with antibodies to HER2, pY1005 HER2, and pY1139 HER2. Scale bar: 20 μm.</p>
Full article ">Figure 6
<p>Domain structure of HER2 and the proteolytic cleavage sites.</p>
Full article ">Figure 7
<p>Inhibition of γ-secretase by chemical inhibitor reduces the production of p75HER2 and increase the amount of p95HER2. (<b>A</b>) Inhibition by RO-4929097. (<b>B</b>) Quantification of the data from A. (<b>C</b>) Inhibition by Ly411575. (<b>D</b>) Quantification of the data from (<b>C</b>). Each data point is the average of at least 3 repeats. *: <span class="html-italic">p</span> &lt; 0.1; **: <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 8
<p>Effects of γ-secretase inhibitor RO-cj4929097, ADAM10 inhibitor G1254023X, and trastuzumab on the cleavage of HER2 to produce p75HER2 and p95HER2 in BT474 and SKBR3 cells. * <span class="html-italic">p</span> &lt; 0.1, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 9
<p>p75HER2 is mitogenic. (<b>A</b>) subcellular localization of GFP-tagged p75HER2, HER2, and ΔNLSp75HER2 in MCF-7 cells by fluorescence microscopy. (<b>B</b>) MDA and MCF-7 cells were transfected with GFP-tagged p75HER2, HER2, and ΔNLSp75HER2. The cell proliferation with or without trastuzumab (10 μg/mL) treatment was revealed by MTT assay.</p>
Full article ">Figure 10
<p>Inhibition of p75HER2 by the combination of GI254023X (5 µM), RO=4929097 (10 µM), and trastuzumab (10 µM) in BT474 and SKBR3 cells. ***, <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">
25 pages, 3010 KiB  
Article
Transient Adaptation of Toxoplasma gondii to Exposure by Thiosemicarbazone Drugs That Target Ribosomal Proteins Is Associated with the Upregulated Expression of Tachyzoite Transmembrane Proteins and Transporters
by Manuela Semeraro, Ghalia Boubaker, Mirco Scaccaglia, Joachim Müller, Anitha Vigneswaran, Kai Pascal Alexander Hänggeli, Yosra Amdouni, Laura Helen Kramer, Alice Vismarra, Marco Genchi, Giorgio Pelosi, Franco Bisceglie, Manfred Heller, Anne-Christine Uldry, Sophie Braga-Lagache and Andrew Hemphill
Int. J. Mol. Sci. 2024, 25(16), 9067; https://doi.org/10.3390/ijms25169067 - 21 Aug 2024
Viewed by 769
Abstract
Thiosemicarbazones and their metal complexes have been studied for their biological activities against bacteria, cancer cells and protozoa. Short-term in vitro treatment with one gold (III) complex (C3) and its salicyl-thiosemicarbazone ligand (C4) selectively inhibited proliferation of T. gondii. Transmission Electron Microscopy [...] Read more.
Thiosemicarbazones and their metal complexes have been studied for their biological activities against bacteria, cancer cells and protozoa. Short-term in vitro treatment with one gold (III) complex (C3) and its salicyl-thiosemicarbazone ligand (C4) selectively inhibited proliferation of T. gondii. Transmission Electron Microscopy (TEM) detected transient structural alterations in the parasitophorous vacuole membrane and the tachyzoite cytoplasm, but the mitochondrial membrane potential appeared unaffected by these compounds. Proteins potentially interacting with C3 and C4 were identified using differential affinity chromatography coupled with mass spectrometry (DAC-MS). Moreover, long-term in vitro treatment was performed to investigate parasitostatic or parasiticidal activity of the compounds. DAC-MS identified 50 ribosomal proteins binding both compounds, and continuous drug treatments for up to 6 days caused the loss of efficacy. Parasite tolerance to both compounds was, however, rapidly lost in their absence and regained shortly after re-exposure. Proteome analyses of six T. gondii ME49 clones adapted to C3 and C4 compared to the non-adapted wildtype revealed overexpression of ribosomal proteins, of two transmembrane proteins involved in exocytosis and of an alpha/beta hydrolase fold domain-containing protein. Results suggest that C3 and C4 may interfere with protein biosynthesis and that adaptation may be associated with the upregulated expression of tachyzoite transmembrane proteins and transporters, suggesting that the in vitro drug tolerance in T. gondii might be due to reversible, non-drug specific stress-responses mediated by phenotypic plasticity. Full article
(This article belongs to the Section Molecular Biology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Chemical structures of three gold (III) complexes and their thiosemicarbazone ligand. <b>C1</b>: salicylaldehyde thiosemicarbazone gold (III) chloride, <b>C2</b>: salicylaldehyde 4,4-Dimethyl-3-thiosemicarbazide gold (III) dichloride, <b>C3</b>: salicylaldehyde 4-phenylthiosemicarbazide gold (III) chloride, and <b>C4</b>: salicylaldehyde 4,4-dimethyl-3-thiosemicarbazide.</p>
Full article ">Figure 2
<p>TEM of non-treated <span class="html-italic">T. gondii</span> ME49 tachyzoites (<b>A</b>–<b>E</b>) and tachyzoites treated with the positive control drug DB745. (<b>A</b>) A tachyzoite undergoing endodyogeny with the two emerging daughter cells (dc) at 12 h post-infection, situated within a parasitophorous vacuole. The yellow boxed area in (<b>A</b>) is shown at higher magnification in (<b>B</b>). The vacuole is surrounded by the parasitophorous vacuole membrane (pvm). (<b>C</b>) A vacuole containing numerous tachyzoites at 24 h p.i., the yellow boxed area is shown at higher magnification in (<b>D</b>), revealing rhoptries. (rop), dense granule (dg), micronemes (mic), the mitochondrion, and the nucleus (nuc), all of them indicated with yellow arrows. (<b>E</b>) Proliferating tachyzoites still attached to a residual body (rb) within a parasitophorous vacuole at 48 h of culture. (<b>F</b>) tachyzoites treated with DB745 during 48 h, showing tachyzoites (yellow T in the figure) with severe structural alterations. Bars in (<b>A</b>) = 1 µm; (<b>B</b>) = 0.3 µm; (<b>C</b>) = 1.2 µm; (<b>D</b>) = 0.6 µm; (<b>E</b>) = 1.5 µm; (<b>F</b>) = 1.2 µm.</p>
Full article ">Figure 3
<p>TEM of <span class="html-italic">T. gondii</span> ME49 tachyzoites grown in HFF and treated with 0.5 µM C3 during 6–12 h (<b>A</b>–<b>C</b>), 24 h (<b>D</b>,<b>E</b>) and 48 h (<b>F</b>). A tachyzoite undergoing endodyogeny with an emerging daughter cell (dc) after 6 h of treatment is depicted in (<b>A</b>) and the respective apical part marked with an asterisk (*), including the conoid of the daughter cell (con) shown at higher magnification in (<b>B</b>). A large portion of the parasitophorous vacuole is surrounded by a multi-layered membrane (marked with small yellow arrows). Similar findings were obtained after 12 h (<b>C</b>). At 24 h, the multilayered membrane surrounding the parasitophorous vacuole was not visible anymore, but many parasites exhibited cytoplasmic alterations, leaving an empty space between the nuclear periphery and the cytoplasm, marked by asterisks (*) as shown in (<b>D</b>,<b>E</b>), and the matrix of the mitochondrion (mito) was partially dissolved. At 48 h, most tachyzoites had lost these ultrastructural alterations and had a normal appearance (<b>F</b>). Bars in (<b>A</b>) = 0.8 µm; (<b>B</b>) = 0.5 µm; (<b>C</b>) = 0.8 µm; (<b>D</b>) = 1 µm; (<b>E</b>) = 1 µm; (<b>F</b>) = 0.8 µm.</p>
Full article ">Figure 4
<p>TEM of <span class="html-italic">T. gondii</span> ME49 tachyzoites grown in HFF and treated with 0.5 µM of C4 during 6 h (<b>A</b>), 12 h (<b>B</b>,<b>C</b>), 24 h (<b>D</b>), and 48 h (<b>E</b>,<b>F</b>). The yellow boxed area in (<b>E</b>) is depicted at higher magnification in (<b>F</b>). A multi-layered membrane, indicated by small arrows surrounding the parasitophorous vacuole is evident after 6 h of treatment (<b>A</b>). After 12 h of treatment (<b>B</b>,<b>C</b>), ultrastructural alterations became visible in a large number of parasites as evidenced by the accumulation of membranous components within the vacuolar space (showed with thick yellow arrows) and increased numbers of cytoplasmic vacuoles (vac), and in some parasites, a separation of the nucleus from the surrounding cytoplasm was visible (asterisks *, see (<b>C</b>)). In addition, pronounced alterations of the mitochondrial matrix (mito) were noted after 24 h (<b>D</b>). At 48 h after initiation of drug treatment (<b>E</b>,<b>F</b>), parasites exhibited a largely normal structural appearance, with the exception of some parasites still maintaining free spaces between nuclear membrane and cytoplasm (asterisks *). Bars in (<b>A</b>) = 0.6 µm; (<b>B</b>) = 1 µm; (<b>C</b>) = 0.5 µm; (<b>D</b>) = 0.9 µm; (<b>E</b>) = 2.2 µm; (<b>F</b>) = 0.9 µm.</p>
Full article ">Figure 5
<p>Identification of <span class="html-italic">Toxoplasma</span> proteins that bind to C3 and/or C4 by affinity chromatography coupled mass spectrometry. (<b>A</b>): box and whisker plot describing protein intensity distributions as calculated by the iBAQ (intensity Based Absolute Quantification) algorithm across eluates from C3 (gold (III) complex), C4 (salicyl-TSC ligand), mock and tyrosine columns. (<b>B</b>): Venn diagram detailing the distribution of identified proteins (512) which were not binding to the mock column.</p>
Full article ">
17 pages, 4891 KiB  
Article
TMEM9B Regulates Endosomal ClC-3 and ClC-4 Transporters
by Margherita Festa, Maria Antonietta Coppola, Elena Angeli, Abraham Tettey-Matey, Alice Giusto, Irene Mazza, Elena Gatta, Raffaella Barbieri, Alessandra Picollo, Paola Gavazzo, Michael Pusch, Cristiana Picco and Francesca Sbrana
Life 2024, 14(8), 1034; https://doi.org/10.3390/life14081034 - 20 Aug 2024
Viewed by 3508
Abstract
The nine-member CLC gene family of Cl chloride-transporting membrane proteins is divided into plasma membrane-localized Cl channels and endo-/lysosomal Cl/H+ antiporters. Accessory proteins have been identified for ClC-K and ClC-2 channels and for the lysosomal ClC-7, but not [...] Read more.
The nine-member CLC gene family of Cl chloride-transporting membrane proteins is divided into plasma membrane-localized Cl channels and endo-/lysosomal Cl/H+ antiporters. Accessory proteins have been identified for ClC-K and ClC-2 channels and for the lysosomal ClC-7, but not the other CLCs. Here, we identified TMEM9 Domain Family Member B (TMEM9B), a single-span type I transmembrane protein of unknown function, to strongly interact with the neuronal endosomal ClC-3 and ClC-4 transporters. Co-expression of TMEM9B with ClC-3 or ClC-4 dramatically reduced transporter activity in Xenopus oocytes and transfected HEK cells. For ClC-3, TMEM9B also induced a slow component in the kinetics of the activation time course, suggesting direct interaction. Currents mediated by ClC-7 were hardly affected by TMEM9B, and ClC-1 currents were only slightly reduced, demonstrating specific interaction with ClC-3 and ClC-4. We obtained strong evidence for direct interaction by detecting significant Förster Resonance Energy Transfer (FRET), exploiting fluorescence lifetime microscopy-based (FLIM-FRET) techniques between TMEM9B and ClC-3 and ClC-4, but hardly any FRET with ClC-1 or ClC-7. The discovery of TMEM9B as a novel interaction partner of ClC-3 and ClC-4 might have important implications for the physiological role of these transporters in neuronal endosomal homeostasis and for a better understanding of the pathological mechanisms in CLCN3- and CLCN4-related pathological conditions. Full article
(This article belongs to the Special Issue Ion Channels and Neurological Disease: 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>TMEM9B as a putative interactor of ClC-3 and ClC-4. (<b>A</b>). Network of ClC-3 interactors, among which is TMEM9B, from the BioGrid 4.4 database [<a href="#B28-life-14-01034" class="html-bibr">28</a>]. The green circle highlights the TMEM9B entry. (<b>B</b>). Network of ClC-4 interactors, among which there is TMEM9B. (<b>C</b>). Network of TMEM9B interactors, among which are ClC-3, -4, and -5, highlighted by green circles. (<b>D</b>). TMEM9B hydrophobicity plot showing a hydrophobic signal peptide (sequence positions 1–32) and a glycosylated asparagine at sequence position 60 in the extracellular/luminal domain. (<b>E</b>). TMEM9B AlphaFold predicted structure, highlighting, in cyan, the hydrophobic region from sequence positions 99 to 144, and, in magenta, the glycosylated asparagine at position 60. The signal peptide was removed from the AlphaFold structure and the image was created with PyMol.</p>
Full article ">Figure 2
<p>Co-expression of ClC-4 and ClC-3 with TMEM9B in <span class="html-italic">Xenopus</span> oocytes. (<b>A</b>). Representative recordings of non-injected oocytes and oocytes injected with ClC-4 and with ClC-4 + TMEM9B evoked by the voltage-clamp protocol are shown on the right. (<b>B</b>). Averaged normalized I-V relationships of ClC-4 with and without TMEM9B. Currents are normalized as described in Methods. (<b>C</b>). Typical voltage clamp current traces of non-injected oocytes and oocytes injected with ClC-3 and co-injected with ClC-3 + TMEM9B in response to the stimulation protocol shown on the right. (<b>D</b>). Averaged normalized I-V currents collected for ClC-3 compared with ClC-3 co-injected with TMEM9B. Note that average currents from non-injected oocytes from the same batches are subtracted in the I-V plots and that, for some data points, error bars are smaller than symbol size.</p>
Full article ">Figure 3
<p>Co-expression of ClC-1 with TMEM9B in <span class="html-italic">Xenopus</span> oocytes. Typical current traces recorded from oocytes injected with ClC-1 alone and with TMEM9B (<b>A</b>). Stimulation protocol is shown as inset. (<b>B</b>) shows the normalized conductance of ClC-1 compared with ClC-1 with TMEM9B. For ClC-1, the slope conductance is the most robust parameter to quantify functional expression [<a href="#B20-life-14-01034" class="html-bibr">20</a>]. The error bar indicates SD (n = 3 injections). The star indicates <span class="html-italic">p</span> &lt; 0.05 (Student’s <span class="html-italic">t</span>-test).</p>
Full article ">Figure 4
<p>Co-expression with TMEM9B strongly reduced transport currents of ClC-4 in HEK cells. (<b>A</b>). Representative ionic currents elicited by the voltage-clamp protocol shown in the inset in control conditions (left trace) and in the presence of TMEM9B (right trace). (<b>B</b>). Average I-V plot shows a strong reduction of outward ClC-4 currents by TMEM9B (mean ± SEM). (<b>C</b>). Average current values at 200 mV (mean ± SD) (red bar, n = 11, I(200 mV) = 1.15 ± 0.45 nA; blue bar, n = 10, I(200 mV) = 0.14 ± 0.07 nA, <span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 5
<p>TMEM9B modulates the biophysical properties of ClC-3. (<b>A</b>). Representative ClC-3 currents elicited by the voltage-clamp protocol shown in the inset in control conditions (left trace) and in the presence of TMEM9B (right trace).0 (<b>B</b>). Representative recordings from a ClC-3 transfected cell (left trace) and from a cell co-transfected with TMEM9B using long (500 ms) pulses as indicated in the inset. (<b>C</b>). Average I-V plot in the absence (orange) and presence of TMEM9B (light blue, mean ± SEM). (<b>D</b>). Average current values at 200 mV (mean ± SD, orange bar, n = 13, I(200 mV) = 2.06 ± 0.97 nA; light blue bar, n = 13, I(200 mV) = 0.64 ± 0.37 nA; The four stars indicate <span class="html-italic">p</span> = 0.0002 (Student’s <span class="html-italic">t</span>-test). (<b>E</b>). Slowing of current activation by TMEM9B. Activation kinetics of cells co-transfected with TMEM9B was fitted with a double exponential function and values of the extracted time constants are shown as mean ± SD (purple bar: τ<sub>fast</sub> =21.0 ± 6.4 ms; pink bar: τ<sub>slow</sub> =120 ± 75 ms). No such slow kinetics were seen in cells transfected only with ClC-3.</p>
Full article ">Figure 6
<p>Co-expression with TMEM9B does not affect ClC-7 transport currents in HEK cells. (<b>A</b>). Representative ClC-7 currents elicited by the voltage-clamp protocol shown in the inset in control conditions (left trace) and in a cell co-transfected with TMEM9B (right trace). (<b>B</b>). Average I-V plot of ClC-7 transfected cells in the absence (green) and presence of TMEM9B (purple, mean ± SEM). (<b>C</b>). Average current values at 140 mV (mean ± SD, green bar, n = 5, I(200 mV) = 1.38 ± 0.58 nA; purple bar, n = 4, I(200 mV) = 1.19 ± 0.59 nA, <span class="html-italic">p</span> = 0.649).</p>
Full article ">Figure 7
<p>Subcellular localization of TMEM9B expressed alone. (<b>A</b>). Confocal images of HEK cells transfected with TMEM9B-GFP (green), stained with CellMask_DeepRed (magenta), merged image, and corresponding bright field image. The squared region is shown zoomed on the right. (<b>B</b>). Confocal images of cells transfected with TMEM9B-GFP (green) and stained with Lysotracker_DeepRed (magenta), merged image, and corresponding bright field image. The squared region is shown zoomed on the right. (<b>C</b>). Similar to B, but using mCherry tagged TMEM9B (red).</p>
Full article ">Figure 8
<p>Subcellular localization of TMEM9B expressed with CLC proteins. (<b>A</b>). Confocal images of HEK cells co-transfected with TMEM9B-mCherry (red) and ClC-4-GFP (green), stained with CellMask_DeepRed (cyan), the merged image, and the corresponding bright field image. The squared region is shown zoomed on the right. (<b>B</b>). Similar results with inverted tags, i.e., TMEM9B-GFP (green) and ClC-4-mCherry (red). (<b>C</b>). Similar results for cells co-transfected with TMEM9B-mCherry (red) and ClC-3-GFP (green). (<b>D</b>). Similar results for cells co-transfected with TMEM9B-mCherry (red) and ClC-1-GFP (green). (<b>E</b>). Similar results for cells co-transfected with TMEM9B-mCherry (red) and ClC-7-GFP (green).</p>
Full article ">Figure 9
<p>FLIM-FRET analysis of TMEM9B co-expressed with CLC proteins. (<b>A</b>). ClC-4-GFP phasor plot with corresponding lifetime value and representative image of a ClC-4-GFP transfected HEK cell. (<b>B</b>). ClC-4-GFP/TMEM9B-mCherry phasor plot with relative lifetime value and representative fluorescent confocal merged image of a ClC-4-GFP/TMEM9B-mCherry co-transfected HEK cell. Here, and in panels D, F, and H, the purple circle indicates the GS-coordinates of the unquenched donor. (<b>C</b>). ClC-3-GFP phasor plot with relative lifetime value and representative fluorescent confocal merged image of a ClC-3-GFP transfected HEK cell. (<b>D</b>). ClC-3-GFP/TMEM9B-mCherry phasor plot with relative lifetime value and representative fluorescent confocal merged image of a ClC-3-GFP/TMEM9B-mCherry co-transfected HEK cell. (<b>E</b>). ClC-1-GFP phasor plot with relative lifetime value and representative image of a ClC-1-GFP transfected HEK cell. (<b>F</b>). ClC-1-GFP/TMEM9B-mCherry phasor plot with relative lifetime value and representative image of a ClC-1-GFP/TMEM9B-mCherry co-transfected HEK cell. (<b>G</b>). ClC-7-GFP phasor plot with relative lifetime value and representative image of a ClC-7-GFP transfected HEK cell. (<b>H</b>). ClC-7-GFP/TMEM9B-mCherry phasor plot with relative lifetime value and representative image of a ClC-7-GFP/TMEM9B-mCherry co-transfected HEK cell. (<b>I</b>). FRET Efficiency analysis comparison with **** <span class="html-italic">p</span> &lt; 0.0001 compared to the other groups.</p>
Full article ">
Back to TopTop