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

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

Editor’s Choice Articles

Editor’s Choice articles are based on recommendations by the scientific editors of MDPI journals from around the world. Editors select a small number of articles recently published in the journal that they believe will be particularly interesting to readers, or important in the respective research area. The aim is to provide a snapshot of some of the most exciting work published in the various research areas of the journal.

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
32 pages, 407 KiB  
Article
The Medicinal Phage—Regulatory Roadmap for Phage Therapy under EU Pharmaceutical Legislation
by Timo Faltus
Viruses 2024, 16(3), 443; https://doi.org/10.3390/v16030443 - 12 Mar 2024
Cited by 2 | Viewed by 2268
Abstract
Bacteriophage therapy is a promising approach to treating bacterial infections. Research and development of bacteriophage therapy is intensifying due to the increase in antibiotic resistance and the faltering development of new antibiotics. Bacteriophage therapy uses bacteriophages (phages), i.e., prokaryotic viruses, to specifically target [...] Read more.
Bacteriophage therapy is a promising approach to treating bacterial infections. Research and development of bacteriophage therapy is intensifying due to the increase in antibiotic resistance and the faltering development of new antibiotics. Bacteriophage therapy uses bacteriophages (phages), i.e., prokaryotic viruses, to specifically target and kill pathogenic bacteria. The legal handling of this type of therapy raises several questions. These include whether phage therapeutics belong to a specially regulated class of medicinal products, and which legal framework should be followed with regard to the various technical ways in which phage therapeutics can be manufactured and administered. The article shows to which class of medicinal products phage therapeutics from wild type phages and from genetically modified (designer) phages do or do not belong. Furthermore, the article explains which legal framework is relevant for the manufacture and administration of phage therapeutics, which are manufactured in advance in a uniform, patient-independent manner, and for tailor-made patient-specific phage therapeutics. For the systematically coherent, successful translation of phage therapy, the article considers pharmaceutical law and related legal areas, such as genetic engineering law. Finally, the article shows how the planned legislative revisions of Directive 2001/83/EC and Regulation (EC) No 726/2004 may affect the legal future of phage therapy. Full article
(This article belongs to the Section Bacterial Viruses)
22 pages, 3968 KiB  
Article
Recommendations for Uniform Variant Calling of SARS-CoV-2 Genome Sequence across Bioinformatic Workflows
by Ryan Connor, Migun Shakya, David A. Yarmosh, Wolfgang Maier, Ross Martin, Rebecca Bradford, J. Rodney Brister, Patrick S. G. Chain, Courtney A. Copeland, Julia di Iulio, Bin Hu, Philip Ebert, Jonathan Gunti, Yumi Jin, Kenneth S. Katz, Andrey Kochergin, Tré LaRosa, Jiani Li, Po-E Li, Chien-Chi Lo, Sujatha Rashid, Evguenia S. Maiorova, Chunlin Xiao, Vadim Zalunin, Lisa Purcell and Kim D. Pruittadd Show full author list remove Hide full author list
Viruses 2024, 16(3), 430; https://doi.org/10.3390/v16030430 - 11 Mar 2024
Viewed by 13395
Abstract
Genomic sequencing of clinical samples to identify emerging variants of SARS-CoV-2 has been a key public health tool for curbing the spread of the virus. As a result, an unprecedented number of SARS-CoV-2 genomes were sequenced during the COVID-19 pandemic, which allowed for [...] Read more.
Genomic sequencing of clinical samples to identify emerging variants of SARS-CoV-2 has been a key public health tool for curbing the spread of the virus. As a result, an unprecedented number of SARS-CoV-2 genomes were sequenced during the COVID-19 pandemic, which allowed for rapid identification of genetic variants, enabling the timely design and testing of therapies and deployment of new vaccine formulations to combat the new variants. However, despite the technological advances of deep sequencing, the analysis of the raw sequence data generated globally is neither standardized nor consistent, leading to vastly disparate sequences that may impact identification of variants. Here, we show that for both Illumina and Oxford Nanopore sequencing platforms, downstream bioinformatic protocols used by industry, government, and academic groups resulted in different virus sequences from same sample. These bioinformatic workflows produced consensus genomes with differences in single nucleotide polymorphisms, inclusion and exclusion of insertions, and/or deletions, despite using the same raw sequence as input datasets. Here, we compared and characterized such discrepancies and propose a specific suite of parameters and protocols that should be adopted across the field. Consistent results from bioinformatic workflows are fundamental to SARS-CoV-2 and future pathogen surveillance efforts, including pandemic preparation, to allow for a data-driven and timely public health response. Full article
Show Figures

Figure 1

Figure 1
<p>Flow chart of the platforms from each participating organization’s workflows at the time of analysis. Shown are the schematics for (<b>A</b>) Illumina platform variant calling and (<b>B</b>) Oxford Nanopore Technologies (ONT) variant calling. For each sequencing platform, the main steps of variant calling are captured in each box, including: read retrieval, host removal, read trimming, alignment, variant calling, variant filtering, and variant normalization. For each step, the software used by each workflow is noted.</p>
Full article ">Figure 2
<p>The impact of host contamination removal and primer trimming. (<b>A</b>) The removal of host reads from RNAseq SARS-CoV-2 sequencing result, SRA run SRR12245095, reduced the potential for false-positive variant calls. In the top panel, additional mutations were present in aligned reads between positions 3049–3076 of NC_045512 when host reads were not removed. After excluding host reads (bottom panel), reads containing the mutations were no longer observed. (<b>B</b>) Allele frequencies of variants called after trimming primer sequences from aligned reads (corrected allele frequencies) are plotted against allele frequencies of the same variants called without primer trimming (uncorrected allele frequencies). Primer trimming increases the allele frequencies of most within-primer binding sites variants. Blue lines represent the allele-frequency thresholds used in this study to filter variant calls (allele frequency; AF ≥ 0.15) and to call consensus variants (AF ≥ 0.5).</p>
Full article ">Figure 3
<p>The effect of Alternate Allele Depth and Alternate Allele Frequency on variant calling agreement across workflows and platforms. For each panel, calls made by all but one workflow (<b>A</b>–<b>D</b>) or both platforms (<b>E</b>–<b>H</b>) were considered true-positives, while calls made by only a single workflow (or technology) were considered false-positives, thus the ROC AUCs cannot be directly compared between groups. For the right panels, points represent an Allele Frequency (AF) cut-off of 1 at the lower-leftmost point, and the cut-off decreases by 0.1 along the length of the line. For the left panels, the points represent a minimum Alternate Allele Depth (AltDP) going from 4,000 at the lower-left most point to 10 along each line. (<b>A</b>,<b>B</b>) Impact of AltDP and AF, respectively, on Illumina workflow accuracy and specificity across workflows. (<b>C</b>,<b>D</b>) Impact of AltDP and AF, respectively, on Illumina workflow accuracy and specificity across platforms. (<b>E</b>,<b>F</b>) Impact of AltDP and AF, respectively, on ONT workflow accuracy and specificity across workflows. (<b>G</b>,<b>H</b>) Impact of AltDP and AF, respectively, on ONT workflow accuracy and specificity across platforms.</p>
Full article ">Figure 4
<p>Agreement across workflows with and without recommended parameters. (<b>A</b>–<b>D</b>) Agreement across workflows, without recommended parameters. (<b>E</b>–<b>H</b>) Agreement across workflows, with recommended parameters. (<b>A</b>,<b>E</b>) Agreement on Illumina SNP calls. (<b>B</b>,<b>F</b>) Agreement on Illumina InDel calls. (<b>C</b>,<b>G</b>) Agreement on Oxford Nanopore (ONT) SNP calls. (<b>D</b>,<b>H</b>) Agreement on ONT InDel Calls. For each figure, the bars indicate the number of variants called by the groups, indicated by filled circles below, across the whole dataset.</p>
Full article ">Figure 5
<p>Application of recommended parameters results in increased agreement across platforms. Graphical representation of the agreement between platforms without the application of recommended parameters of SNP (<b>A</b>) and InDel (<b>B</b>) calls. (<b>C</b>) (SNP) and (<b>D</b>) (InDel) represent the agreement between platforms after the application of the recommended parameters. For each figure, only those samples for which both Illumina and ONT platform data had at least one variant call that passed all of the filters were considered. The total height is normalized to the total number of calls made by each workflow, with light blue portion indicating calls made on both platforms for a given sample, medium blue indicating calls made only for Illumina data, and dark blue indicating calls made only for ONT data.</p>
Full article ">Figure 6
<p>Variant calling workflow recommendations. Outline of the recommendations for each step in a variant calling workflow, from read cleanup to variant filtering, are illustrated. Additionally, the benefit of implementing the recommendations at each step are noted.</p>
Full article ">
16 pages, 5599 KiB  
Article
Discovery of a Novel Antiviral Effect of the Restriction Factor SPOC1 against Human Cytomegalovirus
by Anna K. Kuderna, Anna Reichel, Julia Tillmanns, Maja Class, Myriam Scherer and Thomas Stamminger
Viruses 2024, 16(3), 363; https://doi.org/10.3390/v16030363 - 27 Feb 2024
Viewed by 1043
Abstract
The chromatin-remodeler SPOC1 (PHF13) is a transcriptional co-regulator and has been identified as a restriction factor against various viruses, including human cytomegalovirus (HCMV). For HCMV, SPOC1 was shown to block the onset of immediate-early (IE) gene expression under low multiplicities of infection (MOI). [...] Read more.
The chromatin-remodeler SPOC1 (PHF13) is a transcriptional co-regulator and has been identified as a restriction factor against various viruses, including human cytomegalovirus (HCMV). For HCMV, SPOC1 was shown to block the onset of immediate-early (IE) gene expression under low multiplicities of infection (MOI). Here, we demonstrate that SPOC1-mediated restriction of IE expression is neutralized by increasing viral titers. Interestingly, our study reveals that SPOC1 exerts an additional antiviral function beyond the IE phase of HCMV replication. Expression of SPOC1 under conditions of high MOI resulted in severely impaired viral DNA replication and viral particle release, which may be attributed to inefficient viral transcription. With the use of click chemistry, the localization of viral DNA was investigated at late time points after infection. Intriguingly, we detected a co-localization of SPOC1, RNA polymerase II S5P and polycomb repressor complex 2 (PRC2) components in close proximity to viral DNA in areas that are hypothesized to harbor viral transcription sites. We further identified the N-terminal domain of SPOC1 to be responsible for interaction with EZH2, a subunit of the PRC2 complex. With this study, we report a novel and potent antiviral function of SPOC1 against HCMV that is efficient even with unrestricted IE gene expression. Full article
(This article belongs to the Special Issue Molecular Biology of Human Cytomegalovirus)
Show Figures

Figure 1

Figure 1
<p>Increasing HCMV doses are able to antagonize SPOC1-mediated repression of IE1 and IE2 expression. (<b>A</b>) The 24 hpi lysates of HCMV-infected control fibroblasts (HFF/Ctrl) and fibroblasts expressing SPOC1 (HFF/SPOC1) infected with HCMV TB40/E at MOIs of 0.01 (lanes 1 and 2), 1 (lanes 3 and 4) and 3 (lanes 5 and 6) were investigated by Western blotting. Expression levels of viral immediate-early proteins IE1 and IE2, β-actin and SPOC1 were analyzed. (<b>B</b>) Quantification of IE1 and IE2 signal intensities normalized to β-actin levels in HFF/SPOC1 relative to normalized IE1 and IE2 levels in HFF/Ctrl of three independent experiments. Statistical analysis was performed using Student’s <span class="html-italic">t</span>-test (one sample, two-tailed); **** <span class="html-italic">p</span> &lt; 0.0001, * <span class="html-italic">p</span> &lt; 0.05, <span class="html-italic">n.s.</span> = not significant.</p>
Full article ">Figure 2
<p>SPOC1 expression negatively affects HCMV DNA replication and viral particle release. HFF/Ctrl and HFF/SPOC1 were infected in triplicate with HCMV strain TB40/E at an MOI of 1 or 3. (<b>A</b>) At 96 hpi, supernatants were analyzed for viral genome equivalents via qPCR. (<b>B</b>) At 96 hpi, intracellular DNA was isolated and HCMV genome equivalents were quantified via qPCR and normalized to albumin copy numbers. One out of three independent experiments is shown. Statistical analysis was performed using Student’s <span class="html-italic">t</span>-test (unpaired, two-tailed); *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001. (<b>C</b>) Experimental set-up: Doxycycline (Dox)-inducible HFF/SPOC1 cells were either treated with Dox 24 h prior to or 24 h post infection with AD169 at MOI 0.1. (<b>D</b>) Dox-inducible HFF/SPOC1 was infected with AD169, at MOI 0.1, in triplicate. The infected cells were either left untreated or treated with Dox 24 h prior to or 24 h post infection (<b>C</b>). At 96 hpi, the supernatant was harvested and analyzed for viral genome equivalents via qPCR. One out of two experiments is shown. Statistical analysis was performed utilizing the one sample <span class="html-italic">t</span>-test; * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 3
<p>SPOC1 leads to lower expression levels of viral early and late proteins. (<b>A</b>) Lysates of mock-infected (m) or TB40/E- (MOI of 3) or AD169-infected (MOI of 3 and 1) HFF/Ctrl and HFF/SPOC1 cells were analyzed at 24 to 72 hpi by separation on a 10 % polyacrylamide gel followed by Western blot detection of indicated proteins. Expression kinetics of viral immediate-early protein IE1, viral early protein pUL44 and viral late proteins pp28 and MCP were investigated. Asterisks (*) highlight the protein bands that are attenuated upon SPOC1 expression. (<b>B</b>) Quantification of IE1 levels in HFF/SPOC1 normalized to β-actin are depicted as fold change of the normalized IE1 level of HFF/Ctrl (indicated by dashed line at y = 1). (<b>C</b>) Quantification of early and late viral proteins of HFF/SPOC1 normalized to β-actin are depicted as fold change of the regarding normalized protein levels of HFF/Ctrl (indicated by dashed line at y = 1).</p>
Full article ">Figure 4
<p>Impaired transcription of viral early and late genes in HFF/SPOC1 cells at late times in the replicative cycle. HFF/Ctrl and HFF/SPOC1 cells were infected with AD169 at an MOI of 1. (<b>A</b>) The 24 hpi IE1 and IE2 protein levels were analyzed via SDS PAGE and Western blotting. The relative intensity values were normalized to β-actin. Quantification of three independent experiments is shown. Statistical analysis was performed using Student’s <span class="html-italic">t</span>-test (one sample, two-tailed); <span class="html-italic">n.s.</span> = not significant. (<b>B</b>) IE1, IE2 and US3 transcript levels normalized to GAPDH were evaluated at 8 hpi using qPCR. Shown are the mean values of triplicates of one out of two experiments. Statistical analysis was performed using Student’s <span class="html-italic">t</span>-test (one sample, two-tailed). (<b>C</b>) At 48 and 72 hpi, total cellular RNA was isolated, cDNA was synthesized and viral mRNA levels of two immediate-early, early and late genes were quantified via qPCR, respectively. Shown are the mean values of triplicates of one out of three experiments. Statistical analysis was performed with Student’s <span class="html-italic">t</span>-test (unpaired, two-tailed); ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 5
<p>SPOC1 localization during the time course of HCMV infection. (<b>A</b>) HFF/mCherry-SPOC1 cells were infected with AD169 at MOI 1 and fixed at the indicated time points. An antibody against pUL44 was used in combination with the secondary antibody Alexa-488. DAPI staining was used to visualize the nucleus. (<b>B</b>) At 72 hpi, AD169-infected HFF/SPOC1 (MOI of 1) was treated with F-Ara-EdU prior to fixation at 96 hpi. Samples were stained for SPOC1 (secondary antibody: Alexa-488) as well as with an antibody against pUL44 in combination with the Alexa-647 antibody. Click chemistry was performed to visualize viral DNA (modified from [<a href="#B15-viruses-16-00363" class="html-bibr">15</a>]).</p>
Full article ">Figure 6
<p>SPOC1 co-localizes with PRC2 components close to viral DNA. HFF/SPOC1 was infected with AD169 at an MOI of 1. At 96 hpi, the cells were fixed and treated with antibodies directed against SPOC1, pUL44 and either (<b>A</b>) EZH2 or (<b>B</b>) SUZ12. As secondary antibodies, Alexa-488 (SPOC1) and a combination of either mouse or rabbit Alexa-555 and mouse or rabbit Alexa-647 were used. DAPI signals visualize the nucleus (modified from [<a href="#B15-viruses-16-00363" class="html-bibr">15</a>]). (<b>C</b>,<b>D</b>) At 72 hpi, EdC was added to SPOC1-expressing cells prior to fixation at 96 hpi. The same antibodies against EZH2 (<b>C</b>) or SUZ12 (<b>D</b>) were used as for (<b>A</b>) and (<b>B</b>); viral DNA was visualized by click chemistry. Nuclei were visualized by DAPI staining. Merge images were created from the two images above.</p>
Full article ">Figure 7
<p>Interaction between SPOC1 deletion mutants and EZH2. (<b>A</b>) Schematic representation of the generated FLAG-SPOC1 deletion mutants with indicated amino acid sequence. (<b>B</b>) HEK293T cells were co-transfected with an empty control plasmid or FLAG-SPOC1 deletion mutants together with an EZH2-expressing plasmid. FLAG-SPOC1 was precipitated, and lysate controls as well as immuno-precipitated (IP) samples were analyzed via Western blotting. SPOC1 was visualized using an anti-FLAG antibody. CoIP = Co-Immunoprecipitaton.</p>
Full article ">Figure 8
<p>Co-localization of SPOC1 with RNA pol II S5P. HFF/SPOC1 was infected with AD169 (MOI 1) and fixed at 96 hpi. Indirect immunostaining was performed using antibodies against SPOC1 and (<b>A</b>) 4H8 antibody detecting specifically Ser5 phosphorylated RNA pol II (modified from [<a href="#B15-viruses-16-00363" class="html-bibr">15</a>]) or (<b>B</b>) 8WG16 antibody to mark general RNA Pol II localization. DAPI was used to stain cell nuclei.</p>
Full article ">
12 pages, 8282 KiB  
Commentary
Viroids, Satellite RNAs and Prions: Folding of Nucleic Acids and Misfolding of Proteins
by Gerhard Steger, Detlev Riesner and Stanley B. Prusiner
Viruses 2024, 16(3), 360; https://doi.org/10.3390/v16030360 - 26 Feb 2024
Viewed by 2556
Abstract
Theodor (“Ted”) Otto Diener (* 28 February 1921 in Zürich, Switzerland; † 28 March 2023 in Beltsville, MD, USA) pioneered research on viroids while working at the Plant Virology Laboratory, Agricultural Research Service, USDA, in Beltsville. He coined the name viroid and defined [...] Read more.
Theodor (“Ted”) Otto Diener (* 28 February 1921 in Zürich, Switzerland; † 28 March 2023 in Beltsville, MD, USA) pioneered research on viroids while working at the Plant Virology Laboratory, Agricultural Research Service, USDA, in Beltsville. He coined the name viroid and defined viroids’ important features like the infectivity of naked single-stranded RNA without protein-coding capacity. During scientific meetings in the 1970s and 1980s, viroids were often discussed at conferences together with other “subviral pathogens”. This term includes what are now called satellite RNAs and prions. Satellite RNAs depend on a helper virus and have linear or, in the case of virusoids, circular RNA genomes. Prions, proteinaceous infectious particles, are the agents of scrapie, kuru and some other diseases. Many satellite RNAs, like viroids, are non-coding and exert their function by thermodynamically or kinetically controlled folding, while prions are solely host-encoded proteins that cause disease by misfolding, aggregation and transmission of their conformations into infectious prion isoforms. In this memorial, we will recall the work of Ted Diener on subviral pathogens. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Timeline of research development of viroids (green text) [<a href="#B4-viruses-16-00360" class="html-bibr">4</a>,<a href="#B13-viruses-16-00360" class="html-bibr">13</a>,<a href="#B14-viruses-16-00360" class="html-bibr">14</a>,<a href="#B15-viruses-16-00360" class="html-bibr">15</a>,<a href="#B16-viruses-16-00360" class="html-bibr">16</a>,<a href="#B17-viruses-16-00360" class="html-bibr">17</a>,<a href="#B18-viruses-16-00360" class="html-bibr">18</a>,<a href="#B19-viruses-16-00360" class="html-bibr">19</a>,<a href="#B20-viruses-16-00360" class="html-bibr">20</a>,<a href="#B21-viruses-16-00360" class="html-bibr">21</a>,<a href="#B22-viruses-16-00360" class="html-bibr">22</a>,<a href="#B23-viruses-16-00360" class="html-bibr">23</a>,<a href="#B24-viruses-16-00360" class="html-bibr">24</a>,<a href="#B25-viruses-16-00360" class="html-bibr">25</a>,<a href="#B26-viruses-16-00360" class="html-bibr">26</a>,<a href="#B27-viruses-16-00360" class="html-bibr">27</a>,<a href="#B28-viruses-16-00360" class="html-bibr">28</a>,<a href="#B29-viruses-16-00360" class="html-bibr">29</a>,<a href="#B30-viruses-16-00360" class="html-bibr">30</a>,<a href="#B31-viruses-16-00360" class="html-bibr">31</a>,<a href="#B32-viruses-16-00360" class="html-bibr">32</a>,<a href="#B33-viruses-16-00360" class="html-bibr">33</a>,<a href="#B34-viruses-16-00360" class="html-bibr">34</a>,<a href="#B35-viruses-16-00360" class="html-bibr">35</a>,<a href="#B36-viruses-16-00360" class="html-bibr">36</a>,<a href="#B37-viruses-16-00360" class="html-bibr">37</a>,<a href="#B38-viruses-16-00360" class="html-bibr">38</a>,<a href="#B39-viruses-16-00360" class="html-bibr">39</a>,<a href="#B40-viruses-16-00360" class="html-bibr">40</a>,<a href="#B41-viruses-16-00360" class="html-bibr">41</a>,<a href="#B42-viruses-16-00360" class="html-bibr">42</a>,<a href="#B43-viruses-16-00360" class="html-bibr">43</a>,<a href="#B44-viruses-16-00360" class="html-bibr">44</a>], satellite RNAs (orange text) [<a href="#B11-viruses-16-00360" class="html-bibr">11</a>,<a href="#B12-viruses-16-00360" class="html-bibr">12</a>,<a href="#B45-viruses-16-00360" class="html-bibr">45</a>,<a href="#B46-viruses-16-00360" class="html-bibr">46</a>,<a href="#B47-viruses-16-00360" class="html-bibr">47</a>,<a href="#B48-viruses-16-00360" class="html-bibr">48</a>,<a href="#B49-viruses-16-00360" class="html-bibr">49</a>], prions (red text) [<a href="#B7-viruses-16-00360" class="html-bibr">7</a>,<a href="#B50-viruses-16-00360" class="html-bibr">50</a>,<a href="#B51-viruses-16-00360" class="html-bibr">51</a>,<a href="#B52-viruses-16-00360" class="html-bibr">52</a>,<a href="#B53-viruses-16-00360" class="html-bibr">53</a>,<a href="#B54-viruses-16-00360" class="html-bibr">54</a>,<a href="#B55-viruses-16-00360" class="html-bibr">55</a>,<a href="#B56-viruses-16-00360" class="html-bibr">56</a>,<a href="#B57-viruses-16-00360" class="html-bibr">57</a>,<a href="#B58-viruses-16-00360" class="html-bibr">58</a>,<a href="#B59-viruses-16-00360" class="html-bibr">59</a>,<a href="#B60-viruses-16-00360" class="html-bibr">60</a>,<a href="#B61-viruses-16-00360" class="html-bibr">61</a>,<a href="#B62-viruses-16-00360" class="html-bibr">62</a>,<a href="#B63-viruses-16-00360" class="html-bibr">63</a>], and a few general hallmarks (black text) [<a href="#B64-viruses-16-00360" class="html-bibr">64</a>].</p>
Full article ">Figure 2
<p>At the 7th International Conference of Virology in Edmonton, Canada, 1987. From left to right: Robert A. Owens, Detlev Riesner, Theodor O. Diener, Heinz Ludwig Sänger.</p>
Full article ">Figure 3
<p>A poster and its visitor at the 10th International Congress of Biochemistry, Hamburg, Germany (1976). Prusiner in front of the poster sketched from memory by Riesner.</p>
Full article ">
12 pages, 969 KiB  
Article
Bivalent VSV Vectors Mediate Rapid and Potent Protection from Andes Virus Challenge in Hamsters
by Joshua Marceau, David Safronetz, Cynthia Martellaro, Andrea Marzi, Kyle Rosenke and Heinz Feldmann
Viruses 2024, 16(2), 279; https://doi.org/10.3390/v16020279 - 11 Feb 2024
Viewed by 985
Abstract
Orthohantaviruses may cause hemorrhagic fever with renal syndrome or hantavirus cardiopulmonary syndrome. Andes virus (ANDV) is the only orthohantavirus associated with human–human transmission. Therefore, emergency vaccination would be a valuable public health measure to combat ANDV-derived infection clusters. Here, we utilized a promising [...] Read more.
Orthohantaviruses may cause hemorrhagic fever with renal syndrome or hantavirus cardiopulmonary syndrome. Andes virus (ANDV) is the only orthohantavirus associated with human–human transmission. Therefore, emergency vaccination would be a valuable public health measure to combat ANDV-derived infection clusters. Here, we utilized a promising vesicular stomatitis virus (VSV)-based vaccine to advance the approach for emergency applications. We compared monovalent and bivalent VSV vectors containing the Ebola virus (EBOV), glycoprotein (GP), and ANDV glycoprotein precursor (GPC) for protective efficacy in pre-, peri- and post-exposure immunization by the intraperitoneal and intranasal routes. Inclusion of the EBOV GP was based on its favorable immune cell targeting and the strong innate responses elicited by the VSV-EBOV vaccine. Our data indicates no difference of ANDV GPC expressing VSV vectors in pre-exposure immunization independent of route, but a potential benefit of the bivalent VSVs following peri- and post-exposure intraperitoneal vaccination. Full article
Show Figures

Figure 1

Figure 1
<p>Generation and characterization of the VSV vectors. (<b>A</b>) Schematic design of VSV vectors. VSV vectors were generated through reverse genetics, recovered, titrated and sequence confirmed. The following VSV vectors were used: VSVwt, VSV–ANDV (expressing ANDV GPC), VSV–EBOV (expressing EBOV GP), VSV–ANDV–EBOV (expressing ANDV GPC and EBOV GP), and VSV–EBOV–ANDV (expressing EBOV GP and ANDV GPC). (<b>B</b>) In vitro attenuation. BHK-21 cells were infected with the different VSV vectors (MOI of 0.001), and supernatants were harvested at the indicated time points. Infectivity was determined by a TCID<sub>50</sub> assay. (<b>C</b>) In vivo attenuation. Syrian hamsters (<span class="html-italic">n</span> = 8) were infected with the different VSV vectors and monitored for clinical signs. The graph shows the survival curves. A log-rank test was used to determine significance (<span class="html-italic">p</span> &lt; 0.01 = **). ANDV = Andes virus; EBOV = Ebola virus; MOI = multiplicity of infection; TCID50 = median tissue culture infectious dose.</p>
Full article ">Figure 2
<p>Viral load in lung tissue and blood after pre- and post-challenge immunization. (<b>A</b>) Experimental design. Hamsters (<span class="html-italic">n</span> = 9 per group) were immunized IP or IN with a single dose of 1 × 10<sup>5</sup> PFU pre (D-28 and D-3) or post (D + 1) lethal ANDV challenge (200 FFU/animal). Lung tissue and blood were taken from 3 animals of each group on 8 DPC and analyzed for viral replication using RT-PCR. (<b>B</b>) D-28, IP immunization; (<b>C</b>) D-28, IN immunization; (<b>D</b>) D-3, IP immunization; (<b>E</b>) D-3, IN immunization; (<b>F</b>) D + 1, IP immunization; (<b>G</b>) D + 1, IN immunization. Significance between groups was determined using a two-way ANOVA with Tukey’s multiple comparisons test (<span class="html-italic">p</span> &lt; 0.0001 = ****, <span class="html-italic">p</span> &lt; 0.001 = ***, <span class="html-italic">p</span> &lt; 0.01 = **, <span class="html-italic">p</span> &lt; 0.05 = *). ANDV = Andes virus; D = day; EBOV = Ebola virus; IN = intranasal; IP = intraperitoneal; ‘-’ = pre challenge; ‘+’ = post challenge.</p>
Full article ">
19 pages, 2290 KiB  
Technical Note
Robust Approaches to the Quantitative Analysis of Genome Formula Variation in Multipartite and Segmented Viruses
by Marcelle L. Johnson and Mark P. Zwart
Viruses 2024, 16(2), 270; https://doi.org/10.3390/v16020270 - 8 Feb 2024
Viewed by 948
Abstract
When viruses have segmented genomes, the set of frequencies describing the abundance of segments is called the genome formula. The genome formula is often unbalanced and highly variable for both segmented and multipartite viruses. A growing number of studies are quantifying the genome [...] Read more.
When viruses have segmented genomes, the set of frequencies describing the abundance of segments is called the genome formula. The genome formula is often unbalanced and highly variable for both segmented and multipartite viruses. A growing number of studies are quantifying the genome formula to measure its effects on infection and to consider its ecological and evolutionary implications. Different approaches have been reported for analyzing genome formula data, including qualitative description, applying standard statistical tests such as ANOVA, and customized analyses. However, these approaches have different shortcomings, and test assumptions are often unmet, potentially leading to erroneous conclusions. Here, we address these challenges, leading to a threefold contribution. First, we propose a simple metric for analyzing genome formula variation: the genome formula distance. We describe the properties of this metric and provide a framework for understanding metric values. Second, we explain how this metric can be applied for different purposes, including testing for genome-formula differences and comparing observations to a reference genome formula value. Third, we re-analyze published data to illustrate the applications and weigh the evidence for previous conclusions. Our re-analysis of published datasets confirms many previous results but also provides evidence that the genome formula can be carried over from the inoculum to the virus population in a host. The simple procedures we propose contribute to the robust and accessible analysis of genome-formula data. Full article
(This article belongs to the Special Issue Plant Virus Epidemiology and Control 2023)
Show Figures

Figure 1

Figure 1
<p>We provide a schematic illustration of the variation in the distribution of genome segments (nucleic acid molecules) over virus particles. A legend is given on the far right. In each case shown, we assume the virus genome consists of the two identical coding genome regions, identified by blue and red fills, forming one or two segments. (<b>a</b>) Monopartite viruses have a single genome segment. Note that the two genome regions form a single molecule in the illustration. (<b>b</b>) Segmented viruses have multiple genome segments: two genome segments in this example. These viruses package a full complement of genome segments into each virus particle. (<b>c</b>) A multipartite virus with two genome segments is shown. Each segment is packaged individually into a virus particle. Infection will depend on the transmission of multiple virus particles, as both a blue and a red segment are needed. (<b>d</b>) A segmented virus with non-selective packaging is shown. The illustration is a hypothetical distribution based only on the observation that for some segmented viruses, many virus particles have an incomplete set of genome segments [<a href="#B5-viruses-16-00270" class="html-bibr">5</a>,<a href="#B19-viruses-16-00270" class="html-bibr">19</a>]. This organization is included to highlight that many distributions of genome segments over virus particles are possible, and that the genome formula of segmented viruses does not have to be balanced (i.e., not 1:1 ratio of genome segments).</p>
Full article ">Figure 2
<p>Here, we illustrate the genome formula distance metric (<b>top</b> panels, green lines) and its maximum possible distance for different numbers of genome segments (<b>bottom</b> panels, purple arrows). Figure axes are genome segment frequencies (<span class="html-italic">f</span>) for 2 (panels (<b>a</b>,<b>b</b>)), 3 (panels (<b>b</b>,<b>c</b>,<b>f</b>,<b>g</b>)), or 4 genome segments (panels (<b>d</b>,<b>h</b>)). (<b>a</b>) For a bipartite virus, we illustrate two possible genome formula values with green points and the distance between them with a line. Note that for the bipartite virus, all possible genome formula values fall on the dotted line connecting (1,0) and (0,1). (<b>b</b>) For a tripartite virus, we illustrate two possible genome formula values in three-dimensional genome formula space. As the sum of relative frequencies is 1, all possible genome formula values fall in the triangular plane illustrated by the dotted lines and light blue shading. (<b>c</b>) As all values fall in the same plane in panel b, genome formula values for a tri-segmented virus are often illustrated in only this plane, resulting in a ternary plot. (<b>d</b>) Two genome formula values and their distance are illustrated for a tetrapartite virus in a quarternary plot. All values in the tetrahedron represent possible genome formula values, as indicated by the light blue shading. (<b>e</b>) The maximum possible genome formula distance for a bipartite virus is simply the line connecting the points (1,0) and (0,1). (<b>f</b>) For the tripartite virus, the longest possible distance in the genome formula space is attained along its borders, resulting in an identical maximum genome formula distance to the bipartite virus. The light blue shading indicates the possible space for genome formula values. (<b>g</b>) The outcome described in panel f is clearer in the ternary plot of the genome formula space. (<b>h</b>) For a tetrapartite virus, there is no distance between two points in the genome formula space that is longer than the maximum distance for the bipartite and tripartite viruses. This maximum distance occurs at the edges of the genome formula space, as indicated by the light blue shading, connecting the vertices, which represent the presence of a single segment. To keep the panel clear, we only illustrate this for one edge for a tetrapartite virus, although there are six such edges.</p>
Full article ">Figure 3
<p>The effects of the number of segments and bottleneck size on the predicted genome formula distance are illustrated. The <span class="html-italic">x</span>-axis indicates the number of virus genome segments, whereas the <span class="html-italic">y</span>-axis indicates the log-transformed number of infection founders (<span class="html-italic">λ</span>). For all combinations of these values, we predicted the mean genome formula distance <math display="inline"><semantics> <mrow> <msub> <mrow> <mover accent="true"> <mrow> <mi>D</mi> </mrow> <mo>¯</mo> </mover> </mrow> <mrow> <mi>a</mi> <mo>,</mo> <mi>b</mi> </mrow> </msub> </mrow> </semantics></math>, a value indicated by the heat according to the legend on the far right. We used these simulation results to determine the highest value of <math display="inline"><semantics> <mrow> <msub> <mrow> <mover accent="true"> <mrow> <mi>D</mi> </mrow> <mo>¯</mo> </mover> </mrow> <mrow> <mi>a</mi> <mo>,</mo> <mi>b</mi> </mrow> </msub> </mrow> </semantics></math> for each number of genome segments, a value we term <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mover accent="true"> <mrow> <mi>D</mi> </mrow> <mo>¯</mo> </mover> </mrow> <mrow> <mi>a</mi> <mo>,</mo> <mi>b</mi> </mrow> <mrow> <mi>d</mi> <mi>r</mi> <mi>i</mi> <mi>f</mi> <mi>t</mi> </mrow> </msubsup> </mrow> </semantics></math>. Note that the highest mean distance values occur at intermediate values of <span class="html-italic">λ</span>, as well as being associated with higher values of <span class="html-italic">λ</span> as the number of segments is increased.</p>
Full article ">Figure 4
<p>Resampling approach to testing for an effect of inoculum on the genome formula measured in the inoculated leaf. The blue bars in the histogram indicate the frequency of predicted mean genome formula distance for 10<sup>4</sup> resampled datasets, in which observations in the inoculated leaf were randomly assigned to an inoculum. The red line indicates the genome formula distance for the actual data.</p>
Full article ">Figure A1
<p>Resampling approach for testing for an effect of inoculum on the AMV genome formula measured in different tissues. The blue bars in the histogram indicate the frequency of predicted mean genome formula distance for 10<sup>4</sup> resampled datasets, in which observations in the inoculated leaf were randomly assigned to an inoculum. The red line indicates the genome formula distance for the actual data, which in all cases falls well within the 99% confidence interval of the distribution predicted by resampling (see <a href="#viruses-16-00270-t004" class="html-table">Table 4</a>). (<b>a</b>) Results for the middle leaf of the plant are shown. (<b>b</b>) Results for the upper leaf are shown. (<b>c</b>) Results for the rest of the plant tissues are shown.</p>
Full article ">
18 pages, 3792 KiB  
Article
Low gH/gL (Sub)Species-Specific Antibody Levels Indicate Elephants at Risk of Fatal Elephant Endotheliotropic Herpesvirus Hemorrhagic Disease
by Tabitha E. Hoornweg, Willem Schaftenaar, Victor P. M. G. Rutten and Cornelis A. M. de Haan
Viruses 2024, 16(2), 268; https://doi.org/10.3390/v16020268 - 8 Feb 2024
Viewed by 1445
Abstract
Elephant endotheliotropic herpesviruses (EEHVs), of which eleven (sub)species are currently distinguished, infect either Asian (Elephas maximus) or African elephants (Loxodonta species). While all adult elephants are latently infected with at least one EEHV (sub)species, young elephants, specifically those with low [...] Read more.
Elephant endotheliotropic herpesviruses (EEHVs), of which eleven (sub)species are currently distinguished, infect either Asian (Elephas maximus) or African elephants (Loxodonta species). While all adult elephants are latently infected with at least one EEHV (sub)species, young elephants, specifically those with low to non-detectable EEHV-specific antibody levels, may develop fatal hemorrhagic disease (EEHV-HD) upon infection. However, animals with high antibody levels against EEHV(1A) gB, an immunodominant antigen recognized by antibodies elicited against multiple (sub)species, may also occasionally succumb to EEHV-HD. To better define which animals are at risk of EEHV-HD, gB and gH/gL ELISAs were developed for each of the Asian elephant EEHV subspecies and assessed using 396 sera from 164 Asian elephants from European zoos. Antibody levels measured against gB of different (sub)species correlated strongly with one another, suggesting high cross-reactivity. Antibody levels against gH/gL of different subspecies were far less correlated and allowed differentiation between these (sub)species. Importantly, while high gB-specific antibody levels were detected in the sera of several EEHV-HD fatalities, all fatalities (n = 23) had low antibody levels against gH/gL of the subspecies causing disease. Overall, our data indicate that (sub)species-specific gH/gL ELISAs can be used to identify animals at risk of EEHV-HD when infected with a particular EEHV (sub)species. Full article
(This article belongs to the Special Issue Animal Herpesvirus)
Show Figures

Figure 1

Figure 1
<p>EEHV gB-specific antibody levels in young Asian elephants that did not develop EEHV-HD as compared to EEHV-HD fatalities. EEHV-specific antibody levels measured using the multiple-EEHV-species (EEHV1A) gB ELISA for a cross-sectional cohort of (<b>A</b>) 102 serum samples from 49 individual Asian elephants below the age of 10 that had not (yet) developed EEHV-HD and (<b>B</b>) 56 serum samples from 23 individual EEHV-HD fatalities. All samples were tested at 1:100 dilution, and ΔOD values were obtained by subtraction of serum-specific background signals from gB-specific signals. Values were normalized as described in Materials and Methods. Obtained ΔOD values are plotted according to age at sampling. Individual animals are distinguished by different colors, and longitudinal sera are connected by lines. (<b>C</b>) ΔOD values depicted in panels A and B grouped per category. Individual values and mean ± standard deviation (SD) are shown. Statistical significance was tested by Mann–Whitney test using GraphPad Prism: **** indicates <span class="html-italic">p</span> &lt; 0.0001. OD = optical density.</p>
Full article ">Figure 2
<p>Maximum likelihood trees inferred from gB and gH/gL amino acid sequences of different EEHV (sub)species. Full-length amino acid sequences of gB (<b>A</b>) and gH and gL (<b>B</b>) were retrieved from GenBank for one viral strain per EEHV subspecies. To facilitate phylogenetic analysis, gH and gL sequences were concatenated. Sequences were aligned using Clustal Omega, and trees were constructed using IQTree based on the (<b>A</b>) JTT + G4 (gB) and (<b>B</b>) WAG + F + I + G4 (gH/gL) evolutionary model using 1000 bootstrap replicates. Inferred trees were visualized and edited in FigTree (<a href="http://tree.bio.ed.ac.uk/software/figtree/" target="_blank">http://tree.bio.ed.ac.uk/software/figtree/</a> (accessed on 23 February 2023)). Presented trees are midpoint rooted. Only bootstrap values  ≥  70 are shown. EEHV (sub)species, viral strain analyzed, and GenBank accession numbers are indicated. EEHV subspecies infecting Asian elephants are highlighted in red; subspecies infecting African elephants are shown in black. Branches including the AT-rich EEHVs [<a href="#B1-viruses-16-00268" class="html-bibr">1</a>] are colored dark blue; branches including the GC-rich EEHVs [<a href="#B1-viruses-16-00268" class="html-bibr">1</a>] are shown in teal.</p>
Full article ">Figure 3
<p>Production of recombinant gB and gH/gL proteins of the Asian elephant EEHV (sub)species. Gelcode Blue-stained gels electrophoresed with gB-3×ST (<b>A</b>) and gH-3×ST/gL-6×His (<b>B</b>) for EEHV subspecies 1A, 1B, 4, and 5A affinity purified using the StrepTag. Molecular mass markers are indicated on the left side of the gels. (<b>A</b>) Expected molecular weights of glycosylated gB proteins are listed in the table on the right. (<b>B</b>) gH/gL protein fractions were deglycosylated by PNGaseF (~35kDa) prior to electrophoresis. Expected molecular weights of deglycosylated gH and gL proteins are listed in the table on the right. (<b>C</b>) Western blot of gels electrophoresed with secreted gH-3×ST/gL-6×His proteins for EEHV subspecies 1A, 1B, 4, and 5A, stained using anti-StrepTag and anti-HisTag antibodies. Protein fractions were deglycosylated by PNGaseF prior to electrophoresis. Expected molecular weights as in panel (<b>B</b>).</p>
Full article ">Figure 4
<p>Antibody levels against gB and gH/gL of different EEHV (sub)species measured in sera of Asian elephants &lt; 10 years of age. Antibody levels were measured against gB (<b>A</b>) and gH/gL (<b>B</b>) of all EEHV subspecies for 76 sera of 51 individual animals &lt; 10 years of age. A maximum of two sera per animal were included. Samples were tested at a 1:100 dilution, and ΔOD values were calculated and normalized as described in Materials and Methods. Values are grouped per EEHV (sub)species showing both individual values and mean ± standard deviation (SD).</p>
Full article ">Figure 5
<p>Linear regression analyses of antibody levels against gB of the different EEHV subspecies. (<b>A</b>) Pairwise simple linear regression analyses for ΔOD values presented in <a href="#viruses-16-00268-f004" class="html-fig">Figure 4</a>A. R<sup>2</sup> and <span class="html-italic">p</span>-values calculated for each regression are shown in the top left corner of each panel. (<b>B</b>) Correlation between the R<sup>2</sup> levels obtained in (<b>A</b>) and amino acid identity (in %) as shown in <a href="#viruses-16-00268-t002" class="html-table">Table 2</a>. Calculated R<sup>2</sup> and <span class="html-italic">p</span>-values are shown in the top left corner.</p>
Full article ">Figure 6
<p>Linear regression analyses of antibody levels against gH/gL of different EEHV subspecies. (<b>A</b>) Pairwise simple linear regression analyses for ΔOD values presented in <a href="#viruses-16-00268-f004" class="html-fig">Figure 4</a>B. R<sup>2</sup> and <span class="html-italic">p</span>-values calculated for each regression are shown in the top left corner of each panel. (<b>B</b>–<b>D</b>) Correlation between the R<sup>2</sup> levels obtained in (<b>A</b>) and pairwise amino acid identity of gH/gL dimer ((<b>B</b>); shown in <a href="#viruses-16-00268-t003" class="html-table">Table 3</a>), gH (<b>C</b>), and gL (<b>D</b>). Calculated R<sup>2</sup> and <span class="html-italic">p</span>-values are shown in the top left corner of each panel.</p>
Full article ">Figure 7
<p>Antibody levels primarily reactive to gH/gL of a particular EEHV (sub)species as detected in sera of Asian elephants &lt; 10 years of age. Normalized ΔOD values detected in (<b>A</b>) sera of 11 individual elephants and (<b>B</b>) paired sera of 3 individual elephants using gH/gL ELISAs for the different EEHV subspecies. Sera of individual elephants are identified by capital letters (A–M), and for each serum sample, elephant age at time of sampling is indicated. Samples of elephants that previously tested PCR positive for a specific EEHV subspecies are indicated by ((sub)species)<sup>+</sup>. All samples were tested and ΔOD values were calculated and normalized as described in <a href="#viruses-16-00268-f004" class="html-fig">Figure 4</a>.</p>
Full article ">Figure 8
<p>Antibody levels detected in sera of 23 fatal EEHV-HD cases using the multiple-EEHV-species gB ELISA and (sub)species-specific gH/gL ELISAs. (<b>A</b>) gH/gL-specific antibody levels of 11 animals that showed virtually no gB-specific antibody levels in the last serum sample taken before death. EEHV-HD cases shown succumbed to EEHV1A (7 individuals), EEHV1B (2 individuals), or EEHV5 (1 individual). One animal succumbed to an EEHV1A/EEHV4 co-infection. (<b>B</b>–<b>E</b>) Individual panels showing gH/gL-specific antibody levels in (longitudinal) serum samples of 12 EEHV-HD cases with detectable gB-specific antibodies. The respective EEHV-HD cases died due to either an EEHV1A (<b>B</b>,<b>D</b>,<b>E</b>) or EEHV1B (<b>C</b>) infection. In each panel, the antibody level against gH/gL of the subspecies the animal succumbed to is indicated by a red dotted line, and the time of death is indicated by †. Antibody levels against gB are included for reference purposes. All samples were tested and values were normalized as described previously.</p>
Full article ">Figure 9
<p>Antibody levels measured against gH/gL of the different EEHV (sub)species with increasing age. (<b>A</b>) Antibody levels against gH/gL of the different EEHV (sub)species were assessed for a total of 298 sera, divided over 8 different age categories. Number of samples included per age category ranged from 9 to 61, with no more than two samples per individual animal included per age category. Proportion of samples for which ΔOD levels above an (arbitrary) cutoff level of 0.25 were detected in (<span class="html-italic">n</span> = 0–4)/4 gH/gL ELISAs is indicated by different colors. (<b>B</b>) Normalized ΔOD values detected in paired sera of five individual elephants &gt; 10 years of age using gH/gL ELISAs for the different EEHV (sub)species. Individual elephants are identified by a capital letter, and for each sample, elephant age at time of sampling is indicated. All samples were tested and values were normalized as described previously.</p>
Full article ">
19 pages, 3155 KiB  
Article
The Inovirus Pf4 Triggers Antiviral Responses and Disrupts the Proliferation of Airway Basal Epithelial Cells
by Medeea C. Popescu, Naomi L. Haddock, Elizabeth B. Burgener, Laura S. Rojas-Hernandez, Gernot Kaber, Aviv Hargil, Paul L. Bollyky and Carlos E. Milla
Viruses 2024, 16(1), 165; https://doi.org/10.3390/v16010165 - 22 Jan 2024
Viewed by 1281
Abstract
Background: The inovirus Pf4 is a lysogenic bacteriophage of Pseudomonas aeruginosa (Pa). People with Cystic Fibrosis (pwCF) experience chronic airway infection with Pa and a significant proportion have high numbers of Pf4 in their airway secretions. Given the known severe damage [...] Read more.
Background: The inovirus Pf4 is a lysogenic bacteriophage of Pseudomonas aeruginosa (Pa). People with Cystic Fibrosis (pwCF) experience chronic airway infection with Pa and a significant proportion have high numbers of Pf4 in their airway secretions. Given the known severe damage in the airways of Pa-infected pwCF, we hypothesized a high Pf4 burden can affect airway healing and inflammatory responses. In the airway, basal epithelial cells (BCs) are a multipotent stem cell population critical to epithelium homeostasis and repair. We sought to investigate the transcriptional responses of BCs under conditions that emulate infection with Pa and exposure to high Pf4 burden. Methods: Primary BCs isolated from pwCF and wild-type (WT) donors were cultured in vitro and exposed to Pf4 or bacterial Lipopolysaccharide (LPS) followed by transcriptomic and functional assays. Results: We found that BCs internalized Pf4 and this elicits a strong antiviral response as well as neutrophil chemokine production. Further, we found that BCs that take up Pf4 demonstrate defective migration and proliferation. Conclusions: Our findings are highly suggestive of Pf4 playing a role in the pathogenicity of Pa in the airways. These findings provide additional evidence for the ability of inoviruses to interact with mammalian cells and disrupt cell function. Full article
(This article belongs to the Special Issue Inoviruses)
Show Figures

Figure 1

Figure 1
<p>Single cell RNA sequencing of WT and CF BCs demonstrate distinct responses to Pf4. (<b>A</b>) Diagram of experimental workflow. (<b>B</b>) UMAP plots for all experimental conditions (Pf4, LPS and/or PBS control) for CF and WT cells identify specific functional clusters. Among cells exposed to Pf with or without LPS a cluster with an antiviral gene expression signature was clearly apparent (Cluster 8, circled) (<b>C</b>) Dot plot showing higher average expression of ISG15, IFIT3, OAS1, OAS2, and IRF7 by cluster identity. (<b>D</b>) Proportion of cells in cluster 8 by experimental condition. (<b>E</b>) Dot plot showing average expression of markers ISG15, IFIT3, OAS1, OAS2, and IRF7 across samples. (<b>F</b>) Dot plot showing average expression of markers CXCL1, CXCL2, CXCL5, CXCL6, and CXCL8 across all samples. Graphical schematics created with <a href="http://BioRender.com" target="_blank">BioRender.com</a>.</p>
Full article ">Figure 2
<p>WT and CF cells exhibit differing responses between Pf4 and Pf4 + LPS experimental exposure conditions. (<b>A</b>) UMAP of WT and CF BCs comparing only cells stimulated with Pf ± LPS. Cluster corresponding to antiviral responses circled in black. (<b>B</b>) Proportion of cells in defined clusters for WT and CF cells between LPS and Pf + LPS conditions. (<b>C</b>) GO terms corresponding to each cluster across categories.</p>
Full article ">Figure 3
<p>BCs initiate antiviral responses to Pf4 phage. (<b>A</b>) GO term enrichment for top upregulated genes in antiviral cluster. (<b>B</b>) Gene expression pattern across all clusters for top upregulated genes of antiviral cluster. (<b>C</b>) Gene expression for individual antiviral genes across control and Pf-stimulated conditions cystic fibrotic (left) and WT (right) cells. (<b>D</b>) Overall expression of individual antiviral genes in WT and CF cells exposed to LPS or Pf + LPS.</p>
Full article ">Figure 4
<p>Basal cells from additional donors validate antiviral signature in response to Pf4 phage. RT-qPCR of WT (<b>A</b>) and CF (<b>B</b>) basal cells incubated with Pf4 10<sup>10</sup> pfu/mL ± LPS 5 ug/mL for 24 h. (<b>C</b>) Expression of the selected validation genes in sequencing dataset. Protein quantification of CXCL10 in supernatants from WT (<b>D</b>) and CF (<b>E</b>) basal cells stimulated as in (<b>A</b>) as determined by ELISA, with 10<sup>10</sup> pfu/mL of additional phage controls. Significance between different conditions (** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001) was assessed by ANOVA followed by Holm-Šídák test for multiple comparisons.</p>
Full article ">Figure 5
<p>CF BCs internalize Pf4 phage. (<b>A</b>) Untreated BCs or (<b>B</b>) labelled Pf4 treated BCs stained with tubulin-CF555 demonstrate uptake of Pf4. To further determine intracellular localization, (<b>C</b>) Untreated BCs or (<b>D</b>) labelled Pf4 treated BCs stained with endosomal marker EEA-1-AF647 demonstrate Pf4 localization with endosomes. Confocal images taken at 100× magnification.</p>
Full article ">Figure 5 Cont.
<p>CF BCs internalize Pf4 phage. (<b>A</b>) Untreated BCs or (<b>B</b>) labelled Pf4 treated BCs stained with tubulin-CF555 demonstrate uptake of Pf4. To further determine intracellular localization, (<b>C</b>) Untreated BCs or (<b>D</b>) labelled Pf4 treated BCs stained with endosomal marker EEA-1-AF647 demonstrate Pf4 localization with endosomes. Confocal images taken at 100× magnification.</p>
Full article ">Figure 6
<p>Pf4 induces production of neutrophil chemotactic factors in WT and CF basal cells. Expression (<b>A</b>) and ridge plots (<b>B</b>) of selected neutrophil chemoattractant genes in cluster 0 of the 10x sequencing dataset. (<b>C</b>) Feature plots showing neutrophil chemokine expression across basal cell clusters. Protein concentration of neutrophil chemoattractants in supernatants from WT (<b>D</b>) and CF (<b>E</b>) basal cells incubated with 5 ug/mL LPS ± 10<sup>10</sup> pfu/mL Pf4 phage for 24 h as determined by ELISA. Significance between different conditions (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001) relative to PBS control (dotted line) was assessed by ANOVA followed by Holm-Šídák test for multiple comparisons.</p>
Full article ">Figure 7
<p>Pf4 decreases BCs progression towards confluence. (<b>A</b>) Representative 200× images of WT cells treated with LPS (<b>left</b>) and Pf + LPS (<b>right</b>) shown at 48 h. CF cells treated with LPS (<b>left</b>) and Pf + LPS (<b>right</b>) shown at 48 h. (<b>B</b>) Progression towards confluence shown for WT and CF cells treated with Pf4 and control conditions. (<b>C</b>) Mean area increase in confluence at 48 h for both genotypes and all conditions shown. Means and SD estimated by generalized linear model (GLM) and comparisons between different groups assessed by Tukey’s studentized range test (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01). (<b>D</b>) Time to 50% confluence (dotted line) of the seeded area for WT cells in the presence of Pf4 or PBS control was assessed between 54 and 86 h (<span class="html-italic">p</span> = 0.012).</p>
Full article ">
18 pages, 32199 KiB  
Article
Partial Atomic Model of the Tailed Lactococcal Phage TP901-1 as Predicted by AlphaFold2: Revelations and Limitations
by Jennifer Mahony, Adeline Goulet, Douwe van Sinderen and Christian Cambillau
Viruses 2023, 15(12), 2440; https://doi.org/10.3390/v15122440 - 15 Dec 2023
Viewed by 2340
Abstract
Bacteria are engaged in a constant battle against preying viruses, called bacteriophages (or phages). These remarkable nano-machines pack and store their genomes in a capsid and inject it into the cytoplasm of their bacterial prey following specific adhesion to the host cell surface. [...] Read more.
Bacteria are engaged in a constant battle against preying viruses, called bacteriophages (or phages). These remarkable nano-machines pack and store their genomes in a capsid and inject it into the cytoplasm of their bacterial prey following specific adhesion to the host cell surface. Tailed phages possessing dsDNA genomes are the most abundant phages in the bacterial virosphere, particularly those with long, non-contractile tails. All tailed phages possess a nano-device at their tail tip that specifically recognizes and adheres to a suitable host cell surface receptor, being proteinaceous and/or saccharidic. Adhesion devices of tailed phages infecting Gram-positive bacteria are highly diverse and, for the majority, remain poorly understood. Their long, flexible, multi-domain-encompassing tail limits experimental approaches to determine their complete structure. We have previously shown that the recently developed protein structure prediction program AlphaFold2 can overcome this limitation by predicting the structures of phage adhesion devices with confidence. Here, we extend this approach and employ AlphaFold2 to determine the structure of a complete phage, the lactococcal P335 phage TP901-1. Herein we report the structures of its capsid and neck, its extended tail, and the complete adhesion device, the baseplate, which was previously partially determined using X-ray crystallography. Full article
(This article belongs to the Section Bacterial Viruses)
Show Figures

Figure 1

Figure 1
<p>Schematic representation of the structural module of TP901-1. The functions are indicated above the arrows and the scale bar is presented at the base of the schematic, measured in base pairs (bp). This figure and associated functions are based on those described in reference [<a href="#B17-viruses-15-02440" class="html-bibr">17</a>].</p>
Full article ">Figure 2
<p>TP901-1 capsid’s hexon and penton structures. (<b>A</b>) Ribbon view of the predicted hexon structure. One monomer is rainbow colored and its domains are labeled. The monomer at the bottom of the figure is colored according to the pLDDT values (the quality of the prediction), from blue (low) to red (high). Note the low-confidence structure of the N-terminus (bottom dashed box). (<b>B</b>) Surface representation of the hexon (same orientation and scale as in (<b>A</b>)). (<b>C</b>) Ribbon view of the predicted penton structure. (<b>D</b>) Surface representation of the penton (same orientation and scale as in (<b>C</b>)).</p>
Full article ">Figure 3
<p>The procapsid’s scaffolding proteins (SPs). (<b>A</b>) Amino acid alignment between the scaffolding proteins of staphylococcal phage 80α and lactococcal phage TP901-1. (<b>B</b>) Ribbon representation of the cryoEM structure of a phage 80α MCP hexamer (blue) in complex with the C-terminus of the scaffolding protein (orange; see also the orange box in (<b>A</b>)) [<a href="#B1-viruses-15-02440" class="html-bibr">1</a>]. (<b>C</b>) AF2 prediction of the full-length phage 80α scaffolding protein. The orange box corresponds to the orange helix in (<b>B</b>). (<b>D</b>) AF2 prediction of the full-length phage TP901-1 scaffolding protein. The green-boxed helices correspond to the green box in (<b>A</b>). (<b>E</b>, <b>top</b>) Ribbon representation of a phage TP901-1 MCP hexamer (light blue) in complex with the C-terminus of the scaffolding protein (orange; see also the green box in (<b>A</b>)). (<b>E</b>, <b>bottom</b>) Surface view of the same structure rotated 90°.</p>
Full article ">Figure 4
<p>The structures of the dodecameric portal and adaptor. (<b>A</b>) The portal dodecamer has a size of 160 × 110 Å. The C-terminal segments exhibit low pLDDT values and have been boxed in gray. (<b>B</b>) The structure of the portal monomer with the different domains conserved in all portals. The C-terminal extension is poorly predicted and is disordered. (<b>C</b>) Lateral view of the adaptor with the extended C-terminal α-helices and the central β-sheet. (<b>D</b>) The same view rotated 90°, relative to (<b>C</b>).</p>
Full article ">Figure 5
<p>Protein structure from the portal to the major tail portein. (<b>A</b>) Dodecameric portal (yellow) and adaptor (violet), hexameric stopper (light blue), tail terminator (TT, green), and major tail protein (MTP, gray). The C-terminal segments exhibit low pLDDT values and have been boxed in gray. (<b>B</b>) The portal/adaptor interface (close-up view in inset one, <b>i1</b>) involves four b-strands: portal i, s13 and s11 (anti-parallel); portal i + 1, s12 and adaptor C-terminal b-strand stacks against a-helix 6; and portal i + 1, a-strands s11 and s13. (<b>C</b>) The adaptor/stopper interface involves two adaptors (beige and orange) and one stopper (violet). Contacts originate from the stopper’s N-terminus and loops joining the a-strands of the three monomers. The insertion of the adaptor displaces a-helix 6 (inset two, <b>i2</b>). (<b>D</b>) The stopper/TT interface: The long stopper’s b-hairpin loop s2–s3 (light green) is inserted between two TT monomers (red and beige). Other contacts involve the long TT’s loops h1–s1 and s3–s5. (<b>E</b>) The TT/MTP interface: The MTP i − 1 N-terminus (beige) is inserted between MTP i (orange) and TT (light blue), and comprises the majority of TT/MTP interactions. The MTP C-terminus is inserted vertically between two TT monomers.</p>
Full article ">Figure 6
<p>The fit of the neck in the nsEM density map. (<b>A</b>) The fit of the neck (portal, 12 mer, yellow; adaptor, 12 mer, pink; stopper, 6 mer, light blue; NPS N-termini, 12 mer, white) in the nsEM map at a 20.0 Å resolution. (<b>B</b>) Same view as in (<b>A</b>), but split at half its diameter. The NPS N-termini are located just below the portal/adaptor junction. (<b>C</b>) View at 90° of (<b>A</b>). The NPS N-termini cover the cavities observed at the portal/adaptor interface. (<b>D</b>) The phage 80α fiber structure was superimposed onto the NPS N-termini to illustrate a possible structure of the NPSs. (<b>E</b>) Surface view of the structural prediction of the NPS trimer with a total length of 48 nm.</p>
Full article ">Figure 7
<p>The tail tube. (<b>A</b>) Surface view of four MTP rings forming a section of the tail tube (the top of the view is the direction of capsid). Note the insertion of each monomer C-terminus inside a monomer cavity in the above ring. (<b>B</b>) Same orientation as in (<b>A</b>) but split at its half-diameter and colored according to electrostatics. (<b>C</b>) Fit of the MTPs in the nsEM map at 20Å resolution (EMD-2228). (<b>D</b>) MTP/MTP interactions between the stacked rings. Interactions involve the N- and C-termini and the b-hairpin with strands s3 and s4.</p>
Full article ">Figure 8
<p>The distal MTP and the central baseplate. (<b>A</b>) Surface view of the distal hexameric MTP (gray), the hexameric Dit (blue), and the trimeric Tal (orange). Ga: the Dit’s galectin domain; St: the Tal’s conserved structural domain; Ld1 and Ld2: linker domains 1 and 2; pgd and endopeptidase (ep) domains. (<b>B</b>) Ribbon view of the MTP/Dit interface. (<b>C</b>) Ribbon view of the Dit/Tal interface. (<b>D</b>) Ribbon side-view of the Tal trimer N-terminus (residues 1–484). (<b>E</b>) Same view rotated by 90° within the phage’s main axis.</p>
Full article ">Figure 9
<p>The peripheral baseplate. (<b>A</b>) Surface view of the dodecameric BppU N-terminal domain (yellow) complexed to the hexameric Dit (blue): MTP: gray; Tal: orange. (<b>B</b>) Split view of the nsEM map (EMD-1793; 25 Å resolution) of the complete baseplate with the distal hexameric MTP (gray), the hexameric Dit (blue), the trimeric Tal (orange), and the dodecameric BppU N-terminal domain (yellow) fitted inside. The gray ribbon inside the map represents the X-ray structure (PDB id 4v96). (<b>C</b>) The AF2-predicted structure of the “tripod”, formed of the trimeric BppU C-terminal domain holding three RBP trimers. (<b>D</b>) Close-up of the BppU/RBP interface with the hydrophobic residues Ile219, Phe226, and Phe232 also identified in the X-ray structure. (<b>E</b>) Surface view of the “tripod” fit in the nsEM map.</p>
Full article ">
38 pages, 2700 KiB  
Review
Evolving Horizons: Adenovirus Vectors’ Timeless Influence on Cancer, Gene Therapy and Vaccines
by Prasad D. Trivedi, Barry J. Byrne and Manuela Corti
Viruses 2023, 15(12), 2378; https://doi.org/10.3390/v15122378 - 3 Dec 2023
Cited by 3 | Viewed by 4242
Abstract
Efficient and targeted delivery of a DNA payload is vital for developing safe gene therapy. Owing to the recent success of commercial oncolytic vector and multiple COVID-19 vaccines, adenovirus vectors are back in the spotlight. Adenovirus vectors can be used in gene therapy [...] Read more.
Efficient and targeted delivery of a DNA payload is vital for developing safe gene therapy. Owing to the recent success of commercial oncolytic vector and multiple COVID-19 vaccines, adenovirus vectors are back in the spotlight. Adenovirus vectors can be used in gene therapy by altering the wild-type virus and making it replication-defective; specific viral genes can be removed and replaced with a segment that holds a therapeutic gene, and this vector can be used as delivery vehicle for tissue specific gene delivery. Modified conditionally replicative–oncolytic adenoviruses target tumors exclusively and have been studied in clinical trials extensively. This comprehensive review seeks to offer a summary of adenovirus vectors, exploring their characteristics, genetic enhancements, and diverse applications in clinical and preclinical settings. A significant emphasis is placed on their crucial role in advancing cancer therapy and the latest breakthroughs in vaccine clinical trials for various diseases. Additionally, we tackle current challenges and future avenues for optimizing adenovirus vectors, promising to open new frontiers in the fields of cell and gene therapies. Full article
(This article belongs to the Special Issue Research and Clinical Application of Adenovirus (AdV), 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Highlights of adenovirus capsid features with overview of viral genome. In the top right corner, the species (A–G) of AdVs with their known tropism are indicated. The schematic representation of the gene map is for understanding purposes only and is not normalized for actual gene size. Kb: kilobases; mu: map unit. Created with <a href="http://BioRender.com" target="_blank">BioRender.com</a> (accessed on 1 December 2023).</p>
Full article ">Figure 2
<p>Represents an overview of the scope of the modifications and packaging capacity of rAdV vectors. Red delta signifies the possibility of gene deletion for creating multi-generational rAdVs. The suggested approximate insert size of the gene of interest (GOI) depends on the specific application. Early and late genes are explained in previous sections. The schematic representation of the gene map is for understanding purposes only and is not normalized for actual gene size. Kb: kilobases; mu: map unit Created with <a href="http://BioRender.com" target="_blank">BioRender.com</a> (accessed on 1 December 2023).</p>
Full article ">Figure 3
<p>Shows an overview of the clinical trials trend. The red highlighted point on the graph indicates the decline in adenovirus vector application and the green highlighted section represents the recent pandemic, which points towards the upward trajectory of revamped interest in adenovirus vectors. Created with <a href="http://BioRender.com" target="_blank">BioRender.com</a> (accessed on 1 December 2023).</p>
Full article ">Figure 4
<p>Represents simplified version of large-scale operation for manufacturing adenovirus vector using specified platform. During clone development, a master cell bank (MCB) and Master Virus Seed is generated, then a working cell bank can be made for further use. This initial stage changes a great deal depending on the process development required for the specific generation of AdV vector. As optimized in this initial stage, the upstream manufacturing platform is chosen to be performed by infecting in either adherent cell line format or in suspension cell line format. The purification and final polishing of the vector is conducted by using combination of optimized chromatography steps such as immobilized metal-ion affinity chromatography (IMAC), anion exchange chromatography (AEX), and/or gel filtration chromatography (GFC). Depending on the required final concentration of vector in specific formulation buffer for increased stability, tangential flow filtration and diafilteration (TFF) can be performed before the fill-finish process. Created with <a href="http://BioRender.com" target="_blank">BioRender.com</a> (accessed on 1 December 2023).</p>
Full article ">
16 pages, 2322 KiB  
Article
Enhanced Susceptibility to Tomato Chlorosis Virus (ToCV) in Hsp90- and Sgt1-Silenced Plants: Insights from Gene Expression Dynamics
by Irene Ontiveros, Noé Fernández-Pozo, Anna Esteve-Codina, Juan José López-Moya and Juan Antonio Díaz-Pendón
Viruses 2023, 15(12), 2370; https://doi.org/10.3390/v15122370 - 30 Nov 2023
Cited by 1 | Viewed by 1453
Abstract
The emerging whitefly-transmitted crinivirus tomato chlorosis virus (ToCV) causes substantial economic losses by inducing yellow leaf disorder in tomato crops. This study explores potential resistance mechanisms by examining early-stage molecular responses to ToCV. A time-course transcriptome analysis compared naïve, mock, and ToCV-infected plants [...] Read more.
The emerging whitefly-transmitted crinivirus tomato chlorosis virus (ToCV) causes substantial economic losses by inducing yellow leaf disorder in tomato crops. This study explores potential resistance mechanisms by examining early-stage molecular responses to ToCV. A time-course transcriptome analysis compared naïve, mock, and ToCV-infected plants at 2, 7, and 14 days post-infection (dpi). Gene expression changes were most notable at 2 and 14 dpi, likely corresponding to whitefly feeding and viral infection. Gene Ontology (GO) and Kyoto Encyclopedia of Genes and Genomes (KEGG) enrichment analyses revealed key genes and pathways associated with ToCV infection, including those related to plant immunity, flavonoid and steroid biosynthesis, photosynthesis, and hormone signaling. Additionally, virus-derived small interfering RNAs (vsRNAs) originating from ToCV predominantly came from RNA2 and were 22 nucleotides in length. Furthermore, two genes involved in plant immunity, Hsp90 (heat shock protein 90) and its co-chaperone Sgt1 (suppressor of the G2 allele of Skp1) were targeted through viral-induced gene silencing (VIGS), showing a potential contribution to basal resistance against viral infections since their reduction correlated with increased ToCV accumulation. This study provides insights into tomato plant responses to ToCV, with potential implications for developing effective disease control strategies. Full article
(This article belongs to the Special Issue Plant Virus Resistance)
Show Figures

Figure 1

Figure 1
<p>Exploratory analysis of transcriptomic data. (<b>A</b>) Principal component analysis (PCA) of all replicates from naïve (no whitefly and no virus), Mock (non-viruliferous whiteflies), and ToCV (ToCV-viruliferous whiteflies) samples at 2, 7, and 14 days post-infection (dpi). (<b>B</b>) Number of up-regulated and down-regulated differentially expressed genes (DEGs) at each time point after ToCV infection. Red and blue bars represent numbers of up-regulated and down-regulated genes, respectively. Total indicates the total number of both up-regulated and down-regulated DEGs. (<b>C</b>) Venn diagram displaying the number of shared and distinct DEGs across the three specified time points.</p>
Full article ">Figure 2
<p>Gene ontology (GO) enrichment analysis of the differentially expressed genes (DEGs) in response to ToCV infection at 2, 7, and 14 days post inoculation (dpi). Enrichment of GO terms among the up-regulated (<b>A</b>) and down-regulated (<b>B</b>) DEGs. Each cell is colored based on the number of genes associated with the respective GO term.</p>
Full article ">Figure 3
<p>KEGG enrichment analysis. Top KEGG pathways enriched with up-regulated (<b>A</b>) and down-regulated (<b>B</b>) differentially expressed genes (DEGs) triggered by ToCV infection at 2, 7, and 14 days post-inoculation (dpi).</p>
Full article ">Figure 4
<p>Global analysis of virus-derived small RNAs (vsRNAs) in tomato plants infected with ToCV. (<b>A</b>) Percentage of vsRNAs in the 20–25 nt reads pool mapped to RNA1 and RNA2 of the ToCV genome. (<b>B</b>) Single-nucleotide resolution maps of vsRNAs from tomato plants challenged by ToCV. Positive- and negative-strand reads are shown in blue and red, respectively. Genome organization of each viral genomic RNA is shown.</p>
Full article ">Figure 5
<p>Impact of silencing of tomato <span class="html-italic">Hsp90</span> and <span class="html-italic">Sgt1</span> genes through virus-induced gene silencing on the susceptibility to ToCV infection. (<b>A</b>) Phenotypes observed in ToCV-infected tomato plants at 14 days post-inoculation (dpi) that were agroinfiltrated 7 days earlier with tobacco rattle virus (TRV) vector alone (TRV2), with the TRV Sgt1- and Hsp90-silencing constructs (TRV-<span class="html-italic">Sgt1</span> and TRV-<span class="html-italic">Hsp90</span>) and mock-inoculated (Mock). (<b>B</b>) Relative accumulation of ToCV RNA at 12 dpi analyzed by reverse transcription–quantitative polymerase chain reaction (RT-qPCR). The plants were agroinfiltrated 7 days earlier with TRV2, TRV-<span class="html-italic">Sgt1</span>, TRV-<span class="html-italic">Hsp90,</span> or Mock. Values were normalized using tomato elongation factor 1-α and Sand as reference genes, with Mock serving as the calibrator. Error bars represent standard errors of five biological replicates and an asterisk indicates a significant difference according to one-way ANOVA with <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">
24 pages, 3513 KiB  
Review
The Functional Implications of Broad Spectrum Bioactive Compounds Targeting RNA-Dependent RNA Polymerase (RdRp) in the Context of the COVID-19 Pandemic
by Brittany A. Comunale, Robin J. Larson, Erin Jackson-Ward, Aditi Singh, Frances L. Koback and Lilly D. Engineer
Viruses 2023, 15(12), 2316; https://doi.org/10.3390/v15122316 - 25 Nov 2023
Cited by 1 | Viewed by 1453
Abstract
Background: As long as COVID-19 endures, viral surface proteins will keep changing and new viral strains will emerge, rendering prior vaccines and treatments decreasingly effective. To provide durable targets for preventive and therapeutic agents, there is increasing interest in slowly mutating viral proteins, [...] Read more.
Background: As long as COVID-19 endures, viral surface proteins will keep changing and new viral strains will emerge, rendering prior vaccines and treatments decreasingly effective. To provide durable targets for preventive and therapeutic agents, there is increasing interest in slowly mutating viral proteins, including non-surface proteins like RdRp. Methods: A scoping review of studies was conducted describing RdRp in the context of COVID-19 through MEDLINE/PubMed and EMBASE. An iterative approach was used with input from content experts and three independent reviewers, focused on studies related to either RdRp activity inhibition or RdRp mechanisms against SARS-CoV-2. Results: Of the 205 records screened, 43 studies were included in the review. Twenty-five evaluated RdRp activity inhibition, and eighteen described RdRp mechanisms of existing drugs or compounds against SARS-CoV-2. In silico experiments suggested that RdRp inhibitors developed for other RNA viruses may be effective in disrupting SARS-CoV-2 replication, indicating a possible reduction of disease progression from current and future variants. In vitro, in vivo, and human clinical trial studies were largely consistent with these findings. Conclusions: Future risk mitigation and treatment strategies against forthcoming SARS-CoV-2 variants should consider targeting RdRp proteins instead of surface proteins. Full article
(This article belongs to the Special Issue Broad-Spectrum Antivirals and Interaction with Viruses)
Show Figures

Figure 1

Figure 1
<p>Flow chart of the study selection process.</p>
Full article ">Figure 2
<p>Key amino acid residues of the SARS-CoV-2 RNA-dependent RNA polymerase (RdRp) nsp12 complex with cofactors nsp7 and nsp8 interact with bioactive compound inhibitors targeting RdRp. Catalytic residues (highlighted in blue: Asp760, Asp761, and Ser759) and other active site residues (Asp618 and Tyr619) have the strongest contact with inhibitors via hydrophobic interactions and hydrogen bonding. Electrostatic and steric interactions are seen with amino acid residues positioned near the catalytic center of functional motif C (Arg553 and Lys551). When inhibitors interact with the residues of conserved motifs, the protein declines in functionality, as normal replication processes are disrupted.</p>
Full article ">
12 pages, 2187 KiB  
Article
A Robust Phenotypic High-Throughput Antiviral Assay for the Discovery of Rabies Virus Inhibitors
by Xinyu Wang, Winston Chiu, Hugo Klaassen, Arnaud Marchand, Patrick Chaltin, Johan Neyts and Dirk Jochmans
Viruses 2023, 15(12), 2292; https://doi.org/10.3390/v15122292 - 23 Nov 2023
Viewed by 1807
Abstract
Rabies virus (RABV) causes severe neurological symptoms in mammals. The disease is almost inevitably lethal as soon as clinical symptoms appear. The use of rabies immunoglobulins (RIG) and vaccination in post-exposure prophylaxis (PEP) can provide efficient protection, but many people do not receive [...] Read more.
Rabies virus (RABV) causes severe neurological symptoms in mammals. The disease is almost inevitably lethal as soon as clinical symptoms appear. The use of rabies immunoglobulins (RIG) and vaccination in post-exposure prophylaxis (PEP) can provide efficient protection, but many people do not receive this treatment due to its high cost and/or limited availability. Highly potent small molecule antivirals are urgently needed to treat patients once symptoms develop. In this paper, we report on the development of a high-throughput phenotypic antiviral screening assay based on the infection of BHK-21 cells with a fluorescent reporter virus and high content imaging readout. The assay was used to screen a repurposing library of 3681 drugs (all had been studied in phase 1 clinical trials). From this series, salinomycin was found to selectively inhibit viral replication by blocking infection at the entry stage. This shows that a high-throughput assay enables the screening of large compound libraries for the purposes of identifying inhibitors of RABV replication. These can then be optimized through medicinal chemistry efforts and further developed into urgently needed drugs for the treatment of symptomatic rabies. Full article
(This article belongs to the Special Issue Rabies Virus: Treatment and Prevention)
Show Figures

Figure 1

Figure 1
<p>Optimization of the RABV 384-well antiviral screening. (<b>A</b>) Representative photomicrographs at 4 days after the infection of BHK-21 cells with RABV-mCherry. Unprocessed images of cell nuclei (Hoechst staining), RABV-mCherry, and the merger of both channels are depicted in the upper panels. For image analysis (bottom panels), the areas defined by a fixed pixel intensity threshold were selected and, based on nuclei staining, the cells were encircled with blue lines and counted as objects. Similarly, virus infection is detected in green circles; an orange circle within an object is counted as an uninfected cell. (<b>B</b>) Optimization of RABV dilutions and incubation time. Viral infection increased and viable cell number (nuclei count compared with uninfected controls) decreased with time. At an MOI of 0.019 TCID<sub>50</sub>/cell and a readout on day 4, an optimal condition was obtained with 86% infected cells and 71% cell viability. Data are from 3 independent experiments, and mean and standard errors are presented. (<b>C</b>) Based on the same dataset, the value of the Z’ factor and the percentage of cell viability at a MOI 0.019 TCID<sub>50</sub>/cell at different time points of readout are calculated.</p>
Full article ">Figure 2
<p>Schematic representation of the 384-well anti-RABV assay with high content imaging (HCI) used for HTS and the 96-well anti-RABV assay with whole-well fluorescence readout with a plate reader (PR) used for hit confirmation. (<b>A</b>) Schematic representation of the 384-well assay with HCI. After a 24 h incubation of 2000 BHK-21 cells with the test compounds, the cells were infected with mCherry-RABV (MOI 0.019 TCID<sub>50</sub>/cell). At 4 dpi, Hoechst was added to each well; virus infection and cell viability were assessed via HCI analysis. (<b>B</b>) Schematic representation of the 96-well assay. Test compounds were serially diluted and incubated with 1.5 × 10<sup>4</sup> BHK-21 cells per well in 96-well plates. The next day, cultures were infected with mCherry-RABV (MOI 0.01 TCID<sub>50</sub>/cell), or the medium without the virus (viability assay) was added. On day 5 pi, the fluorescence intensity of the virus infection and cell viability (MTS assay) were quantified (using a plate reader (PR)).</p>
Full article ">Figure 3
<p>Results of the antiviral screen of a repurposing library of 3681 compounds. Each dot represents the result of the testing of a single compound at 10 µM. The area with red dots indicates the selection criteria for the hits (&gt;60% cell viability, &lt;25% virus infection), resulting in the identification of 4 hits.</p>
Full article ">Figure 4
<p>Salinomycin inhibits RABV at an early stage of infection. (<b>A</b>) Chemical structure of salinomycin. (<b>B</b>–<b>E</b>) Dose response effect of salinomycin on RABV replication and on the viability of BHK-21 and SH-SY5Y cells. Each condition was tested in three independent assays; averages and STDEV are indicated. Calculated EC<sub>50</sub>s in BHK-21 cells are 0.051 µM and 0.18 µM on SAD B19 and CVS-11, respectively; and 0.9 µM and 0.67 µM on SAD B19 and CVS-11, respectively, in SH-SY5Y cells. Calculated CC<sub>50</sub>s are 4.0 µM on BHK-21 and SH-SY5Y cells. (<b>F</b>) Schematic of the time-of-drug-addition assay. BHK-21 cells were infected with RABV (CVS-11) either with or without the compound and incubated at 4 °C for 1 h (−1 to 0 hpi). After 3 washes with cold PBS, the plate was transferred to 37 °C (at 0 hpi). Salinomycin (0.5 µM) was added at various time points (−1, 0, 1, 0.5, 1, 2, 4 hpi) before collecting the cells at 16 hpi. Infected cells without the compound treatment were defined as untreated controls, and the sample that was collected at 1 hpi was considered the input virus (background). (<b>G</b>) Intracellular viral RNA of samples from different time points. Hydroxychloroquine (HOQ) was used as a reference compound for the virus entry. Data were from three independent assays; averages and STDEV are given. Dunnett’s multiple comparisons test was used to calculate the statistical significance. (* <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span>&lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 4 Cont.
<p>Salinomycin inhibits RABV at an early stage of infection. (<b>A</b>) Chemical structure of salinomycin. (<b>B</b>–<b>E</b>) Dose response effect of salinomycin on RABV replication and on the viability of BHK-21 and SH-SY5Y cells. Each condition was tested in three independent assays; averages and STDEV are indicated. Calculated EC<sub>50</sub>s in BHK-21 cells are 0.051 µM and 0.18 µM on SAD B19 and CVS-11, respectively; and 0.9 µM and 0.67 µM on SAD B19 and CVS-11, respectively, in SH-SY5Y cells. Calculated CC<sub>50</sub>s are 4.0 µM on BHK-21 and SH-SY5Y cells. (<b>F</b>) Schematic of the time-of-drug-addition assay. BHK-21 cells were infected with RABV (CVS-11) either with or without the compound and incubated at 4 °C for 1 h (−1 to 0 hpi). After 3 washes with cold PBS, the plate was transferred to 37 °C (at 0 hpi). Salinomycin (0.5 µM) was added at various time points (−1, 0, 1, 0.5, 1, 2, 4 hpi) before collecting the cells at 16 hpi. Infected cells without the compound treatment were defined as untreated controls, and the sample that was collected at 1 hpi was considered the input virus (background). (<b>G</b>) Intracellular viral RNA of samples from different time points. Hydroxychloroquine (HOQ) was used as a reference compound for the virus entry. Data were from three independent assays; averages and STDEV are given. Dunnett’s multiple comparisons test was used to calculate the statistical significance. (* <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span>&lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">
19 pages, 8713 KiB  
Article
An Inducible ESCRT-III Inhibition Tool to Control HIV-1 Budding
by Haiyan Wang, Benoit Gallet, Christine Moriscot, Mylène Pezet, Christine Chatellard, Jean-Philippe Kleman, Heinrich Göttlinger, Winfried Weissenhorn and Cécile Boscheron
Viruses 2023, 15(12), 2289; https://doi.org/10.3390/v15122289 - 22 Nov 2023
Viewed by 1237
Abstract
HIV-1 budding as well as many other cellular processes require the Endosomal Sorting Complex Required for Transport (ESCRT) machinery. Understanding the architecture of the native ESCRT-III complex at HIV-1 budding sites is limited due to spatial resolution and transient ESCRT-III recruitment. Here, we [...] Read more.
HIV-1 budding as well as many other cellular processes require the Endosomal Sorting Complex Required for Transport (ESCRT) machinery. Understanding the architecture of the native ESCRT-III complex at HIV-1 budding sites is limited due to spatial resolution and transient ESCRT-III recruitment. Here, we developed a drug-inducible transient HIV-1 budding inhibitory tool to enhance the ESCRT-III lifetime at budding sites. We generated autocleavable CHMP2A, CHMP3, and CHMP4B fusion proteins with the hepatitis C virus NS3 protease. We characterized the CHMP-NS3 fusion proteins in the absence and presence of protease inhibitor Glecaprevir with regard to expression, stability, localization, and HIV-1 Gag VLP budding. Immunoblotting experiments revealed rapid and stable accumulation of CHMP-NS3 fusion proteins. Notably, upon drug administration, CHMP2A-NS3 and CHMP4B-NS3 fusion proteins substantially decrease VLP release while CHMP3-NS3 exerted no effect but synergized with CHMP2A-NS3. Localization studies demonstrated the relocalization of CHMP-NS3 fusion proteins to the plasma membrane, endosomes, and Gag VLP budding sites. Through the combined use of transmission electron microscopy and video-microscopy, we unveiled drug-dependent accumulation of CHMP2A-NS3 and CHMP4B-NS3, causing a delay in HIV-1 Gag-VLP release. Our findings provide novel insight into the functional consequences of inhibiting ESCRT-III during HIV-1 budding and establish new tools to decipher the role of ESCRT-III at HIV-1 budding sites and other ESCRT-catalyzed cellular processes. Full article
(This article belongs to the Section Human Virology and Viral Diseases)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic illustrating our method for transient expression of ESCRT-III fused to NS3 protease and a fluorescent protein using a drug. ESCRT-III proteins are fused to the NS3 cleavage site, NS3 protease, and fluorescent protein. Cells transfected with this construct express wild-type ESCRT-III proteins. Upon addition of the drug, a full-length fusion protein accumulates. For imaging and immunoblotting detection, fluorescent proteins (mostly mNeonGreen) and a Flag tag were further added to the NS3 C-terminus, creating CHMP4B/2A/3-NS3-FP-Flag constructs (hereafter collectively referred to as CHMP-NS3-FP).</p>
Full article ">Figure 2
<p>Complete CHMP-NS3-FP proteins are stably expressed over time. (<b>a</b>–<b>c</b>) Representative immunoblot experiments were performed using whole cell extracts from HEK293 cells transfected with CHMP2A-NS3-green (2 μg) (<b>a</b>), CHMP3-NS3-green (2 μg) (<b>b</b>), or CHMP4B-NS3-green (1 μg) (<b>c</b>) and treated with DMSO or Glecaprevir for the indicated duration. (<b>d</b>–<b>f</b>) Pulse-chase experiment. HEK293 cells transfected with CHMP2A-NS3-green (<b>d</b>), CHMP3-NS3-green (<b>e</b>), or CHMP4B-NS3-green (<b>f</b>) were treated with DMSO or Glecaprevir for 4 h. Subsequently, the medium was washed away and replaced with fresh medium, and whole cell extracts were obtained 24 h later.</p>
Full article ">Figure 3
<p>In vivo localization of CHMPs-NS3-FP proteins. (<b>a</b>) Distribution on CHMP-NS3-FP in HeLA CCL2 cells: Cells transfected with CHMP2A-NS3-blue (left panels), CHMP3-NS3-blue (middle panels), or CHMP4B-NS3-blue (right panels) were treated for 4 h with DMSO or Glecaprevir as indicated. (<b>b</b>) CHMPs-NS3-blue accumulate in endosomes: Cells cotransfected with GFP-p40Phox and CHMP2A-NS3-blue (upper panels), CHMP3-NS3-blue (middle panels), or CHMP4B-NS3-blue (lower panels) were treated for 4 h with DMSO or Glecaprevir as indicated. (<b>c</b>) Colocalization of CHMP-NS3-green proteins with Gag-mCherry: Cells cotransfected with Gag, Gag-mCherry and CHMP2A-NS3-green (upper panels), CHMP3-NS3-green (middle panels), or CHMP4B-NS3-green (lower panels) were treated for 4 h with DMSO or Glecaprevir as indicated. Scale bars are 10 μm. (<b>d</b>) Quantification of Gag-mCherry spots as a proportion colocalizing with CHMP2A-NS3-green or CHMP4B-NS3-green (mean ± SD, n &gt; 230 spots for each condition from 3 cells). The statistical significance (***, indicating <span class="html-italic">p</span> &lt; 0.001), was determined through a Mann–Whitney test.</p>
Full article ">Figure 4
<p>Inhibition of VLP release by CHMP4B-NS3-green, CHMP2A-NS3-green, and CHMP3- NS3-green: immunoblot analysis. (<b>a</b>–<b>c</b>) Inhibition of VLP release in cells cotransfected by Gag (0.5 μg), GFP-Vps4A E228Q (1 μg), and CHMP4B-NS3-green (1 μg) (<b>a</b>), CHMP2A-NS3-green (4 μg) (<b>b</b>), and CHMP3-NS3-green (4 μg) (<b>c</b>). Representative immunoblot experiment depicting the following: Left panels (i) upper line: HIV-1 VLP pellet, (ii) second line: Gag HIV-1 cellular expression (WCE: whole cell extract), and right panels: CHMPs-NS3-green cellular expression. Hek293 cells were transfected with the following: column 1: Gag, column 2: Gag and CHMPs-NS3-green treated with DMSO for 2 h, column 3: Gag and GFP-Vps4A E228Q, column 4: Gag and CHMPs-mut-NS3-green, column 5: Gag and CHMPs-NS3-green treated with Glecaprevir for 4 h. (<b>d</b>) Representative immunoblot depicting Hek293 cells transfected with Gag and CHMP3-YFP. The upper lines represent HIV-1 VLPs released, the second lines indicate HIV-1 cellular expression (WCE: whole cell extract), and the third lines depict CHMP3-YFP cellular expression. (<b>e</b>) Synergy between CHMP3-NS3-green and CHMP2A-NS3-green. To enhance sensitivity, the amount of transfected CHMP2A-NS3-green was reduced to 2 μg, allowing for a slight impairment of VLP release. Cotransfection of Gag (0.5 μg), CHMP2A-NS3-green (2 μg), and CHMP3-NS3-green (2 μg) clearly enhance the VLP release inhibition. Representative immunoblot depicting Hek293 cells transfected with left panels: Gag and CHMP2A-NS3-green, right panels: Gag, CHMP2A-NS3-green, and CHMP3-NS3-green, treated with DMSO or Glecaprevir as indicated. The upper lines represent released HIV-1 VLPs, the second lines indicate HIV-1 cellular expression, and the third lines depict CHMPs-NS3-green cellular expression. (<b>f</b>) VLP release of the tested constructs analyzed by Western blot. Data are presented for Gag, GFP-Vps4A E228Q, GFP-Vps4B R253A, CHMP4B-NS3-green, CHMP2A-NS3-green, and CHMP3-NS3-green, as indicated (mean ± SD, n = 3 for each condition). The statistical significance (* for <span class="html-italic">p</span> &lt; 0.05 and *** for <span class="html-italic">p</span> &lt; 0.001), were determined through a Mann–Whitney test. HIV-1 Gag proteins were detected using anti-p24 antibody, CHMP3-YFP were detected using anti-GFP antibody, while CHMPs-NS3-green were detected using Anti-Flag antibody.</p>
Full article ">Figure 5
<p>Dose- and cell-line-dependent analysis via immunoblotting. (<b>a</b>,<b>b</b>) Inhibition of VLP release in cells cotransfected with Gag (0.5 μg) alongside varying quantities of either CHMP2A-NS3mut-green or CHMP2A-NS3-green, treated with DMSO or Glecaprevir. (<b>a</b>) Illustrative immunoblot experiment showing (top line) HIV-1 VLP pellet, (second line) Gag HIV-1 cellular expression (WCE: whole cell extract), and (lower line) CHMP2A-NS3-green cellular expression. Hek293 cells or HeLa Kyoto (Bst2−) were transfected as indicated. (<b>b</b>) Quantitative analysis of the assessed constructs via Western blot. (<b>c</b>,<b>d</b>) Inhibition of VLP release in cells cotransfected with Gag (0.5 μg) and varied quantities of either CHMP4B-NS3mut-green or CHMP4B-NS3-green, treated with DMSO or Glecaprevir. (<b>c</b>) Representative immunoblot experiment exhibiting (top line) HIV-1 VLP pellet, (second line) Gag HIV-1 cellular expression (WCE: whole cell extract), and (lower line) CHMP4B-NS3-green cellular expression. Hek293 cells, HeLa Kyoto (Bst2−), or HeLa CCL2 were transfected as indicated. (<b>d</b>) Quantitative assessment of the tested constructs via Western blot. HIV-1 Gag proteins were detected using an anti-p24 antibody, while CHMPs-NS3-green were detected using an Anti-Flag antibody. Note that the exposure time for Western blot revelation was typically 30 s for Hek293 cells and 5 min for HeLa CCL2 and HeLa Kyoto cells.</p>
Full article ">Figure 6
<p>Tracking time duration of Gag-mCherry spots. (<b>a</b>,<b>c</b>,<b>d</b>) Frequency distributions of spot tracking time duration in individual HeLa CCL2 cells. The cells were transfected by Gag/Gag-mCherry alone or Gag/Gag-mCherry along with Vps4A E228Q (<b>a</b>), CHMP4B-NS3-green (<b>c</b>), and CHMP2A-NS3-green (<b>d</b>), and then treated or not by Glecaprevir, as indicated. The error bars represent the standard deviation (n = 649; 231; 1874; 1669; 1142; 1527 spots from 4; 5; 8; 8; 8; 8 cells for Gag/Gag-mCherry, along with Vps4A E228Q, CHMP2A-NS3-green DMSO, CHMP2A-NS3-green Glecaprevir, CHMP4B-NS3-green DMSO, and CHMP4B-NS3-green Glecaprevir, respectively). (<b>b</b>) Frequency distribution of spot tracking time in individual HeLa Kyoto Bst2- cells. These cells were transfected with Gag/Gag-mCherry alone, Gag/Gag-mCherry with Vps4A E228Q, or Gag/Gag-mCherry with Vps4B R253A. The error bars represent the standard deviation (n = 883; 424; 1744 spots from 9; 6; 9 cells for Gag/Gag-mCherry along with Vps4A E228Q and Vps4B R253A, respectively).</p>
Full article ">Figure 7
<p>Electron microscopy images of 293FS cells showing VLP arrested in budding at the plasma membrane. (<b>a</b>–<b>e</b>) 293FS cells transfected with Gag and CHMP2A-NS3 and treated for 2 h with either DMSO (<b>a</b>) or Boceprevir (<b>b</b>–<b>e</b>). (<b>f</b>–<b>j</b>) 293FS cells transfected with Gag and CHMP4B-NS3 and treated for 2 h with either DMSO (<b>f</b>) or Boceprevir (<b>g</b>–<b>j</b>). EM images depict NS3 inhibitor action in arresting VLPs budding at the plasma membrane. Scale bars indicate 50 nm (<b>c</b>), 100 nm (<b>b</b>,<b>e</b>), and 200 nm (<b>a</b>,<b>d</b>,<b>f</b>–<b>j</b>). (<b>k</b>) Quantification of VLPs connected to the membrane per field of view in 293FS cells transfected with either CHMP4B-NS3 or CHMP2A-NS3 and subjected to treatment with DMSO or Glecaprevir as indicated (cell count is n = 3 for each condition).</p>
Full article ">
16 pages, 1742 KiB  
Article
The Circulating miRNA Profile of Chronic Hepatitis D and B Patients Is Comparable but Differs from That of Individuals with HBeAg-Negative HBV Infection
by Daniela Cavallone, Eric David B. Ornos, Gabriele Ricco, Filippo Oliveri, Barbara Coco, Piero Colombatto, Laura De Rosa, Leslie Michelle M. Dalmacio, Ferruccio Bonino and Maurizia Rossana Brunetto
Viruses 2023, 15(11), 2257; https://doi.org/10.3390/v15112257 - 15 Nov 2023
Viewed by 1167
Abstract
miRNAs circulating in whole serum and HBsAg-particles are differentially expressed in chronic hepatitis B (CHB) and HBeAg-negative-HBV infection (ENI); their profiles are unknown in chronic hepatitis D (CHD). Serum- and HBsAg-associated miRNAs were analyzed in 75 subjects of 3 well-characterized groups (CHB 25, [...] Read more.
miRNAs circulating in whole serum and HBsAg-particles are differentially expressed in chronic hepatitis B (CHB) and HBeAg-negative-HBV infection (ENI); their profiles are unknown in chronic hepatitis D (CHD). Serum- and HBsAg-associated miRNAs were analyzed in 75 subjects of 3 well-characterized groups (CHB 25, CHD 25, ENI 25) using next-generation sequencing (NGS). Overall miRNA profiles were consonant in serum and HBsAg-particles but significantly different according to the presence of hepatitis independently of Hepatitis D Virus (HDV)-co-infection. Stringent (Bonferroni Correction < 0.001) differential expression analysis showed 39 miRNAs upregulated in CHB vs. ENI and 31 of them also in CHD vs. ENI. miRNA profiles were coincident in CHB and CHD with only miR-200a-3p upregulated in CHB. Three miRNAs (miR-625-3p, miR-142-5p, and miR-223-3p) involved in immune response were upregulated in ENI. All 3 hepatocellular miRNAs of MiR-B-Index (miR-122-5p, miR-99a-5p, miR-192-5p) were overexpressed in both CHB and CHD patients. In conclusion, CHD and CHB patients showed highly similar serum miRNA profiling that was significantly different from that of individuals with HBeAg-negative infection and without liver disease. Full article
(This article belongs to the Section Human Virology and Viral Diseases)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Hierarchical clustering in HBsAg-immunoprecipitated particles. A variance-stabilized transformation was performed on the raw count matrix, and 35 genes with the highest variance across samples were selected for hierarchical clustering. Each row represents one gene, and each column represents one sample. The color represents the difference of the count value to the row mean.</p>
Full article ">Figure 2
<p>Hierarchical clustering of whole cohort study. The heatmap shows the result of the two-way hierarchical clustering of microRNAs and samples. Each row represents one gene, and each column represents one sample. The color represents the difference of the count value to the row mean.</p>
Full article ">Figure 3
<p>MA plot of differentially expressed genes (DEG). Differential expression of serum miRNAs between HBsAg carriers with HBeAg-negative infection (ENI) and CHB patients.</p>
Full article ">Figure 4
<p>MA plot of differentially expressed genes (DEG). Differential expression of serum miRNAs between HBsAg carriers with HBeAg-negative infection (ENI) vs. CHD patients.</p>
Full article ">Figure 5
<p>miRNAs were differentially expressed in individuals with HBeAg-negative infection (ENI), chronic hepatitis B (CHB), and chronic hepatitis D (CHD) patients; 31 miRNAs were downregulated in ENI when compared to both CHB and CHD. An additional 8 miRNAs were downregulated and 3 were upregulated when comparing ENI to CHB, and 3 were downregulated in the comparison between ENI and CHD.</p>
Full article ">Figure 6
<p>Box plot of the MiR-B-Index in the three groups of HBsAg-positive/HBeAg-negative individuals.</p>
Full article ">
13 pages, 1603 KiB  
Article
Identification of Host Factors for Rift Valley Fever Phlebovirus
by Velmurugan Balaraman, Sabarish V. Indran, Yonghai Li, David A. Meekins, Laxmi U. M. R. Jakkula, Heidi Liu, Micheal P. Hays, Jayme A. Souza-Neto, Natasha N. Gaudreault, Philip R. Hardwidge, William C. Wilson, Friedemann Weber and Juergen A. Richt
Viruses 2023, 15(11), 2251; https://doi.org/10.3390/v15112251 - 13 Nov 2023
Cited by 1 | Viewed by 1736
Abstract
Rift Valley fever phlebovirus (RVFV) is a zoonotic pathogen that causes Rift Valley fever (RVF) in livestock and humans. Currently, there is no licensed human vaccine or antiviral drug to control RVF. Although multiple species of animals and humans are vulnerable to RVFV [...] Read more.
Rift Valley fever phlebovirus (RVFV) is a zoonotic pathogen that causes Rift Valley fever (RVF) in livestock and humans. Currently, there is no licensed human vaccine or antiviral drug to control RVF. Although multiple species of animals and humans are vulnerable to RVFV infection, host factors affecting susceptibility are not well understood. To identify the host factors or genes essential for RVFV replication, we conducted CRISPR-Cas9 knockout screening in human A549 cells. We then validated the putative genes using siRNA-mediated knock-downs and CRISPR-Cas9-mediated knock-out studies. The role of a candidate gene in the virus replication cycle was assessed by measuring intracellular viral RNA accumulation, and the virus titers were analyzed using plaque assay or TCID50 assay. We identified approximately 900 genes with potential involvement in RVFV infection and replication. Further evaluation of the effect of six genes on viral replication using siRNA-mediated knock-downs revealed that silencing two genes (WDR7 and LRP1) significantly impaired RVFV replication. For further analysis, we focused on the WDR7 gene since the role of the LRP1 gene in RVFV replication was previously described in detail. WDR7 knockout A549 cell lines were generated and used to dissect the effect of WRD7 on a bunyavirus, RVFV, and an orthobunyavirus, La Crosse encephalitis virus (LACV). We observed significant effects of WDR7 knockout cells on both intracellular RVFV RNA levels and viral titers. At the intracellular RNA level, WRD7 affected RVFV replication at a later phase of its replication cycle (24 h) when compared with the LACV replication, which was affected in an earlier replication phase (12 h). In summary, we identified WDR7 as an essential host factor for the replication of two different viruses, RVFV and LACV, both of which belong to the Bunyavirales order. Future studies will investigate the mechanistic role through which WDR7 facilitates phlebovirus replication. Full article
(This article belongs to the Special Issue Emerging Highlights in the Study of Rift Valley Fever Virus)
Show Figures

Figure 1

Figure 1
<p>Schematics of GeCKO-A549 cells generation, selection, NGS, and data analysis. A549 cells were transduced with the lentivirus-CRISPR-Cas9 library to generate GeCKO-A549 cells. Then, the GeCKO-A549 cells were subjected to three rounds of infection with the RVFV MP-12 (1 MOI) virus. The genomic DNA of round 0 GeCKO-A549 cells, round 1, and round 3 GeCKO-A549cells were sequenced using the Illumina NextSeq 550 platform. The output NGS data were analyzed using the MaGeCK program to generate the list of genes involved in RVFV replication.</p>
Full article ">Figure 2
<p>Validation of gene hits via siRNA gene knockdown study. A549 cells were transfected with 50 nM of siRNAs. At 48 h post-transfection, the cells were infected with RVFV MP-12 virus at 0.1 MOI. At 24 h post-infection, the supernatant was collected and titered using plaque assay. NTC-non-target control siRNA; si46N-anti-RVFV siRNA; and <span class="html-italic">WDR7</span>, <span class="html-italic">SLC35B2</span>, <span class="html-italic">EXOC4</span>, <span class="html-italic">LRP1</span>, <span class="html-italic">EMC3</span>, <span class="html-italic">CT47A1</span> gene-specific siRNAs were transfected. Each bar represents the average virus titer (pfu/mL) along with the corresponding standard deviation. Statistical analysis was performed on two independent experiments with four replicates for each, using the Mann–Whitney U test and independent Student’s <span class="html-italic">t</span>-test (** <span class="html-italic">p</span>-value &lt; 0.005, *** <span class="html-italic">p</span>-value &lt; 0.001).</p>
Full article ">Figure 3
<p>Effect of <span class="html-italic">WDR7</span> gene knockout (KO) on virus production of bunyaviruses: (<b>A</b>) A549 cells CT (non-knockout control) cells and <span class="html-italic">WDR7</span> gene KO cell lines 1 and 2 were analyzed for WDR7 protein expression via Western blot using a WDR7-specific polyclonal antibody. (<b>B</b>–<b>E</b>) CT cells and WDR7 KO A549 cells were infected with RVFV MP-12 vaccine strain (<b>B</b>), with the wild-type RVFV Kenya 128B-15 strain (<b>C</b>), or with La Crosse encephalitis virus (<b>D</b>,<b>E</b>) at 0.1 MOI. The supernatant was collected at 6, 12 or 24 h post-infection (h pi) and titered using plaque assay (RVFV) or TCID<sub>50</sub>-CPE assay (LACV). RVFV MP-12 testing on A549 CT cells and WDR7 KO lines 1 or 2, involved three to five independent experiments with three to four technical replicates each. RVFV Kenya 128B-15 testing involved independent experiments with three technical replicates each. LACV testing was performed in two independent experiments with eight technical replicates each. Statistical analysis was performed using the Mann–Whitney U test and independent Student’s <span class="html-italic">t</span>-test (* <span class="html-italic">p</span>-value &lt; 0.05, ** <span class="html-italic">p</span>-value &lt; 0.005, *** <span class="html-italic">p</span>-value &lt; 0.001).</p>
Full article ">Figure 4
<p>Viral RNA accumulation at various time points post-infection in WDR7 knockout (KO) cells. CT and WDR7 KO 1 cells were infected with (<b>A</b>) RVFV MP-12 vaccine strain or (<b>B</b>) LACV, both at 0.1 MOI. Total cellular RNA was harvested at various hours post-infection (h pi). One-step RT-qPCR was performed to detect the level of viral RNA using the <span class="html-italic">PGK1</span> gene as an internal control. CT and WDR7 KO 1 cells were utilized. Each bar graph represents the average fold change in viral RNA expression, along with the corresponding standard deviation. Statistical analysis was performed on three independent experiments with two to three technical replicates for each, using the Mann–Whitney U test and independent Student’s <span class="html-italic">t</span>-test (* <span class="html-italic">p</span>-value &lt; 0.05, *** <span class="html-italic">p</span>-value &lt; 0.001, ns, non-significant).</p>
Full article ">
12 pages, 2184 KiB  
Article
Nanoluciferase Reporter Zika Viruses as Tools for Assessing Infection Kinetics and Antibody Potency
by Yanqun Xu, Devin Vertrees, Yong He, Sanaz Momben-Abolfath, Xiaohong Li, Yambasu A. Brewah, Dorothy E. Scott, Krishnamurthy Konduru, Maria Rios and Evi B. Struble
Viruses 2023, 15(11), 2190; https://doi.org/10.3390/v15112190 - 31 Oct 2023
Cited by 1 | Viewed by 1244
Abstract
Zika virus (ZIKV) has become endemic in multiple tropical and subtropical regions and has the potential to become widespread in countries with limited prior exposure to this infection. One of the most concerning sequelae of ZIKV infection is the teratogenic effect on the [...] Read more.
Zika virus (ZIKV) has become endemic in multiple tropical and subtropical regions and has the potential to become widespread in countries with limited prior exposure to this infection. One of the most concerning sequelae of ZIKV infection is the teratogenic effect on the developing fetus, with the mechanisms of viral spread to and across the placenta remaining largely unknown. Although vaccine trials and prophylactic or therapeutic treatments are being studied, there are no approved treatments or vaccines for ZIKV. Appropriate tests, including potency and in vivo assays to assess the safety and efficacy of these modalities, can greatly aid both the research of the pathophysiology of the infection and the development of anti-ZIKV therapeutics. Building on previous work, we tested reporter ZIKV variants that express nanoluciferase in cell culture and in vivo assays. We found that these variants can propagate in cells shown to be susceptible to the widely used clinical isolate PRVABC59, including Vero and human placenta cell lines. When used in neutralization assays with bioluminescence as readout, these variants gave rise to neutralization curves similar to those produced by PRVABC59, while being better suited for performing high-throughput assays. In addition, the engineered reporter variants can be useful research tools when used in other in vitro and in vivo assays, as we illustrated in transcytosis experiments and a pilot study in guinea pigs. Full article
(This article belongs to the Special Issue Arbovirus Diagnostics)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Producing an infectious clone of Zika virus (ZIKV) with teLuc, a version of the nanoluciferase gene with improved spectral properties. (<b>a</b>) Schematic representation of the genetic composition of teLuc-ZIKV infectious clone. Red vertical lines represent the position of the primers used for cloning. (<b>b</b>) Size and purity of amplicons generated during cloning. Agarose gel (1.2%) Lane 1: 1 kb DNA ladder; Lane 2: 100 bp DNA ladder; Lane 3: Amplicon 3; Lane 4: Amplicon 1; Lane5: Amplicon 2. The size and mass of the DNA ladders loaded in Lanes 1 (<b>c</b>) and 2 (<b>d</b>).</p>
Full article ">Figure 2
<p>Titration and infection kinetics of reporter ZIKV in Vero cells. (<b>a</b>) Kinetics of viral growth for nLuc-ZIKV in Vero cells at three different dilutions, 1:10<sup>3</sup> (●), 1:3 × 10<sup>3</sup> (■), and 1:10<sup>4</sup> (▲), represented as luminescence arbitrary units (AU). Means of four replicates and error bars representing standard deviations are shown. The experiment was performed once. (<b>b</b>) Titration of nLuc- and teLuc-ZIKV in Vero cells. Serial dilutions of teLuc- and nLuc-ZIKV in quadruplicates were added onto Vero cells for 48 h, and then media removed, and luciferase activity measured. Means from four repeats with error bars representing standard deviations are shown.</p>
Full article ">Figure 3
<p>A comparison of the infectivity of two nanoluciferase ZIKV reporter viruses (<b>a</b>) nLuc-ZIKV, (<b>b</b>) teLuc-ZIKV in three different cell lines (Vero, JEG3 and BeWo). Similar dilutions of each variant were added onto 75–90% confluent Vero, JEG3 or BeWo cells. Each data point represents the average of eight repeats; error bars representing standard deviations are shown.</p>
Full article ">Figure 4
<p>In vitro applications of nLuc-ZIKV reporter variant. (<b>a</b>) Neutralization of nLuc-ZIKV by monoclonal antibodies in Vero cells. Anti-ZIKV (mAb14, line) and anti-flavivirus (mAb17, broken line) monoclonal antibodies were serially diluted, mixed with nLuc-ZIKV then used to inoculate Vero cells. The infection was quantified using bioluminescence readout and the ratio with the no-antibody controls was plotted as a function of antibody concentration. Each data point represents averages from eight inoculations. Error bars representing standard deviations are shown. (<b>b</b>) Assessing the susceptibility to nLuc-ZIKV infection in cells that overexpress FcRn versus those that do not. MDCK/FcRn cells were more susceptible to nLuc-ZIKV infection than control cells that express an empty vector, recapitulating the findings seen with PRVABC59 strain [<a href="#B12-viruses-15-02190" class="html-bibr">12</a>]. Three different dilutions and up to eight replicates per dilution were used for each cell line. Differences are statistically significant by Student <span class="html-italic">t</span>-test (<span class="html-italic">p</span> = 0.036). (<b>c</b>) Evaluating transcytosis of infectious nLuc-ZIKV particles through a semipermeable membrane supporting confluent monolayers of MDCK/FcRn or control cells. Infectious viruses were quantified after transferring the contents of the basal chamber onto Vero cells (shown in the inset panel) and quantifying the infection using bioluminescence readout. Up to six replicates per experiment and cell line were used; the experiment was performed twice. The error bars are outside the lower limit of the y-axis. (<b>d</b>–<b>f</b>) Evaluating transcytosis through placenta trophoblast cells of infectious nLuc-ZIKV particles alone or as immune complexes (IC) with neutralizing antibody. Either BeWo (<b>d</b>) or Jeg-3 (<b>e</b>,<b>f</b>) cells were grown onto semi-permeable trans-well membranes and nLuc-ZIKV alone or as preformed immune complexes with 10 µg/mL anti-ZIKV antibody mAb14 was added inside the trans-well. The contents of the basolateral chamber were transferred onto Vero or BeWo cells as indicated in the inset schematic (top right corner of each panel). Addition of 10 µg/mL mAb14 reduces the infectivity in Vero or BeWo cells. Up to six replicates per experiment/cell line were used; each experiment was performed twice. In panel (<b>e</b>)/(−) IgG column, the error bars are outside the lower limit for <span class="html-italic">y</span>-axis. * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 5
<p>Use of nLuc-ZIKV for in vivo bioluminescence imaging (BLI). (<b>a</b>–<b>c</b>) Representative whole-body images of bioluminescent signals in control and inoculated guinea pigs on post-inoculation day (PID) 3. (<b>a</b>) Sham-inoculated, (<b>b</b>) nLuc-ZIKV-inoculated and infected, and (<b>c</b>) nLuc-ZIKV-inoculated but non-infected guinea pigs. Red line outlining each animal represents the region of interest. (<b>d</b>) Kinetics of bioluminescent signals in n = 4 sham and n = 6 nLuc-ZIKV-inoculated juvenile guinea pigs. Two out of six nLuc-ZIKV-inoculated guinea pigs have higher signal on PID 3, compared to either other animals or time points. (<b>e</b>) A comparison of bioluminescence on PID 3 for the two inoculated and infected animals shown on panel (<b>b</b>) with the signal from the control and other inoculated animals, the differences are statistically significant. **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">
17 pages, 2446 KiB  
Article
Pteropus vampyrus TRIM40 Is an Interferon-Stimulated Gene That Antagonizes RIG-I-like Receptors
by Sarah van Tol, Adam Hage, Ricardo Rajsbaum and Alexander N. Freiberg
Viruses 2023, 15(11), 2147; https://doi.org/10.3390/v15112147 - 25 Oct 2023
Viewed by 1405
Abstract
Nipah virus (NiV; genus: Henipavirus; family: Paramyxoviridae) naturally infects Old World fruit bats (family Pteropodidae) without causing overt disease. Conversely, NiV infection in humans and other mammals can be lethal. Comparing bat antiviral responses with those of humans may illuminate the [...] Read more.
Nipah virus (NiV; genus: Henipavirus; family: Paramyxoviridae) naturally infects Old World fruit bats (family Pteropodidae) without causing overt disease. Conversely, NiV infection in humans and other mammals can be lethal. Comparing bat antiviral responses with those of humans may illuminate the mechanisms that facilitate bats’ tolerance. Tripartite motif proteins (TRIMs), a large family of E3-ubiquitin ligases, fine-tune innate antiviral immune responses, and two human TRIMs interact with Henipavirus proteins. We hypothesize that NiV infection induces the expression of an immunosuppressive TRIM in bat, but not human cells, to promote tolerance. Here, we show that TRIM40 is an interferon-stimulated gene (ISG) in pteropodid but not human cells. Knockdown of bat TRIM40 increases gene expression of IFNβ, ISGs, and pro-inflammatory cytokines following poly(I:C) transfection. In Pteropus vampyrus, but not human cells, NiV induces TRIM40 expression within 16 h after infection, and knockdown of TRIM40 correlates with reduced NiV titers as compared to control cells. Bats may have evolved to express TRIM40 in response to viral infections to control immunopathogenesis. Full article
(This article belongs to the Special Issue TRIM Proteins in Antiviral Immunity and Virus Pathogenesis)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Pteropodid TRIM40 is an interferon-stimulated gene. (<b>A</b>,<b>B</b>) Transfection of A549 (human lung), PVK4 (<span class="html-italic">Pteropus vamyprus</span> kidney), and ZFBK13-75 (<span class="html-italic">Eidolon helvum</span> kidney) cells with 1 μg/mL of low-molecular weight (LMW) poly(I:C). (<b>A</b>) Immunoblots with antibodies specific to interferon induction regulators (pTBK1 and pSTAT1) and interferon-stimulated genes (total STAT1 and RIG-I). (<b>B</b>) At four hours post-stimulation, RNA was collected for quantitative PCR of the TRIM40 gene. Fold induction of poly(I:C)-stimulated samples relative to mock-transfected samples is shown. (<b>C</b>,<b>D</b>) A549, PVK4, and ZFBK13-75 cells were stimulated with 500 U/mL of universal type I interferon (U-IFN). (<b>C</b>) Cell lysates were collected for Western blot to probe for activated STAT1 (pSTAT1 Y701) and STAT2 (pSTAT2 Y690) as well as their total protein levels. (<b>D</b>) At four hours post-stimulation, RNA was collected for qPCR, and log<sub>2</sub> fold change is presented for canonical ISGs (Mx1, OAS1, TRIM25, and TRIM56), non-IFN-regulated TRIM23, and the gene of interest, TRIM40.</p>
Full article ">Figure 2
<p>Pteropus vampyrus TRIM40 antagonizes type I interferon and pro-inflammatory cytokine gene expression. siRNA targeting <span class="html-italic">Pteropus vampyrus</span> TRIM40 or scrambled control was transfected in PVK4 cells 48 h prior to the transfection of 1 μg/mL of low-molecular weight (LMW) poly(I:C). (<b>A</b>) The RNA collected at 24 h post-poly(I:C) transfection shows knockdown efficiency. (<b>B</b>) Lysates collected in laemmli were used for Western blot to measure activated (pSTAT2) and total STAT2. (<b>C</b>) RNA from lysates was used to measure fold-induction of ISGs (RIG-I, ISG56, Mx1, and OAS1) and pro-inflammatory cytokines (IL6 and TNFα). <span class="html-italic">p</span> values are reported as * &lt; 0.05, ** &lt; 0.01, *** &lt; 0.001, **** &lt; 0.0001.</p>
Full article ">Figure 3
<p>Pteropus vampyrus TRIM40 interacts with RIG-I-like receptors RIG-I and MDA5. Cells were co-transfected with <span class="html-italic">Pteropus vampyrus</span> HA-PV-TRIM40 and FLAG-PV-RIG-I-like receptor (RLR) RIG-I or MDA5 for 30 h in 293T (<b>A</b>) or ZFBK13-75 cells (<b>B</b>). The whole cell extracts (WCE) were used to check levels of expression and immunoprecipitated (IP) with anti-FLAG antibody beads. (<b>C</b>) 293T cells were transfected with human (Hu) or PV HA-TRIM40 and partially purified using HA peptide. Separate 293T cells were transfected with Hu- or PV-RLRs or HA-Hu-IRF3. Lysates from FLAG-protein expressing cells were immunoprecipitated with FLAG beads and washed prior to the addition of HA-purified TRIM40. (<b>D</b>) ZFBK13-75 cells were co-transfected with FLAG-PV-RIG-I and HA-PV-TRIM40 for 24 h and then stimulated with 1 μg/mL low-molecular weight (LMW) poly(I:C) for 1 or 4 h, and lysates were used for WCE or IP: FLAG. (<b>E</b>) ZFBK13-75 cells were co-transfected with FLAG-PV-MDA5 and HA-PV-TRIM40 for 24 h and then stimulated with 1 ug/mL high-molecular weight (HMW) poly(I:C) for 1 or 4 h, and lysates were used for WCE or IP: FLAG.</p>
Full article ">Figure 4
<p><span class="html-italic">Pteropus vampyrus</span> TRIM40 ubiquitinates RIG-I-like receptors RIG-I and MDA5.</p>
Full article ">Figure 5
<p>Nipah virus infection induces TRIM40 expression in <span class="html-italic">Pteropus vampyrus</span> cells. (<b>A</b>,<b>B</b>) A549 and PVK4 cells were infected with a recombinant Nipah virus (NiV) that express GFP at a multiplicity of infection of 0.05 or 1.0, and supernatants were collected throughout the 72 h post-infection period (<b>A</b>) with fluorescence microscopy images taken at 36 h (<b>B</b>). (<b>C</b>) Ephrin A and B gene expression in unstimulated PVK4 and A549 cells reported as relative expression normalized to GAPDH. (<b>D</b>) TRIM40 gene expression in A549 and PVK4 cells 16 h post-infection (h.p.i) with MOI 0.05 or 1.0 rNiV-GFP reported as fold induction (top panel), and IFNβ mRNA expression (bottom panel). <span class="html-italic">p</span> values are reported as * &lt; 0.05, ** &lt; 0.01, **** &lt; 0.0001.</p>
Full article ">Figure 6
<p>TRIM40 is proviral in bat and human cells. PVK4 cells infected with recombinant Nipah virus (rNiV) expressing GFP at a multiplicity of infection of 1.0 and measured knockdown efficiency of TRIM40 (<b>A</b>) and NiV titer (<b>B</b>) at 24 h post-infection. CRISPR-knockout or control PVK4 cells were infected with rNiV-GFP MOI 1.0 for 24 h to measure titer (<b>C</b>). A549 cells infected with rNiV-GFP at MOI 0.05 and measured knockdown efficiency of TRIM40 (qPCR) (<b>D</b>) and NiV titer (<b>E</b>) at 24 h post-infection. <span class="html-italic">p</span> values are reported as ** &lt; 0.01, *** &lt; 0.001, **** &lt; 0.0001.</p>
Full article ">Figure 7
<p>TRIM40 function and conservation. (<b>A</b>) Schematic of TRIM40′s role in regulating the RIG-I-like receptor pathway in pteropid bat and human cells. (<b>B</b>) Protein sub-domain organization of <span class="html-italic">Pteropus vampyrus</span> and human TRIM40. (<b>C</b>) Percent identity and similarity of <span class="html-italic">P. vampyrus</span> and human TRIM40.</p>
Full article ">
13 pages, 1217 KiB  
Review
Diversity and Current Classification of dsRNA Bacteriophages
by Sari Mäntynen, Meri M. Salomaa and Minna M. Poranen
Viruses 2023, 15(11), 2154; https://doi.org/10.3390/v15112154 - 25 Oct 2023
Cited by 1 | Viewed by 1621
Abstract
Half a century has passed since the discovery of Pseudomonas phage phi6, the first enveloped dsRNA bacteriophage to be isolated. It remained the sole known dsRNA phage for a quarter of a century and the only recognised member of the Cystoviridae family until [...] Read more.
Half a century has passed since the discovery of Pseudomonas phage phi6, the first enveloped dsRNA bacteriophage to be isolated. It remained the sole known dsRNA phage for a quarter of a century and the only recognised member of the Cystoviridae family until the year 2018. After the initial discovery of phi6, additional dsRNA phages have been isolated from globally distant locations and identified in metatranscriptomic datasets, suggesting that this virus type is more ubiquitous in nature than previously acknowledged. Most identified dsRNA phages infect Pseudomonas strains and utilise either pilus or lipopolysaccharide components of the host as the primary receptor. In addition to the receptor-mediated strictly lytic lifestyle, an alternative persistent infection strategy has been described for some dsRNA phages. To date, complete genome sequences of fourteen dsRNA phage isolates are available. Despite the high sequence diversity, similar sets of genes can typically be found in the genomes of dsRNA phages, suggesting shared evolutionary trajectories. This review provides a brief overview of the recognised members of the Cystoviridae virus family and related dsRNA phage isolates, outlines the current classification of dsRNA phages, and discusses their relationships with eukaryotic RNA viruses. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic presentation of phi6 virion. The size of the enveloped virion is about 85 nm [<a href="#B27-viruses-15-02154" class="html-bibr">27</a>]. The lipid-protein envelope encloses the nucleocapsid, which comprises two concentric protein shells: the nucleocapsid surface shell and the polymerase complex. Hexamers of the packaging NTPase P4 are attached on the icosahedral five-fold vertices of the polymerase complex and protrude through the nucleocapsid surface shell [<a href="#B28-viruses-15-02154" class="html-bibr">28</a>]. The spooled dsRNA genome is tightly packed inside the polymerase complex [<a href="#B29-viruses-15-02154" class="html-bibr">29</a>] together with about ten copies of the viral polymerase subunit P2 [<a href="#B30-viruses-15-02154" class="html-bibr">30</a>].</p>
Full article ">Figure 2
<p>Taxonomy of eukaryotic and bacterial dsRNA viruses currently recognised by the International Committee on Taxonomy of Viruses. Two of the five phyla under the kingdom <span class="html-italic">Orthornavirae</span> are depicted. These two phyla, <span class="html-italic">Pisuviricota</span> and <span class="html-italic">Duplornaviricota</span>, comprise all the current dsRNA virus families, except <span class="html-italic">Birnaviridae,</span> which has not been assigned to any phylum. All the three classes of <span class="html-italic">Duplornaviricota</span> (<span class="html-italic">Vidaverviricetes</span>, <span class="html-italic">Resentoviricetes</span> and <span class="html-italic">Chrymotiviricetes</span>) comprise solely dsRNA viruses, while <span class="html-italic">Pisuviricota</span> contains both ssRNA (two classes; grey boxes) and dsRNA viruses (class <span class="html-italic">Duplopiviricetes</span>). dsRNA virus taxa are in bold, dsRNA phage taxa are underlined and in blue boxes. ssRNA virus families under <span class="html-italic">Duplopiviricetes</span> are in grey font.</p>
Full article ">
11 pages, 1559 KiB  
Brief Report
A Triple Gene-Deleted Pseudorabies Virus-Vectored Subunit PCV2b and CSFV Vaccine Protect Pigs against a Virulent CSFV Challenge
by Ediane Silva, Elizabeth Medina-Ramirez, Selvaraj Pavulraj, Douglas P. Gladue, Manuel Borca and Shafiqul I. Chowdhury
Viruses 2023, 15(11), 2143; https://doi.org/10.3390/v15112143 - 25 Oct 2023
Viewed by 1180
Abstract
Classical swine fever (CSF) remains one of the most economically significant viral diseases affecting domestic pigs and wild boars worldwide. To develop a safe and effective vaccine against CSF, we have constructed a triple gene-deleted pseudorabies virus (PRVtmv)-vectored bivalent subunit vaccine against porcine [...] Read more.
Classical swine fever (CSF) remains one of the most economically significant viral diseases affecting domestic pigs and wild boars worldwide. To develop a safe and effective vaccine against CSF, we have constructed a triple gene-deleted pseudorabies virus (PRVtmv)-vectored bivalent subunit vaccine against porcine circovirus type 2b (PCV2b) and CSFV (PRVtmv+). In this study, we determined the protective efficacy of the PRVtmv+ against virulent CSFV challenge in pigs. The results revealed that the sham-vaccinated control group pigs developed severe CSFV-specific clinical signs characterized by pyrexia and diarrhea, and became moribund on or before the seventh day post challenge (dpc). However, the PRVtmv+-vaccinated pigs survived until the day of euthanasia at 21 dpc. A few vaccinated pigs showed transient diarrhea but recovered within a day or two. One pig had a low-grade fever for a day but recovered. The sham-vaccinated control group pigs had a high level of viremia, severe lymphocytopenia, and thrombocytopenia. In contrast, the vaccinated pigs had a low–moderate degree of lymphocytopenia and thrombocytopenia on four dpc, but recovered by seven dpc. Based on the gross pathology, none of the vaccinated pigs had any CSFV-specific lesions. Therefore, our results demonstrated that the PRVtmv+ vaccinated pigs are protected against virulent CSFV challenge. Full article
(This article belongs to the Special Issue Strategies for Preventing Viral Diseases of Domestic Animals)
Show Figures

Figure 1

Figure 1
<p>PRVtmv+ immunization in pigs. (<b>A</b>) Schematic showing the PRVtmv+ vaccination, sample collection, and CSFV challenge scheme for the animal experiment. (<b>B</b>) Pigs (<span class="html-italic">n</span> = 5 in each group) were immunized with PRVtmv+ vaccine (for each pig, intranasally—8 × 10<sup>7</sup> plaque forming units (PFU), and subcutaneously—4 × 10<sup>7</sup> PFU; 0 days post vaccination (dpv)), or sham-vaccinated and sera samples were collected on 0 and 28 dpv to determine the classical swine fever virus (CSFV)-specific neutralizing antibody titer [<a href="#B33-viruses-15-02143" class="html-bibr">33</a>]. The dot plot graph shows each animal’s individual CSFV titer with the mean values of the group (<span class="html-italic">n</span> = 5). *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 2
<p>Clinical assessment, survival curve, and classical swine fever virus (CSFV)-associated viremia in sham-vaccinated control and PRVtmv+ vaccinated pigs following CSFV challenge. (<b>A</b>) Pigs were challenged with CSFV at 28 days post vaccination (0 days post challenge (dpc)), and rectal temperature was recorded until 21 dpc. Mean rectal temperatures with standard deviation (SD) are given (<span class="html-italic">n</span> = 5). All pigs in the CSFV control group were euthanized on days 6 and 7 dpc. (<b>B</b>) A survival rate is given in terms of percentage in each group (<span class="html-italic">n</span> = 5). (<b>C</b>) Blood samples were collected from pigs on days 0, 4, 7, 14, and 21 post challenge, and the CSFV titer was determined in cell culture. The mean CSFV titer in the blood of each animal from both groups is shown. The dot plot graph represents mean + individual values in each group (<span class="html-italic">n</span> = 5). TCID50—50% tissue culture infectious dose; * <span class="html-italic">p</span> &lt; 0.05 and *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 3
<p>Percent changes in leukocyte, lymphocyte, and platelet count following the classical swine fever virus (CSFV) Brescia strain (BICv) challenge in both groups. Whole blood was collected from pigs on 28 dpv/0 dpc and 49 dpv/21 dpc, (<b>A</b>) leukocyte, (<b>B</b>) lymphocyte, and (<b>C</b>) platelet counts were determined, and their percent changes were calculated. The normal range in pigs: (i) leukocytes—11–22 × 10<sup>3</sup>/μL; (ii) lymphocytes—4.6–12 × 10<sup>3</sup>/μL; (iii) platelets—200–500 × 10<sup>3</sup>/μL [<a href="#B34-viruses-15-02143" class="html-bibr">34</a>].</p>
Full article ">Figure 4
<p>Photograph of spleen, kidney, and tonsils of PRVtmv+ vaccinated pigs (#84, #87, #88, #91, and #92) euthanized at 21 days post challenge. Spleen (<b>A</b>–<b>E</b>), kidney (<b>F</b>–<b>J</b>), and tonsils (<b>K</b>–<b>O</b>). The spleen, kidney, and tonsils lack CSFV-specific lesions. The focal areas marked with circles in D most likely represent areas of blood pooling/incomplete extrusion of blood; this is sometimes seen in the spleens of various species during postmortems after euthanasia using barbiturate compounds. Splenic infarctions, even red-type infarcts, are usually sharply demarcated from the rest of the parenchyma.</p>
Full article ">
31 pages, 1731 KiB  
Review
Polyomavirus Wakes Up and Chooses Neurovirulence
by Arrienne B. Butic, Samantha A. Spencer, Shareef K. Shaheen and Aron E. Lukacher
Viruses 2023, 15(10), 2112; https://doi.org/10.3390/v15102112 - 18 Oct 2023
Viewed by 2484
Abstract
JC polyomavirus (JCPyV) is a human-specific polyomavirus that establishes a silent lifelong infection in multiple peripheral organs, predominantly those of the urinary tract, of immunocompetent individuals. In immunocompromised settings, however, JCPyV can infiltrate the central nervous system (CNS), where it causes several encephalopathies [...] Read more.
JC polyomavirus (JCPyV) is a human-specific polyomavirus that establishes a silent lifelong infection in multiple peripheral organs, predominantly those of the urinary tract, of immunocompetent individuals. In immunocompromised settings, however, JCPyV can infiltrate the central nervous system (CNS), where it causes several encephalopathies of high morbidity and mortality. JCPyV-induced progressive multifocal leukoencephalopathy (PML), a devastating demyelinating brain disease, was an AIDS-defining illness before antiretroviral therapy that has “reemerged” as a complication of immunomodulating and chemotherapeutic agents. No effective anti-polyomavirus therapeutics are currently available. How depressed immune status sets the stage for JCPyV resurgence in the urinary tract, how the virus evades pre-existing antiviral antibodies to become viremic, and where/how it enters the CNS are incompletely understood. Addressing these questions requires a tractable animal model of JCPyV CNS infection. Although no animal model can replicate all aspects of any human disease, mouse polyomavirus (MuPyV) in mice and JCPyV in humans share key features of peripheral and CNS infection and antiviral immunity. In this review, we discuss the evidence suggesting how JCPyV migrates from the periphery to the CNS, innate and adaptive immune responses to polyomavirus infection, and how the MuPyV-mouse model provides insights into the pathogenesis of JCPyV CNS disease. Full article
(This article belongs to the Special Issue Neurotropic Viral Pathogens)
Show Figures

Figure 1

Figure 1
<p>Comparison of JCPyV and MuPyV genomes and how each virus enters its host cell. Both viruses initially attach to a receptor containing a sialic acid. Entrance is then facilitated by a secondary receptor. Both viruses are then internalized via different means. While JCPyV is internalized via clathrin-mediated endocytosis, MuPyV is internalized via caveolin-mediated endocytosis. The viruses then enter the early endosomes and transition into late endosomes. Both MuPyV and JCPyV are transmitted via similar routes and establish infections in similar tissues. Made in BioRender.</p>
Full article ">Figure 2
<p>Hypothesized infection routes through the three barriers of the CNS: the BBB, the ChP-CSF-ependyma barrier, and the meninges. Generally, for all three barriers, infection may occur directly via free virions or through indirect means like EVs or infected leukocytes. Made in BioRender.</p>
Full article ">
14 pages, 2581 KiB  
Article
Small Heat Shock Protein (sHsp22.98) from Trialeurodes vaporariorum Plays Important Role in Apple Scar Skin Viroid Transmission
by Savita Chaudhary, Vijayanandraj Selvaraj, Preshika Awasthi, Swati Bhuria, Rituraj Purohit, Surender Kumar and Vipin Hallan
Viruses 2023, 15(10), 2069; https://doi.org/10.3390/v15102069 - 9 Oct 2023
Cited by 1 | Viewed by 3617
Abstract
Trialeurodes vaporariorum, commonly known as the greenhouse whitefly, severely infests important crops and serves as a vector for apple scar skin viroid (ASSVd). This vector-mediated transmission may cause the spread of infection to other herbaceous crops. For effective management of ASSVd, it is [...] Read more.
Trialeurodes vaporariorum, commonly known as the greenhouse whitefly, severely infests important crops and serves as a vector for apple scar skin viroid (ASSVd). This vector-mediated transmission may cause the spread of infection to other herbaceous crops. For effective management of ASSVd, it is important to explore the whitefly’s proteins, which interact with ASSVd RNA and are thereby involved in its transmission. In this study, it was found that a small heat shock protein (sHsp) from T. vaporariorum, which is expressed under stress, binds to ASSVd RNA. The sHsp gene is 606 bp in length and encodes for 202 amino acids, with a molecular weight of 22.98 kDa and an isoelectric point of 8.95. Intermolecular interaction was confirmed through in silico analysis, using electrophoretic mobility shift assays (EMSAs) and northwestern assays. The sHsp22.98 protein was found to exist in both monomeric and dimeric forms, and both forms showed strong binding to ASSVd RNA. To investigate the role of sHsp22.98 during ASSVd infection, transient silencing of sHsp22.98 was conducted, using a tobacco rattle virus (TRV)-based virus-induced gene silencing system. The sHsp22.98-silenced whiteflies showed an approximate 50% decrease in ASSVd transmission. These results suggest that sHsp22.98 from T. vaporariorum is associated with viroid RNA and plays a significant role in transmission. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Heterologous expression and purification of sHsp22.98 from <span class="html-italic">T. vaporariorum</span>. (<b>A</b>) Expression and purification of sHsp22.98 in <span class="html-italic">E. coli</span> BL21 (DE3) cells (lane 1). (<b>B</b>) Detection of monomeric and dimeric forms of purified sHsp22.98 by Western blotting using sHsp22.98-specific antisera (lane 1). (<b>C</b>) Dimerization of recombinant sHsp22.98 after storage. Lane 1: purified sHsp22.98 protein fractionated on SDS-PAGE. Lane M: protein ladder (3-color prestained protein ladder, 10–250 kDa, Genetix).</p>
Full article ">Figure 2
<p>In vitro interaction of sHsp22.98 with ASSVd RNA using electrophoretic mobility shift assay (EMSA) and northwestern blot. (<b>A</b>) EMSA: RNA-protein complex was fractionated on 2.5% agarose gel. Lane 1: 1 µg ASSVd RNA; lane 2: 800 ng protein; lane 3: ASSVd RNA+ 50 ng protein; lane 4: ASSVd RNA + 100 ng protein; lane 5: ASSVd RNA + 200 ng protein; lane 6: ASSVd RNA + 400 ng protein; lane 7: ASSVd RNA + 800 ng protein. (<b>B</b>) In vitro interaction analysis of sHsp22.98 protein with digoxigenin-labeled ASSVd probe using northwestern blotting. Lane M: prestained protein ladder (Genetix). Lane 1: 10 µg of purified protein, lane 2: 5 µg of purified protein, lane 3: BSA as negative control.</p>
Full article ">Figure 3
<p>Molecular modeling and prediction of binding sites in ASSVd and <span class="html-italic">T. vaporariorum</span> small heat shock protein22.98 (sHsp22.98). (<b>A</b>) Three-dimensional structure of ASSVd RNA showing three protein-binding sites. Purple: BS1; magenta: BS2; red: BS3. (<b>B</b>) Structure of sHsp22.98 monomer showing the presence of alpha-crystalline domain in yellow. (<b>C</b>) Three-dimensional structure of sHsp22.98 monomer protein amino acids showing predicted RNA-binding sites in circular dots. The ASSVd RNA and sHsp protein were modeled using Discovery Studio software package.</p>
Full article ">Figure 4
<p>Dynamic association of sHsp22.98 predicted protein-binding sites to ASSVd<b>.</b> The predicted interacting position of ASSVd and sHsp22.98 are given in brackets. The free energy of the interaction between ASSVd and sHsp22.98 is listed as a score. The ASSVd–sHsp22.98 interaction was carried out by adopting RNA homology modeling protocol built in Discovery Studio software package. (<b>A</b>) sHsp22.98–site3 and RNA–BS1 complex (51–97), with a score of −159.30. (<b>B</b>) sHsp22.98–site3 and RNA–BS2 complex (201–250), with a score of −132.60. (<b>C</b>) sHsp22.98–site3 and RNA–BS3 complex (251–300), with a score of −121.40.</p>
Full article ">Figure 5
<p>Dynamic association of sHsp22.98 dimer with ASSVd RNA. (<b>A</b>) The predicted dimeric form of sHsp22.98. (<b>B</b>) The predicted interacting position of ASSVd and dimeric form of sHsp22.98. The ASSVd-sHsp22.98 interaction was carried out by adopting RNA homology modeling protocol built in Discovery Studio software package.</p>
Full article ">Figure 6
<p>Quantification of <span class="html-italic">sHsp22.98</span> mRNA and effect of ASSVd transmission by sHsp22.98-silenced whiteflies. (<b>A</b>) Quantification of target <span class="html-italic">sHsp22.98</span> mRNA using semiquantitative PCR from control (lane 1 and lane 2) and <span class="html-italic">sHsp22.98</span>-silenced whiteflies (lanes 3 and 4); lane 5: non-template control. <span class="html-italic">mtCoI</span> gene was used as internal control. Amplicon intensity was quantified using Uvitech software, and healthy whiteflies (lane 1) were used as the control for quantification. Values represent the percent expression. (<b>B</b>) Transmission of ASSVd by <span class="html-italic">sHsp22.98</span>-silenced whiteflies.</p>
Full article ">
20 pages, 12369 KiB  
Article
Structure of Vibrio Phage XM1, a Simple Contractile DNA Injection Machine
by Zhiqing Wang, Andrei Fokine, Xinwu Guo, Wen Jiang, Michael G. Rossmann, Richard J. Kuhn, Zhu-Hua Luo and Thomas Klose
Viruses 2023, 15(8), 1673; https://doi.org/10.3390/v15081673 - 31 Jul 2023
Cited by 3 | Viewed by 1520
Abstract
Antibiotic resistance poses a growing risk to public health, requiring new tools to combat pathogenic bacteria. Contractile injection systems, including bacteriophage tails, pyocins, and bacterial type VI secretion systems, can efficiently penetrate cell envelopes and become potential antibacterial agents. Bacteriophage XM1 is a [...] Read more.
Antibiotic resistance poses a growing risk to public health, requiring new tools to combat pathogenic bacteria. Contractile injection systems, including bacteriophage tails, pyocins, and bacterial type VI secretion systems, can efficiently penetrate cell envelopes and become potential antibacterial agents. Bacteriophage XM1 is a dsDNA virus belonging to the Myoviridae family and infecting Vibrio bacteria. The XM1 virion, made of 18 different proteins, consists of an icosahedral head and a contractile tail, terminated with a baseplate. Here, we report cryo-EM reconstructions of all components of the XM1 virion and describe the atomic structures of 14 XM1 proteins. The XM1 baseplate is composed of a central hub surrounded by six wedge modules to which twelve spikes are attached. The XM1 tail contains a fewer number of smaller proteins compared to other reported phage baseplates, depicting the minimum requirements for building an effective cell-envelope-penetrating machine. We describe the tail sheath structure in the pre-infection and post-infection states and its conformational changes during infection. In addition, we report, for the first time, the in situ structure of the phage neck region to near-atomic resolution. Based on these structures, we propose mechanisms of virus assembly and infection. Full article
(This article belongs to the Special Issue Phage Structural Biology)
Show Figures

Figure 1

Figure 1
<p>Overview of phage XM1 structure. (<b>a</b>) Cryo-EM reconstructions of the icosahedral capsid and the 6-fold-symmetric tail were joined together to generate a model of the entire XM1 virion. (<b>b</b>) The neck region of the XM1 virion. The left side shows a schematic diagram representing the neck proteins. The right side shows the backbones of the proteins (color-coded as noted below the panel) in the cryo-EM map of the neck region (grey). (<b>c</b>) Structure of the XM1 baseplate. The left side shows a schematic diagram representing the baseplate proteins. The right side shows the backbones of the proteins (color-coded as noted below the panel) in the cryo-EM map of the baseplate region (grey). Protein color codes and their functions are provided at the bottom of the figure.</p>
Full article ">Figure 2
<p>Structures of the tail proteins and the proposed baseplate assembly pathway. (<b>a</b>) The ribbon diagrams of gp6 (forest green), gp7 (green), gp11 (red), gp12 (magenta), gp15 (yellow), gp16 (blue), and gp17 (cyan). The trifurcation region of (gp16)<sub>2</sub>–gp17 is outlined with a dashed circle. (<b>b</b>) The asymmetric unit of the baseplate is shown in two orientations, rotated by 90° relative to each other. The sheath initiator protein, gp15, indicated by an arrow in the top orientation, attaches to gp11 and interacts with wedge proteins and the sheath protein, gp6. The gp12 baseplate stabilization protein, indicated by an arrow in the lower image, strongly interacts with gp11 and gp16. (<b>c</b>) Proposed assembly pathway of the XM1 baseplate. The grey density, segmented from the 6-fold-symmetric map of the full tail, represents the hub protein, gp13, and the cell-puncturing protein, gp14. (<b>i</b>) The central baseplate organizing protein, gp11, assembles on top of the hub protein, with gp12 subunits attached to it. (<b>ii</b>) The (gp16)<sub>2</sub>–gp17 complex is bound to the central hub by interactions with gp11 and gp12. (<b>iii</b>) The six baseplate wedges are then fastened together by dimerization of the gp16 subunits from neighboring wedges. (<b>iv</b>) The first hexameric ring of the tail tube protein, gp7, assembles on top of the gp11 ring. (<b>v</b>) The sheath initiator protein, gp15, attaches to the baseplate, and the first hexameric ring of sheath protein, gp6, assembles on the top of the baseplate. (<b>vi</b>) The density (gold), segmented from the map of the full tail, corresponds to the tail spike protein, gp18, which attaches to the baseplate periphery.</p>
Full article ">Figure 3
<p>Structure of tail sheath protein, gp6, in the extended and contracted sheaths. (<b>a</b>) Top and side views of one hexameric gp6 ring of the extended sheath. The gp6 subunits are colored in magenta and yellow alternately to show the “handshake” between the N- and C-termini of two adjacent subunits. (<b>b</b>) Top and side views of two hexameric gp6 rings of the extended sheath. The upper ring is colored as in panel A and the lower ring is shown in green. (<b>c</b>) Enlargement showing that the “handshake” β-strands from the upper sheath ring form an augmented β-sheet with β-strands of the C-terminal domain (CTD) from the lower ring. The augmented β-sheet is indicated by an arrow. (<b>d</b>) Side view of a single gp6 subunit of the extended sheath. This view represents the “standing” orientation of the sheath subunit. The CTD, NTD, and InD domains of gp6 are outlined by squares. (<b>e</b>) One hexameric ring of the sheath in the contracted state. The gp6 subunits are colored as in Panel A to show that the “handshake” between adjacent subunits is conserved after sheath contraction. In the contracted state, the sheath ring is expanded when viewed from the top and compacted when viewed from the side. (<b>f</b>) Two hexameric rings of the contracted sheath with the lower ring shown in forest green. (<b>g</b>) Enlargement (circled) showing that the β-sheet augmentation is conserved in the contracted sheath. (<b>h</b>) Side view of one gp6 subunit of the contracted sheath. This view represents the “lying” orientation of the sheath subunit.</p>
Full article ">Figure 4
<p>The contracted sheath structure and its interactions with tail terminator protein, gp5, and sheath initiator protein, gp15. (<b>a</b>) The 6-fold-symmetric reconstruction of the XM1 tail calculated using phage particles with contracted tail sheaths and empty capsids. The density is colored according to the distance from the central vertical 6-fold axis (see color bar). (<b>b</b>) The sheath protein, gp6, in the contracted state (blue) is superimposed onto the sheath protein in the extended state (forest green). Note that the N-terminus shows the most conformational changes as pointed out by arrows. (<b>c</b>) The sheath initiator protein, gp15 (yellow and circled), is aligned with the C-terminal domain (CTD) of the sheath protein, gp6 (forest green). (<b>d</b>) The tail terminator protein, gp5 (purple), and the sheath subunits, gp6 (forest green), fitted into the map of the contracted tail. The gp5–gp6 interactions are outlined by the circle. After sheath contraction, the gp6 subunits change their positions relative to gp5. (<b>e</b>) Sheath protein subunits fitted into the reconstruction of contracted tails. Densities representing the augmented β-sheets are encircled. (<b>f</b>) Sheath initiator protein subunits, gp15 (yellow and circled), fitted into the density next to the first ring of the contracted sheath.</p>
Full article ">Figure 5
<p>Structure of the XM1 neck. (<b>a</b>) The portal protein, gp49, is colored orange. The head completion proteins, gp1 and gp4, are colored light blue and brown, respectively. The tail terminator protein, gp5, is colored purple. The tail sheath protein, gp6, is forest green, and the tail tube protein, gp7, is green. The collar spike protein, gp40, is dark blue. (<b>b</b>) The top and side views of the portal protein, gp49. In the side view, one of the chains of the gp49 dodecamer is colored cyan to show that the N-terminal region of the protein wraps around four neighboring subunits. (<b>c</b>) Top view of the dodecameric gp1 ring and side view of one gp1 subunit. The β-sheet insertion, facing the DNA helix, is outlined by a circle, and the positions of five serine residues are labeled. (<b>d</b>) A complex of the gp4, gp5, and gp40 rings, viewed from the side, is shown at the top of the panel. The gp40 protein reinforces the head–tail interactions. One gp40 trimer, interacting with one gp4 subunit and one gp5 subunit, is shown at the bottom of the panel. (<b>e</b>) Side view of the hexameric gp4 ring and of one gp4 subunit. The extended loop between β2 and β3 is circled and the three lysine residues belonging to this loop are labeled. (<b>f</b>) A top view of the rings of the tail terminator protein, gp5, and the sheath protein, gp6, is shown on the left. A side view of one gp5 subunit, and one gp6 subunit is shown on the right. The C-terminus of the tail terminator protein, gp5, extends and forms a β-sheet with β-strands of the C-terminal domain of the gp6 sheath protein. The augmented β-sheet is outlined by the circle.</p>
Full article ">Figure 6
<p>Structure of the XM1 capsid. (<b>a</b>) Cryo-EM reconstruction of the icosahedral XM1 capsid. The density is colored according to the distance from the capsid center (see the color bar). (<b>b</b>) One asymmetric unit of the capsid. The major capsid protein, gp54, is colored yellow, and the decoration protein, gp53, is colored red. (<b>c</b>) The XM1 gp54 (yellow) superimposed onto the HK97 major capsid protein (magenta). (<b>d</b>) Side and top views of the XM1 decoration protein (red) superimposed onto the phage TW1 decoration protein (blue). (<b>e</b>) Close-up view of the cryo-EM map (grey mesh) near the center of the hexameric gp54 capsomer. The arrow points to the unidentified density in the center of the capsomer.</p>
Full article ">
13 pages, 2787 KiB  
Article
Apis mellifera Solinvivirus-1, a Novel Honey Bee Virus That Remained Undetected for over a Decade, Is Widespread in the USA
by Eugene V. Ryabov, Anthony J. Nearman, Ashrafun Nessa, Kyle Grubbs, Benjamin Sallmann, Rachel Fahey, Mikayla E. Wilson, Karen D. Rennich, Nathalie Steinhauer, Anne Marie Fauvel, Yanping Chen, Jay D. Evans and Dennis vanEngelsdorp
Viruses 2023, 15(7), 1597; https://doi.org/10.3390/v15071597 - 21 Jul 2023
Cited by 2 | Viewed by 3518
Abstract
A metagenomic analysis of the virome of honey bees (Apis mellifera) from an apiary with high rates of unexplained colony losses identified a novel RNA virus. The virus, which was named Apis mellifera solinvivirus 1 (AmSV1), contains a 10.6 kb positive-strand [...] Read more.
A metagenomic analysis of the virome of honey bees (Apis mellifera) from an apiary with high rates of unexplained colony losses identified a novel RNA virus. The virus, which was named Apis mellifera solinvivirus 1 (AmSV1), contains a 10.6 kb positive-strand genomic RNA with a single ORF coding for a polyprotein with the protease, helicase, and RNA-dependent RNA polymerase domains, as well as a single jelly-roll structural protein domain, showing highest similarity with viruses in the family Solinviviridae. The injection of honey bee pupae with AmSV1 preparation showed an increase in virus titer and the accumulation of the negative-strand of AmSV1 RNA 3 days after injection, indicating the replication of AmSV1. In the infected worker bees, AmSV1 was present in heads, thoraxes, and abdomens, indicating that this virus causes systemic infection. An analysis of the geographic and historic distribution of AmSV1, using over 900 apiary samples collected across the United States, showed AmSV1 presence since at least 2010. In the year 2021, AmSV1 was detected in 10.45% of apiaries (95%CI: 8.41–12.79%), mostly sampled in June and July in Northwestern and Northeastern United States. The diagnostic methods and information on the AmSV1 distribution will be used to investigate the connection of AmSV1 to honey bee colony losses. Full article
(This article belongs to the Section Insect Viruses)
Show Figures

Figure 1

Figure 1
<p>The organization of <span class="html-italic">Apis mellifera</span> solinvivirus-1 genomic RNA. (<b>a</b>) The schematic representation of AmSV1 genomic RNA (GenBank accession number OQ540582). The position of the main ORF and putative protein domains is shown. Amino acid positions of the protein domains are indicated above the ORF. Non-structural proteins: Hel—helicase, Prot—3C protease, structural proteins: JR—jelly roll domain of structural viral protein (VP)1, VP2; (<b>b</b>) NGS coverage of the AmSV1 genome. Genetic diversity of AmSV1 apiary-level population used for NGS analysis: (<b>c</b>) Shannon’s diversity profile (sliding window average for 100 nt positions); (<b>d</b>) distribution of polymorphic nucleotides (n = 419) and amino acids (n = 63), showing an alternate allele exceeding 3% in frequency in the apiary-level NGS library. (<b>e</b>) The maximum likelihood phylogenetic tree was generated based on the full-length protein sequences of AmSV1, as well as classified (marked with asterisk) and putative solinviviruses. For SINV3 and RAAV, the sequences of -1 translational frameshift proteins (ORF1-ORF2 fusions) were used. Bootstrap values above 50%, generated from 1000 replications, are shown to the left of corresponding nodes. The bar indicates a 10% sequence difference.</p>
Full article ">Figure 2
<p>Accumulation of AmSV1 in different body parts of individual adult honeybee workers from AmSV1 positive apiaries. (<b>a</b>) Sections of a frozen worker honey bee used for RNA extraction. (<b>b</b>) AmSV1 loads in the head, thorax, and abdomen of 16 worker bees. (<b>c</b>) Correlation between AmSV1 loads in head, thorax, and abdomen.</p>
Full article ">Figure 3
<p>Pupal injection experiment. (<b>a</b>) The quantification of AmSV1 genome equivalents (GE) in the pupae, injected with partially purified AmSV1 preparation (AmSV1) or buffer control (PBS), which were sampled immediately after injection (Time 0) and after 3 days of incubation at +33 °C (3 dpi). Quantification was carried out by qPCR using cDNA-generated random primers, allowing the detection of AmSV1 RNA of both polarities. Dots indicate levels of AmSV1 in individual pupae. Significantly different levels of AmSV1 RNA are indicated by different red letters (ANOVA <span class="html-italic">p</span> &lt; 0.01). (<b>b</b>) Specific detection of negative-strand RNA, a virus replicative intermediate, in virus preparation (lane 0) used for injection. The pupae was injected with partially purified AmSV1 preparation (AmSV1) and sampled immediately after injection, Time 0 (lanes 2 and 3, individual pupae), or after 3 day incubation at +33 °C, with 3 dpi (lanes 6 and 7, two pools of 2 pupae). Control, buffer-injected pupae—PBS—were sampled 3 days after injection (lanes 4 and 5, two pools of two pupae). M, DNA ladder, base pairs (bp). The cDNA was produced using tagged forward primer, PCR amplification was carried out with the primer that was identical to the tag and reverse primer. The arrow marks the position of the expected 141 bp RT-PCR product.</p>
Full article ">Figure 4
<p>Spatio-temporal distribution of AmSV1 in the USA. (<b>a</b>) The distribution of AmSV1 in the US apiaries in 2010, 2014, and 2021. (<b>b</b>) Monthly loads (0 = not detected, below 2.8 log<sub>10</sub> GE/bee) and (<b>c</b>) monthly distribution of AmSV1 prevalence and loads for the year 2021.</p>
Full article ">Figure 5
<p>Connections between the prevalence of AmSV1 and (<b>a</b>) other honey bee parasites, including (<b>b</b>) field apiary observations in the U.S. 2021 apiaries. In total, 794 apiary-level samples were tested. Dashed line at OR = 1 represents the null hypothesis that AmSV1 does not associate with listed measures. DWV–A—deformed wing virus, type A; DWV-B deformed wing virus type B; ABPV—acute bee paralysis virus; CBPV—chronic bee paralysis virus; IAPV—Israeli acute paralysis virus; KBV—Kashmir bee virus; LSV2—Lake Sinai virus; Varroa—ectoparasitic mite <span class="html-italic">Varroa destructor</span>; Nosema—<span class="html-italic">Vairimorpha ceranae</span>. EFB—European foulbrood (caused by bacterium <span class="html-italic">Melissococcus plutonius</span>), Sacbrood–caused by sacbrood virus, Chalkbrood–fungal disease of honey bee brood caused by fungus <span class="html-italic">Ascosphaera apis</span>, PMS—Parasitic Mite Syndrome (caused by the mite <span class="html-italic">Varroa destructor</span>), Deformed Wings–could be caused by DWV, Shiny Black—hairless bees, SHB—infestation with small hive beetle (<span class="html-italic">Aethina tumida</span>), Wax Moth—infestation with wax moth (<span class="html-italic">Galleria mellonella</span>), Queen Cells presence, Drone Layer–queen lays unfertilized drone eggs, Queenless—queen is absent in at least one of sampled colonies, Any Queen Issues—combined Queen Cells, Drone Layer, and Queenless. A significant <span class="html-italic">p</span>-value is marked with an asterisk.</p>
Full article ">
19 pages, 6114 KiB  
Article
Structure and Function of Hoc—A Novel Environment Sensing Device Encoded by T4 and Other Bacteriophages
by Andrei Fokine, Mohammad Zahidul Islam, Qianglin Fang, Zhenguo Chen, Lei Sun and Venigalla B. Rao
Viruses 2023, 15(7), 1517; https://doi.org/10.3390/v15071517 - 7 Jul 2023
Cited by 3 | Viewed by 2402
Abstract
Bacteriophage T4 is decorated with 155 180 Å-long fibers of the highly antigenic outer capsid protein (Hoc). In this study, we describe a near-atomic structural model of Hoc by combining cryo-electron microscopy and AlphaFold structure predictions. It consists of a conserved C-terminal capsid-binding [...] Read more.
Bacteriophage T4 is decorated with 155 180 Å-long fibers of the highly antigenic outer capsid protein (Hoc). In this study, we describe a near-atomic structural model of Hoc by combining cryo-electron microscopy and AlphaFold structure predictions. It consists of a conserved C-terminal capsid-binding domain attached to a string of three variable immunoglobulin (Ig)-like domains, an architecture well-preserved in hundreds of Hoc molecules found in phage genomes. Each T4-Hoc fiber attaches randomly to the center of gp23* hexameric capsomers in one of the six possible orientations, though at the vertex-proximal hexamers that deviate from 6-fold symmetry, Hoc binds in two preferred orientations related by 180° rotation. Remarkably, each Hoc fiber binds to all six subunits of the capsomer, though the interactions are greatest with three of the subunits, resulting in the off-centered attachment of the C-domain. Biochemical analyses suggest that the acidic Hoc fiber (pI, ~4–5) allows for the clustering of virions in acidic pH and dispersion in neutral/alkaline pH. Hoc appears to have evolved as a sensing device that allows the phage to navigate its movements through reversible clustering–dispersion transitions so that it reaches its destination, the host bacterium, and persists in various ecological niches such as the human/mammalian gut. Full article
(This article belongs to the Special Issue Bacteriophage Bioinformatics)
Show Figures

Figure 1

Figure 1
<p>Molecular architecture of the T4 head. (<b>A</b>) The native prolate T4 capsid. (<b>B</b>) The isometric mutant capsid. The major capsid protein gp23* shell surface is shown in cyan; gp24* vertices are shown in magenta; Soc subunits are shown in green; Hoc fibers are shown in orange. The gp23*, gp24*, and Soc structures were determined in the previous cryo-EM studies of the isometric [<a href="#B7-viruses-15-01517" class="html-bibr">7</a>] and prolate capsids [<a href="#B2-viruses-15-01517" class="html-bibr">2</a>], whereas the model of the full-length Hoc has been generated for this study. Panels (<b>C</b>,<b>D</b>) show one pentameric gp24* vertex, one vertex-proximal gp23* hexamer, Soc molecules surrounding gp23*, and Hoc in the center of the gp23* hexamer in surface view (<b>C</b>) or in ribbon diagram (<b>D</b>).</p>
Full article ">Figure 2
<p>Structural model of the full-length T4 Hoc protein. (<b>A</b>) Model of T4 Hoc generated using AlphaFold. The residue colors are based on the pLDDT values [<a href="#B33-viruses-15-01517" class="html-bibr">33</a>]. The colors range from blue, corresponding to the pLDDT value of 100 (the highest confidence), to red, corresponding to the pLDDT value of 0 (the lowest confidence). (<b>B</b>) The protein chain is rainbow colored from the N-terminus (blue) to the C-terminus (red). (<b>C</b>) Molecular surface colored according to electrostatic potential showing the acidic nature of the Hoc protein. The surface color ranges from red, corresponding to a potential of −5 kT/e<sup>−</sup>, to blue, corresponding to a potential of +5 kT/e<sup>−</sup>. The potential was calculated assuming 0 M concentrations for the +1 and −1 ion species.</p>
Full article ">Figure 3
<p>Hoc orientations on gp23* capsomers. (<b>A</b>) Schematic representation of three gp23* hexameric capsomers. The six possible orientations of Hoc in the center of each capsomer is depicted by arrows. (<b>B</b>) Schematic representation showing the gp24* vertex pentamer as a pentagon and an adjacent gp23* hexamer as a hexagon. The gp23* subunits within the hexamer are shown in different colors. The two preferred orientations of Hoc in the hexamer center are shown by arrows. (<b>C</b>) Surface of the gp24* pentamer (magenta) and an adjacent gp23* hexamer. The gp23* subunits are shown in blue, cyan, green, tan, indigo, and gray. Surfaces of Hoc C-terminal domains corresponding to the two preferred orientations are shown in orange and red. (<b>D</b>) Backbone traces of the Hoc C-terminal domain in the two preferred orientations (yellow and red) fitted into the cryo-EM density of the isometric T4 capsid reconstruction (EMDB-8661) (blue mesh). The letters A–F in panels (<b>B</b>,<b>C</b>) represent the major capsid protein (gp23*) subunits of the hexameric capsomer.</p>
Full article ">Figure 4
<p>Structure of the T4 Hoc C-terminal domain. (<b>A</b>) Bottom view of the Hoc C-terminal domain (residues 281–376) corresponding to the preferred orientation 1. The polypeptide chain color changes from yellow at the residue 281 to red at the C-terminal residue 376. The sidechains of the residues interacting with the gp23* subunits are shown as sticks. (<b>B</b>) Superposition of the Hoc C-terminal domain in orientation 1 (orange) with the C-terminal domain in orientation 2 (red) and the C-terminal domain derived from the AlphaFold model (blue).</p>
Full article ">Figure 5
<p>Structure of the T4 Hoc C-terminal domain attached to the center of a vertex–proximal gp23* hexamer. Different gp23* subunits are depicted in different colors. Panels (<b>A</b>–<b>C</b>) show Hoc C-terminal domain in the preferred orientation 1 (yellow), and panels (<b>D</b>–<b>F</b>) show the same in the preferred orientation 2 (red). The letters A–F in panels A and D represent the major capsid protein (gp23*) subunits of the hexameric capsomer.</p>
Full article ">Figure 6
<p>Interactions of Hoc C-terminal domain with the major capsid protein gp23* subunits. (<b>A</b>) Regions Gly<sup>326</sup>-Arg<sup>344</sup> of the six gp23* subunits forming the Hoc binding sites are shown in different colors. The region Val<sup>353</sup>-Tyr<sup>360</sup> of the Hoc protein in the preferred orientation 1 interacting with gp23* is shown in red. This Hoc region includes the following conserved loop: Glu<sup>355</sup>-Ser<sup>356</sup>-Arg<sup>357</sup>-Asn<sup>358</sup>-Gly<sup>359</sup>. The letters A-F in panel A represent the major capsid protein (gp23*) subunits of the hexameric capsomer. (<b>B</b>) The Hoc C-terminal domain in the preferred orientation 1 interacting with the gp23* surface. The polypeptide chain color changes from yellow at the residue 281 to red at the C-terminal residue 376. The side chains of residues involved in Hoc-gp23* interactions are shown as sticks. (<b>C</b>,<b>D</b>) Close views of some of the Hoc-gp23* interactions.</p>
Full article ">Figure 7
<p>Alignment of Hoc C-terminal domain sequences. Only the nine most homologous sequences found by Blast and the Hoc C-terminal domains of phages RB69, JS98, Muldoon, RB49, RB43, and 44RR2.8t are shown here. For the complete sequence alignment (including 650 sequences), see <a href="#app1-viruses-15-01517" class="html-app">Supplementary File S1</a>. The residue colors are based on the conservation. The color ranges from gray for the most conserved residues to red for the least conserved residues. The highly conserved Glu<sup>355</sup>-Ser<sup>356</sup>-Arg<sup>357</sup>-Asn<sup>358</sup>-Gly<sup>359</sup> region is delineated by the black rectangle.</p>
Full article ">Figure 8
<p>Clustering behavior of T4 phage with and without Hoc. Plaque assays were performed after incubation of Hoc(−) Soc(−) phage with or without the recombinant T4-Hoc or RB49-Hoc at 4 °C. Percentages and fold differences of plaques were calculated by taking the plaque titer of control Hoc(−) Soc(−) phage as 100% (<b>A</b>), or 1 (<b>B</b>). Error bars were determined from quadruplicate assays. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001, one-way ANOVA test.</p>
Full article ">Figure 9
<p>The dispersion of T4 heads (capsids) is dependent on Hoc. (<b>A</b>) Purified T4 capsids were incubated in pH 5.6 (lane 1 to 3) or pH 8 (lane 4 to 6) buffers either with or without the recombinant T4 or RB49 Hoc and in the presence or absence of salts. The samples were sedimented at 8000× <span class="html-italic">g</span> for 45 min, and SDS–PAGE analysis of the pellets was performed. (<b>B</b>) The volume of gp23* band from each lane was determined by ImageDoc, and bar graphs were plotted using the pixel values. The volume of gp23* of Hoc(−) head control was considered as 100%. Error bars were determined from duplicate assays. * <span class="html-italic">p</span> &lt; 0.05, one-way ANOVA test.</p>
Full article ">
15 pages, 3164 KiB  
Article
P3 and NIa-Pro of Turnip Mosaic Virus Are Independent Elicitors of Superinfection Exclusion
by Haritha Nunna, Feng Qu and Satyanarayana Tatineni
Viruses 2023, 15(7), 1459; https://doi.org/10.3390/v15071459 - 28 Jun 2023
Cited by 2 | Viewed by 1676
Abstract
Superinfection exclusion (SIE) is an antagonistic interaction between identical or closely related viruses in host cells. Previous studies by us and others led to the hypothesis that SIE was elicited by one or more proteins encoded in the genomes of primary viruses. Here, [...] Read more.
Superinfection exclusion (SIE) is an antagonistic interaction between identical or closely related viruses in host cells. Previous studies by us and others led to the hypothesis that SIE was elicited by one or more proteins encoded in the genomes of primary viruses. Here, we tested this hypothesis using Turnip mosaic virus (TuMV), a member of the genus Potyvirus of the family Potyviridae, with significant economic consequences. To this end, individual TuMV-encoded proteins were transiently expressed in the cells of Nicotiana benthamiana leaves, followed by challenging them with a modified TuMV expressing the green fluorescent protein (TuMV-GFP). Three days after TuMV-GFP delivery, these cells were examined for the replication-dependent expression of GFP. Cells expressing TuMV P1, HC-Pro, 6K1, CI, 6K2, NIa-VPg, NIb, or CP proteins permitted an efficient expression of GFP, suggesting that these proteins failed to block the replication of a superinfecting TuMV-GFP. By contrast, N. benthamiana cells expressing TuMV P3 or NIa-Pro did not express visible GFP fluorescence, suggesting that both of them could elicit potent SIE against TuMV-GFP. The SIE elicitor activity of P3 and NIa-Pro was further confirmed by their heterologous expression from a different potyvirus, potato virus A (PVA). Plants systemically infected with PVA variants expressing TuMV P3 or NIa-Pro blocked subsequent infection by TuMV-GFP. A +1-frameshift mutation in P3 and NIa-Pro cistrons facilitated superinfection by TuMV-GFP, suggesting that the P3 and NIa-Pro proteins, but not the RNA, are involved in SIE activity. Additionally, deletion mutagenesis identified P3 amino acids 3 to 200 of 352 and NIa-Pro amino acids 3 to 40 and 181 to 242 of 242 as essential for SIE elicitation. Collectively, our study demonstrates that TuMV encodes two spatially separated proteins that act independently to exert SIE on superinfecting TuMV. These results lay the foundation for further mechanistic interrogations of SIE in this virus. Full article
(This article belongs to the Special Issue Crop Resistance to Viral Infections)
Show Figures

Figure 1

Figure 1
<p>Demonstration of superinfection exclusion of Turnip mosaic virus (TuMV). (<b>A</b>) Schematic representation of the genomes of two variants of TuMV with GFP (TuMV-GFP) or RFP (TuMV-RFP) genes inserted between P1 and HC-Pro cistrons. (<b>B</b>) <span class="html-italic">Nicotiana benthamiana</span> leaves were co-agroinfiltrated with TuMV-GFP and TuMV-RFP, followed by an examination of the expression of GFP and RFP under a confocal microscope. The ‘merged’ image is a superimposed image showing the expression of GFP and RFP, depicting the mutual exclusion of two variants of TuMV. Bars:100 µm.</p>
Full article ">Figure 2
<p>Screening of the Turnip mosaic virus (TuMV) genome for elicitors of superinfection exclusion. TuMV-encoded cistrons were engineered into pCASS4 between 35S promoter and terminator sequences, followed by agroinfiltration into <span class="html-italic">Nicotiana benthamiana</span> leaves. The agroinfiltrated leaves were challenge agroinfiltrated with pCB-TuMV-GFP at 24 h post-agroinfiltration. (<b>A</b>) Western blot of total proteins extracted from agroinfiltrated HA-tagged TuMV cistrons with anti-HA antiserum. Arrowheads indicate expected TuMV proteins. The Coomassie-stained SDS-PAGE shows the RUBISCO protein for the amount of protein loaded per well. (<b>B</b>) Expression of GFP in pre-agroinfiltrated <span class="html-italic">N. benthamiana</span> leaves with TuMV cistrons that were challenge-agroinfiltrated with pCB-TuMV-GFP at 3 days post-agroinfiltration. Pre-agroinfiltrated TuMV cistrons were indicated on top of each picture. All TuMV cistrons except P3 and NIa-Pro facilitated the expression of GFP. Bars: 100 µm. (<b>C</b>) Western blot analysis of total proteins from pre-agroinfiltrated <span class="html-italic">N. benthamiana</span> leaves that were challenge-agroinfiltrated with pCB-TuMV-GFP. The blots were treated with anti-GFP monoclonal antiserum. The Coomassie-stained gel expressing the RUBISCO protein for the amount of protein loaded per well.</p>
Full article ">Figure 3
<p>Turnip mosaic virus P3- and NIa-Pro-encoded proteins, but not the RNA sequences, are required for SIE activity. (<b>A</b>) Schematic representation of +1 frameshift mutations (red-colored amino acids represent the change in amino acid due to a +1 frameshift) into HA-tagged TuMV P3 and NIa-Pro cistrons. The +1 frameshift mutations resulted in a stop codon (asterisk) after 14 and 9 amino acid codons in P3 and NIa-Pro cistrons, respectively. (<b>B</b>) Expression of GFP in challenge agroinfiltrated with TuMV-GFP in <span class="html-italic">Nicotiana benthamiana</span> leaves that were pre-agroinfiltrated with agrosuspension harboring TuMV-P3 + 1FS, TuMV-P3, TuMV-NIa-Pro + 1FS, or TuMV-NIa-Pro at 3 days post-challenge agroinfiltration. Bars: 100 µm.</p>
Full article ">Figure 4
<p>Potato virus A and Turnip mosaic virus can co-infect the same cell. (<b>A</b>) Schematic diagrams showing the genomic organizations of TuMV-GFP and PVA-RFP. (<b>B</b>) Expression of GFP and RFP in <span class="html-italic">Nicotiana benthamiana</span> leaves co-agroinfiltrated with pRD400-PVA-RFP and pCB-TuMV-GFP. The ‘merged’ image on the right is a superimposed image expressing both GFP and RFP. The yellow-colored cells indicate the co-infection of both viruses. (<b>C</b>) Expression of GFP and RFP in <span class="html-italic">N. benthamiana</span> leaves that were first agroinfiltrated with pCB-TuMV-GFP, followed by super agroinfiltrated with pRD400-PVA-RFP. The upper non-agroinfiltrated <span class="html-italic">N. benthamiana</span> leaves were observed in a confocal microscope for the expression of RFP and GFP resulting from the replication of PVA and TuMV, respectively. The merged image on the right shows the co-infection of TuMV-GFP and PVA-RFP. Bars in B and C are 20 µm and 100 µm, respectively.</p>
Full article ">Figure 5
<p>Superinfection exclusion assay of select Turnip mosaic virus cistrons using the add-a-gene strategy in potato virus A (PVA). (<b>A</b>) Schematic diagram of PVA-RFP genome containing TuMV P3, NIa-Pro, or CP with 26 aa of CP and 18 aa of NIb at the 5′ and 3′ end of TuMV cistrons, respectively, for efficient cleavage of proteins [<a href="#B44-viruses-15-01459" class="html-bibr">44</a>]. (<b>B</b>) <span class="html-italic">Nicotiana benthamiana</span> leaves showing the expression of RFP from PVA-RFP with TuMV P3, NIa-Pro, or CP and GFP from TuMV-GFP. <span class="html-italic">N. benthamiana</span> leaves expressing PVA-RFP-TuMV-P3 or NIa-Pro abolished co-infection by TuMV-GFP, while PVA-RFP-TuMV CP efficiently permitted the co-infection by TuMV-GFP. Bars: 100 µm. (<b>C</b>) Stability assay of TuMV cistrons in the PVA genome. Agarose gel electrophoresis of RT-PCR products of TuMV cistrons from <span class="html-italic">N. benthamiana</span> leaves infiltrated with PVA-RFP harboring TuMV cistrons at 10 days post-agroinfiltration; lane 1. TuMV P3; lane 2. TuMV NIa-Pro; lane 3. TuMV CP; lane 4. PVA-RFP; lane 5: mock, and lane M: 1 kbp Plus DNA ladder (Invitrogen).</p>
Full article ">Figure 6
<p>Turnip mosaic virus P3 amino acids 3 to 220 are required for SIE activity. (<b>A</b>) Diagrammatic representation of series of non-overlapping deletions in TuMV P3 cistron. Dark rectangle boxes indicate the location of amino acids deleted in the P3 cistron. (<b>B</b>) Superinfection exclusion assay of TuMV P3 deletion mutants. GFP expression in <span class="html-italic">Nicotiana benthamiana</span> leaves challenge-agroinfiltrated with pCB-TuMV-GFP into leaves pre-agroinfiltrated with pCASS4-TuMV P3 deletion mutants. Deletion of P3 amino acids 3 to 200 efficiently permitted infection by TuMV-GFP, while deletion of amino acids 201 to 240 showed a weak expression of GFP, depicting that some of the deleted amino acids in this region are required for SIE. Deletions comprising P3 amino acids 241 to 352 completely abolished superinfection by TuMV GFP, indicating that this region is dispensable for SIE activity. Bars: 100 µm.</p>
Full article ">Figure 7
<p>Mapping Turnip mosaic virus NIa-Pro amino acids required for SIE. (<b>A</b>) Schematic representation of a series of deletions in TuMV NIa-Pro. Dark rectangles represent deleted amino acids in the NIa-Pro cistron. (<b>B</b>) SIE assay of TuMV NIa-Pro deletion mutants. GFP expression in <span class="html-italic">Nicotiana benthamiana</span> leaves challenge-agroinfiltrated with pCB-TuMV-GFP into leaves pre-agroinfiltrated with NIa-Pro deletion mutants. TuMV NIa-Pro deletions comprising amino acids 3 to 40 and 81 to 242 did not elicit SIE, indicating that these amino acids are required for SIE. Bars: 100 µm.</p>
Full article ">
13 pages, 3462 KiB  
Article
Proteomics Identified UDP-Glycosyltransferase Family Members as Pro-Viral Factors for Turnip Mosaic Virus Infection in Nicotiana benthamiana
by Kaida Ding, Zhaoxing Jia, Penghuan Rui, Xinxin Fang, Hongying Zheng, Jianping Chen, Fei Yan and Guanwei Wu
Viruses 2023, 15(6), 1401; https://doi.org/10.3390/v15061401 - 20 Jun 2023
Cited by 1 | Viewed by 1768
Abstract
Viruses encounter numerous host factors that facilitate or suppress viral infection. Although some host factors manipulated by viruses were uncovered, we have limited knowledge of the pathways hijacked to promote viral replication and activate host defense responses. Turnip mosaic virus (TuMV) is one [...] Read more.
Viruses encounter numerous host factors that facilitate or suppress viral infection. Although some host factors manipulated by viruses were uncovered, we have limited knowledge of the pathways hijacked to promote viral replication and activate host defense responses. Turnip mosaic virus (TuMV) is one of the most prevalent viral pathogens in many regions of the world. Here, we employed an isobaric tag for relative and absolute quantitation (iTRAQ)-based proteomics approach to characterize cellular protein changes in the early stages of infection of Nicotiana benthamiana by wild type and replication-defective TuMV. A total of 225 differentially accumulated proteins (DAPs) were identified (182 increased and 43 decreased). Bioinformatics analysis showed that a few biological pathways were associated with TuMV infection. Four upregulated DAPs belonging to uridine diphosphate-glycosyltransferase (UGT) family members were validated by their mRNA expression profiles and their effects on TuMV infection. NbUGT91C1 or NbUGT74F1 knockdown impaired TuMV replication and increased reactive oxygen species production, whereas overexpression of either promoted TuMV replication. Overall, this comparative proteomics analysis delineates the cellular protein changes during early TuMV infection and provides new insights into the role of UGTs in the context of plant viral infection. Full article
(This article belongs to the Section Viruses of Plants, Fungi and Protozoa)
Show Figures

Figure 1

Figure 1
<p>Infectivity test of TuMV-GFP and replication-deficient TuMV-GFP//△GDD in <span class="html-italic">N. benthamiana</span>. (<b>A</b>) Representative photographs of TuMV-GFP and TuMV-GFP//△GDD infection in <span class="html-italic">N. benthamiana</span> plants. Pictures were taken under regular light and UV light at 4 days post infection. (<b>B</b>) Western blot detection of TuMV coat protein accumulation in the inoculated leaves. (<b>C</b>) RT-qPCR results showing viral positive-sense (+) and negative-sense (−) RNA levels in TuMV-GFP- and TuMV-GFP//△GDD-infected local leaves. The error bar represents the standard deviation of three biological replicates of a representative experiment. Statistical analysis was performed using Student’s <span class="html-italic">t</span>-test (**, <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 2
<p>Number and KOG functional classification of differentially abundant host proteins between TuMV-GFP- and TuMV-GFP//△GDD-infected <span class="html-italic">N. benthamiana</span> leaves. (<b>A</b>) Volcano plot illustrating the significantly differentially accumulated proteins. The –log10 (Benjamini-Hochberg corrected <span class="html-italic">p</span> value) is plotted against the log2 (fold change: TuMV/TuMV-△GDD). The non-axial vertical lines denote ±1.3-fold change while the non-axial horizontal line denotes <span class="html-italic">p</span> = 0.05, which is our significance threshold (prior to logarithmic transformation). (<b>B</b>) Histogram displaying the number of differentially abundant proteins within a specific range of fold changes. (<b>C</b>) GO enrichment classification of differential accumulated proteins. Red colour bars indicate upregulated proteins and blue colour bars indicate downregulated proteins.</p>
Full article ">Figure 3
<p>Validation of the selected NbUGTs RNA expression in the leaves of TuMV-GFP- and TuMV-GFP//△GDD-infected <span class="html-italic">N. benthamiana</span> plants by real time RT-PCR. ** and *** represent statistically significant differences by the Student <span class="html-italic">t</span>-test between groups at <span class="html-italic">p</span> &lt; 0.01 and <span class="html-italic">p</span> &lt; 0.001, respectively. NS, not significant.</p>
Full article ">Figure 4
<p>Effect of silencing the selected NbUGTS on TuMV infection in <span class="html-italic">N. benthamiana</span> plants. (<b>A</b>) Phenotypes of <span class="html-italic">N. benthamiana</span> with silenced NbUGTS before and after TuMV infection (top row). The empty vector TRV::00 was used as control. Lower panels show GFP fluorescence resulting from TuMV-GFP infection under a hand-held UV lamp. (<b>B</b>) TuMV and TRV CP accumulation levels in the TRV+TuMV-inoculated plants. Both inoculated and upper non-inoculated tissues were harvested for Western blotting assay. (<b>C</b>) The oxidative burst in the upper NbUGTs-silenced and control leaves observed by staining with diaminobenzidine (DAB) at 6 dpai after TuMV infection. Bars, 5 cm.</p>
Full article ">Figure 5
<p>Effect of NbUGT74F1 and NbUGT91C1 overexpression on TuMV infection. (<b>A</b>) GFP fluorescence in <span class="html-italic">N. benthamiana</span> plants inoculated with TuMV-GFP together with GUS (control), NbUGT74F1, or NbUGT91C1. Plants were photographed under a hand-held UV lamp at 3 dpai. (<b>B</b>) RT-qPCR results showing the quantification of positive-strand viral genomic RNA [(+)RNA] or negative-strand viral genomic RNA [(−) RNA] accumulation in <span class="html-italic">N. benthamiana</span> plants agroinfiltrated with different combinations of plasmids from (<b>A</b>). The infiltrated leaf tissues were collected for RNA purification at 65 h post agroinfiltration (hpai) and RT-qPCR was performed with TuMV nib-specific primers using the actin II transcript level as an internal control. The error bar represents the standard deviation of three biological replicates of a representative experiment. Statistical analysis was performed using Student’s <span class="html-italic">t</span>-test **, <span class="html-italic">p</span> &lt; 0.01). (<b>C</b>) Western blotting analysis of the TuMV coat protein (CP) in the infiltrated leaf tissues from <span class="html-italic">N. benthamiana</span> plants in (<b>A</b>) at 65 hpai. Coomassie Brilliant Blue R-250-stained RuBisco large subunit serves as a loading control. TuMV CP was detected with anti-TuMV CP polyclonal antibody. NbUGT74F1 and NbUGT91C1 were detected with anti-c-Myc monoclonal antibodies.</p>
Full article ">
20 pages, 8185 KiB  
Article
Plant Poly(ADP-Ribose) Polymerase 1 Is a Potential Mediator of Cross-Talk between the Cajal Body Protein Coilin and Salicylic Acid-Mediated Antiviral Defence
by Nadezhda Spechenkova, Viktoriya O. Samarskaya, Natalya O. Kalinina, Sergey K. Zavriev, S. MacFarlane, Andrew J. Love and Michael Taliansky
Viruses 2023, 15(6), 1282; https://doi.org/10.3390/v15061282 - 30 May 2023
Cited by 3 | Viewed by 2072
Abstract
The nucleolus and Cajal bodies (CBs) are sub-nuclear domains with well-known roles in RNA metabolism and RNA-protein assembly. However, they also participate in other important aspects of cell functioning. This study uncovers a previously unrecognised mechanism by which these bodies and their components [...] Read more.
The nucleolus and Cajal bodies (CBs) are sub-nuclear domains with well-known roles in RNA metabolism and RNA-protein assembly. However, they also participate in other important aspects of cell functioning. This study uncovers a previously unrecognised mechanism by which these bodies and their components regulate host defences against pathogen attack. We show that the CB protein coilin interacts with poly(ADP-ribose) polymerase 1 (PARP1), redistributes it to the nucleolus and modifies its function, and that these events are accompanied by substantial increases in endogenous concentrations of salicylic acid (SA), activation of SA-responsive gene expression and callose deposition leading to the restriction of tobacco rattle virus (TRV) systemic infection. Consistent with this, we also find that treatment with SA subverts the negative effect of the pharmacological PARP inhibitor 3-aminobenzamide (3AB) on plant recovery from TRV infection. Our results suggest that PARP1 could act as a key molecular actuator in the regulatory network which integrates coilin activities as a stress sensor for virus infection and SA-mediated antivirus defence. Full article
(This article belongs to the Special Issue Plant Viruses: Pirates of Cellular Pathways)
Show Figures

Figure 1

Figure 1
<p>Effect of TRV infection on <span class="html-italic">N. benthamiana PARP1</span> (<span class="html-italic">NbPARP1</span>) gene expression at 14 dpi. Accumulation of PARP1mRNA (<b>a</b>) and TRV RNA1 used as an indicator of TRV multiplication [<a href="#B33-viruses-15-01282" class="html-bibr">33</a>] (<b>b</b>) (measured by RT-qPCR) in newly emerged leaves of WT and coilin KD <span class="html-italic">N. benthamiana</span> plants infected with or without (mock-inoculated) TRV or TRV∆16K. PARP1 mRNA and TRV RNA1 expression levels were normalized to those of internal <span class="html-italic">N. benthamiana</span> controls, <span class="html-italic">UBIQUITIN3</span> gene (<span class="html-italic">UBI3)</span> and <span class="html-italic">60S ribosomal protein 23</span> gene (<span class="html-italic">L23</span>) [<a href="#B46-viruses-15-01282" class="html-bibr">46</a>]. Statistical analysis was performed on four independent biological replicates. Data are mean ± SD. Each replicate was composed of samples from three plants pooled together (two leaves per plant). Analysis of variance and Tukey’s HSD post hoc tests were performed on the RT-qPCR data. NS, not significant; ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 2
<p>Interaction of PARP1 with coilin. (<b>a</b>) Co-immunoprecipitation of PARP1, coilin and TRV 16K protein. Protein extracts were prepared at 7 dpi from older systemically infected leaves (third and fourth leaves above the inoculated leaf) of WT or coilin KD plants infected with or without (mock-inoculated, M) TRV or TRV∆16K, as indicated. Proteins in the lysate prior to immunoprecipitation (IP) are shown on the left (input). Anti-coilin antibodies (anti-coilin) were used to co-immunoprecipitate coilin and PARP. Anti-PARP antibodies were used to co-immunoprecipitate PARP and coilin. Antibodies to 16 K (anti-16K) were used to co-precipitate 16K, coilin and PARP. Proteins were detected by western blot analysis (immunoblotting, IB) using anti-PARP and anti-coilin antibodies. Positions of molecular mass markers are on the left. Images have been cropped for presentation. Uncropped images are presented in <a href="#app1-viruses-15-01282" class="html-app">Supplementary Materials</a>. (<b>b</b>) Far-western blot analysis of the in vitro interaction between coilin and commercially-sourced PARP1. Bovine serum albumin (BSA) was used as a negative control. The left blot was stained with Ponceau red; the middle and right blots were incubated with and without the recombinant coilin, respectively (as indicated), and then probed with anti-coilin antiserum. Positions of the molecular mass markers are indicated on the left.</p>
Full article ">Figure 3
<p>The interaction between coilin and PARP1 induced by TRV results in partial nucleolar sequestration of PARP1 and over-accumulation of PARylated proteins. (<b>a</b>) Representative images of intranuclear distribution of PARP1 (immunofluorescent staining using primary rabbit anti-PARP1 antibody and secondary fluorescent anti-rabbit antibody, green ) in WT and coilin KD <span class="html-italic">N. benthamiana</span> plants infected with or without (mock-inoculated) TRV or TRV∆16K were taken at 7 dpi in older systemically infected leaves (third and fourth leaves above the inoculated leaf) or at 14 dpi in recovered newly emerging leaves of WT plants systemically infected with TRV [seventh and eighth leaves above the inoculated leaf; WT-TRV(R)] or at 3 days post-agroinfiltration (dpa) in leaves agroinfiltrated with a construct expressing the 16K protein (WT-16K or KD-16K). N, nuclei; No, nucleoli; CBs are shown by arrows. Scale bars, 5 µm. (<b>b</b>) Quantification of results presented in (<b>a</b>). The ratio of nucleolar fluorescence to nucleoplasmic fluorescence (Fno/Fnu) was averaged for at least 100 cells in three independent experiments. Data are mean ± SD. Analysis of variance and Tukey’s HSD post hoc tests were performed on the data obtained. ** <span class="html-italic">p</span> &lt; 0.01. (<b>c</b>) Accumulation of PARylated proteins measured by ELISA using rabbit anti-PAR polyclonal antibody, in plants described in (<b>a</b>). Data are mean ± SD, n = 6 from three independent experiments. Analysis of variance and Tukey’s HSD post hoc tests were performed on the data obtained. ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 4
<p>Effect of 3-aminobenzamide (3AB) on the development of TRV infection and accumulation of PARylated proteins in WT <span class="html-italic">N. benthamiana</span> plants. (<b>a</b>) Symptoms induced in TRV-infected and mock-inoculated plants treated with or without 3AB. (<b>b</b>) Northern blot analysis of TRV RNA1 in inoculated (inoc, 7 dpi) and newly emerging systemically infected leaves (seventh and eighth leaves above the inoculated leaf; sys-N, 14 dpi). The positions of RNA size markers are indicated on the left. Ethidium bromide (EtBr)-stained rRNA (bottom panel) is shown as a loading control. Images have been cropped for presentation. Uncropped images are presented in <a href="#app1-viruses-15-01282" class="html-app">Supplementary Materials</a>. (<b>c</b>) Accumulation of TRV RNA1 (measured using RT-qPCR) in inoculated (inoc; 2 and 7 dpi), older systemically infected leaves (third and fourth leaves above the inoculated leaf; sys-O, 7 and 14 dpi) and newly emerging systemically infected leaves (seventh and eighth leaves above the inoculated leaf; sys-N, 14 and 21 dpi). TRV RNA1 expression levels were normalized to those of the internal controls, <span class="html-italic">UBIQUITIN3</span> gene <span class="html-italic">(UBI3)</span> and <span class="html-italic">60S ribosomal protein 23</span> gene <span class="html-italic">(L23)</span>. (<b>d</b>) Accumulation of PARylated proteins was measured by ELISA using rabbit anti-PAR polyclonal antibody, in TRV- systemically infected or mock-inoculated plants treated with or without 3AB (older third and fourth leaves above the inoculated leaf). Data are mean ± SD, <span class="html-italic">n</span> = 6 from three independent experiments. Analysis of variance and Tukey’s HSD post hoc tests were performed on the data obtained. ** <span class="html-italic">p</span> &lt; 0.01; NS, non-significant (<b>c</b>,<b>d</b>). These data could suggest that a change in PARP1 shuttling activity leading to over-accumulation of PAR (associated with PARP target proteins) in TRV-infected leaves subsequently activates host defence and results in plant recovery. However, given that pharmacological PARP inhibitors including 3AB may not only affect the activity of canonical PARPs, but also have off-target effects [<a href="#B1-viruses-15-01282" class="html-bibr">1</a>,<a href="#B2-viruses-15-01282" class="html-bibr">2</a>,<a href="#B48-viruses-15-01282" class="html-bibr">48</a>], results of pharmacological experiments to infer PARP function in plants should be verified by the genetic inhibition of PARP activity.</p>
Full article ">Figure 5
<p>Effect of virus-induced silencing of <span class="html-italic">PARP1</span> (<span class="html-italic">NbPARP1</span>) expression on TRV infection. Two separate PVX-PARP1 VIGS constructs made in this work (see Materials and Methods) exhibited similar effects on <span class="html-italic">NbPARP1</span> gene expression (<b>a</b>), accumulation of PARylated proteins (<b>b</b>) and TRV infection (<b>c</b>,<b>d</b>), which are exemplified in this figure by the data obtained in the experiments using fragment 1. (<b>a</b>) Virus-induced silencing of the <span class="html-italic">PARP1</span> gene in <span class="html-italic">N. benthamiana</span> mediated by a PVX vector which contains fragment 1 of the <span class="html-italic">NbPARP1</span> gene (PVX-PARP), compared with an empty PVX vector control (PVX-C). Accumulation of PARP1 mRNA was measured using RT-qPCR in inoculated (inoc) and newly emerging systemically infected (sys-N) leaves at 10 dpi. Results from three independent experiments (I, II, III) are shown. (<b>b</b>) Effect of PVX-induced PARP1 silencing on the accumulation of PARylated proteins measured by ELISA using a rabbit anti-PAR polyclonal antibody, in the plant leaves shown in (<b>a</b>). Results from three independent experiments (I, II, III) are shown. (<b>c</b>) Accumulation of TRV RNA1 (measured using RT-qPCR) in inoculated (inoc) and newly emerging systemically infected (sys-N) leaves of PARP-silenced plants at 10 dpi; the same leaves as in (<b>a</b>) and (<b>b</b>) were analysed. Results from three independent experiments (I, II, III) are shown (<b>a</b>–<b>c</b>). Data are mean ± SD, <span class="html-italic">n</span> = 4 from three independent biological replicates. Analysis of variance and Tukey’s HSD post hoc tests were performed on the data obtained. ** <span class="html-italic">p</span> &lt; 0.01; NS, non-significant (<b>c</b>,<b>d</b>). (<b>d</b>) Symptoms induced in plants infected with TRV in PARP1 silenced (PVX-PARP) and non-silenced [PVX-C and (-)PVX] plants.</p>
Full article ">Figure 6
<p>Concentrations of free (<b>a</b>) and conjugated (SA-b-glucoside) salicylic acid (SA) (<b>b</b>) in inoculated (inoc; 7 dpi), older systemically infected leaves (sys-O; 14 dpi) and newly emerging sys-N; 14 dpi) leaves of <span class="html-italic">N. benthamiana</span> plants infected or uninfected (mock-inoculated) with TRV and treated with or without 3AB. Statistical analysis was performed on four independent biological replicates. Data are mean ± SD. Each replicate was composed of samples from three plants pooled together (two leaves per plant). Analysis of variance and Tukey’s HSD post hoc tests were performed on the data obtained. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01. NS, non-significant.</p>
Full article ">Figure 7
<p>Effect of foliar treatment with salicylic acid (SA) on the transcript level of the PR-1a protein gene (measured using RT-qPCR) (<b>a</b>), callose deposition (<b>b</b>) and accumulation of TRV (<b>c</b>) in inoculated (inoc; 7 dpi), older systemically infected leaves (sys-O; 14 dpi) and newly emerging sys-N; 14 dpi) leaves of <span class="html-italic">N. benthamiana</span> plants infected or uninfected (mock-inoculated) with TRV and treated with or without 3AB. Statistical analysis was performed on four independent biological replicates. Data are mean ± SD. Each replicate was composed of samples from three plants pooled together (two leaves per plant). Analysis of variance and Tukey’s HSD post hoc tests were performed on the data obtained. *<span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01. (<b>d</b>) Symptoms induced in plants infected with TRV after treatment with 3AB and SA as indicated.</p>
Full article ">Figure 8
<p>A model of the TRV infection process and the involvements of the 16K protein, coilin and PARP1. In healthy plants (<b>a</b>), coilin is located within CBs and the nucleoplasm but is not present in the nucleolus. PARP1, a nuclear protein, modifies the function and subcellular localisation of a variety of nuclear “target” proteins (acceptors) by attaching chains of ADP ribose (PAR) to them. To re-activate these target proteins, PARP1 shuttles them from both the nucleolus (NO) and chromatin (chromatin not shown) to CBs for PAR removal and recycling. Upon TRV infection (<b>b</b>), the viral 16K protein is produced in the cytoplasm and is targeted to the nucleus. In the nucleus (CBs and nucleoplasm), the 16K protein interacts with coilin and relocalises it to the nucleolus, which in turn traps PARP1 within this sub-nuclear domain, preventing it trafficking to CBs for PAR cleavage and recycling. This leads to over-accumulation of PAR/PARylated proteins and may enhance accumulation of salicylic acid (SA) and increased elicitation of SA-mediated defence responses (represented here by increased expression of the <span class="html-italic">PR-1a</span> gene and by callose deposition). These responses restrict TRV spread into newly emerging leaves, leading to the plant’s recovery from TRV infection. Thus, PARP1 can act as a mediator in a functional link between stress-sensing activities of coilin and SA-mediated antivirus defence.</p>
Full article ">
15 pages, 6010 KiB  
Article
Noncoding RNA of Zika Virus Affects Interplay between Wnt-Signaling and Pro-Apoptotic Pathways in the Developing Brain Tissue
by Andrii Slonchak, Harman Chaggar, Julio Aguado, Ernst Wolvetang and Alexander A. Khromykh
Viruses 2023, 15(5), 1062; https://doi.org/10.3390/v15051062 - 26 Apr 2023
Cited by 3 | Viewed by 2107
Abstract
Zika virus (ZIKV) has a unique ability among flaviviruses to cross the placental barrier and infect the fetal brain causing severe abnormalities of neurodevelopment known collectively as congenital Zika syndrome. In our recent study, we demonstrated that the viral noncoding RNA (subgenomic flaviviral [...] Read more.
Zika virus (ZIKV) has a unique ability among flaviviruses to cross the placental barrier and infect the fetal brain causing severe abnormalities of neurodevelopment known collectively as congenital Zika syndrome. In our recent study, we demonstrated that the viral noncoding RNA (subgenomic flaviviral RNA, sfRNA) of the Zika virus induces apoptosis of neural progenitors and is required for ZIKV pathogenesis in the developing brain. Herein, we expanded on our initial findings and identified biological processes and signaling pathways affected by the production of ZIKV sfRNA in the developing brain tissue. We employed 3D brain organoids generated from induced human pluripotent stem cells (ihPSC) as an ex vivo model of viral infection in the developing brain and utilized wild type (WT) ZIKV (producing sfRNA) and mutant ZIKV (deficient in the production of sfRNA). Global transcriptome profiling by RNA-Seq revealed that the production of sfRNA affects the expression of >1000 genes. We uncovered that in addition to the activation of pro-apoptotic pathways, organoids infected with sfRNA-producing WT, but not sfRNA-deficient mutant ZIKV, which exhibited a strong down-regulation of genes involved in signaling pathways that control neuron differentiation and brain development, indicating the requirement of sfRNA for the suppression of neurodevelopment associated with the ZIKV infection. Using gene set enrichment analysis and gene network reconstruction, we demonstrated that the effect of sfRNA on pathways that control brain development occurs via crosstalk between Wnt-signaling and proapoptotic pathways. Full article
(This article belongs to the Special Issue Molecular Biology of RNA Viruses)
Show Figures

Figure 1

Figure 1
<p>Differential host gene expression in iPSC-derived human brain organoids infected with WT and sfRNA-deficient ZIKV. Human brain organoids at DIV15 were infected with WT or sfRNA-deficient ZIKV mutant at a dose of 10<sup>5</sup> FFU. Total RNA was isolated at 3 dpi and analyzed by RNA-Seq. (<b>A</b>) Volcano plot showing differentially expressed genes in human brain organoids infected with WT ZIKV. (<b>B</b>) Volcano plot showing differentially expressed genes in human brain organoids infected with sfRNA-deficient ZIKV mutant. In (<b>A</b>,<b>B</b>) genes with significantly different expression levels compared to the mock (FDR-corrected <span class="html-italic">p</span>-value &lt; 0.05 and logFC &gt; 1 or &lt;−1) are shown in the color red for upregulated genes and blue for downregulated genes. The top most significant differentially expressed genes (DEGs) are labeled. (<b>C</b>) Comparison of gene expression levels between human brain organoids infected with WT and sfRNA-deficient (Mut) ZIKV. Genes with significantly different expressions (FDR-corrected <span class="html-italic">p</span>-value &lt; 0.05 and logFC &gt; 1 or &lt;−1) are shown in the color red for genes with higher expression in WT infection compared to Mut virus infection and blue for genes with lower expression in WT infection compared to Mut virus infection. (<b>D</b>) Heat map representation of the expression levels of the top 100 most significant genes identified in (<b>C</b>).</p>
Full article ">Figure 2
<p>qRT-PCR validation of differentially expressed host genes identified by global transcriptome profiling. RNA isolated from human brain organoids infected with WT or sfRNA-deficient mutant viruses was subjected to cDNA synthesis and qRT-PCR. RNA from mock-infected organoids was used as a control. (<b>A</b>) Expression of the genes involved in brain development and ISGs determined using the ΔΔC<sub>T</sub> method relative to the mock, with normalization to TBP mRNA level. (<b>B</b>) Viral RNA levels in infected organoids determined as copy numbers per ug of input RNA using the standard curve method. (<b>C</b>) Expression of host genes involved in brain development and ISGs normalized to viral RNA levels. The values are the means of three replicates ± SD. Statistical analysis was performed using Student’s <span class="html-italic">t</span>-test: **** <span class="html-italic">p</span> &lt; 0.0001, *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 3
<p>Functional classification of differentially expressed host genes that responded differently to infection with WT compared to sfRNA-deficient ZIKV in iPSC-derived human brain organoids. Size of bubbles on the plots is proportional to the number of associated genes; significantly enriched categories are labeled.</p>
Full article ">Figure 4
<p>Expression of the host genes associated with brain development and antiviral response in iPSC-derived human brain organoids infected with WT and sfRNA-deficient ZIKV. (<b>A</b>) Expression of host genes associated with biological processes and signaling pathways differentially affected by WT and sfRNA-deficient ZIKV. (<b>B</b>) Heat map showing association with multiple biological processes of individual differentially expressed host genes.</p>
Full article ">Figure 5
<p>Networks of interactions between the genes affected by the production of sfRNA in ZIKV infection of human brain organoids. Genes associated with each enriched functional category shown in <a href="#viruses-15-01062-f005" class="html-fig">Figure 5</a>A were used to reconstruct the networks of genetic, physical, and pathway interactions. Nod sizes indicate betweenness centrality, the logFC values are for ((WT−Mock) − (Mut−Mock)) contrast.</p>
Full article ">Figure 6
<p>Interconnection between sfRNA-affected biological processes in ZIKV-infected human brain organoids. Individual networks were reconstructed from sfRNA-affected genes related to the most enriched pathways and GO BP terms: Wnt signaling pathway, nervous system development, neuron differentiation, axon guidance, brain development, regulation of apoptotic processes, p53 signaling pathway, and nucleosome assembly. Networks were then merged using the union method and gene names as query keys. The resulting network was analyzed to determine betweenness centrality values for the nods (visualized as nod size). The logFC values are for ((WT−Mock) − (Mut−Mock)) comparison.</p>
Full article ">
14 pages, 6843 KiB  
Article
A Candidate Antigen of the Recombinant Membrane Protein Derived from the Porcine Deltacoronavirus Synthetic Gene to Detect Seropositive Pigs
by Francisco Jesus Castañeda-Montes, José Luis Cerriteño-Sánchez, María Azucena Castañeda-Montes, Julieta Sandra Cuevas-Romero and Susana Mendoza-Elvira
Viruses 2023, 15(5), 1049; https://doi.org/10.3390/v15051049 - 25 Apr 2023
Cited by 1 | Viewed by 1675
Abstract
Porcine deltacoronavirus (PDCoV) is an emergent swine coronavirus which infects cells from the small intestine and induces watery diarrhea, vomiting and dehydration, causing mortality in piglets (>40%). The aim of this study was to evaluate the antigenicity and immunogenicity of the recombinant membrane [...] Read more.
Porcine deltacoronavirus (PDCoV) is an emergent swine coronavirus which infects cells from the small intestine and induces watery diarrhea, vomiting and dehydration, causing mortality in piglets (>40%). The aim of this study was to evaluate the antigenicity and immunogenicity of the recombinant membrane protein (M) of PDCoV (rM-PDCoV), which was developed from a synthetic gene obtained after an in silico analysis with a group of 138 GenBank sequences. A 3D model and phylogenetic analysis confirmed the highly conserved M protein structure. Therefore, the synthetic gene was successfully cloned in a pETSUMO vector and transformed in E. coli BL21 (DE3). The rM-PDCoV was confirmed by SDS-PAGE and Western blot with ~37.7 kDa. The rM-PDCoV immunogenicity was evaluated in immunized (BLAB/c) mice and iELISA. The data showed increased antibodies from 7 days until 28 days (p < 0.001). The rM-PDCoV antigenicity was analyzed using pig sera samples from three states located in “El Bajío” Mexico and positive sera were determined. Our results show that PDCoV has continued circulating on pig farms in Mexico since the first report in 2019; therefore, the impact of PDCoV on the swine industry could be higher than reported in other studies. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Phylogenetic tree. (<b>a</b>) Maximum likelihood phylogenetic tree displaying the 138 PDCoV protein M sequences available in Gen Bank. The black arrow shows the consensus M-PDCoV used in this study to develop a recombinant protein. The black dots show the selection of 16 sequences with branch length differences. (<b>b</b>) A comparison of the 16 sequences showed differences at the amino acid level. The sequence CHzm2019 from China showed the longest branch in the phylogeny and showed 10 different amino acids.</p>
Full article ">Figure 2
<p>3D predicted model comparison, secondary structures, antigenic sites and surface properties analysis. (<b>a</b>) 3D model comparison of consensus M-PDCoV, Chzmd2019, USA/Minnesota292/2014, CHzmd2019, HKU15, M-PED CV777 and M-TGE Miller M6. (<b>b</b>) Secondary structure for each predicted model, red: helix, blue: strands, black: coil. (<b>c</b>) Six predicted antigenic sites predicted and their positions along the consensus M-PDCoV. (<b>d</b>) Consensus sequence surface properties indicate the hydrophobic and hydrophilic sites.</p>
Full article ">Figure 3
<p><span class="html-italic">r</span>M-PDCoV expression. (<b>a</b>) Integrity and weight verification (654 bp) of the PDCoV-M synthetic gene. (<b>b</b>) Western blot and SDS-PAGE to verify the <span class="html-italic">r</span>M-PDCoV expression. Lines 1 to 3 are samples of the culture media before the <span class="html-italic">E. coli</span> BL21 (DE3) disruption. Line 4 is a sample after the <span class="html-italic">E. coli</span> BL21 (DE3) disruption. Lines 5 and 6 are samples after centrifuging of the cultures, soluble and insoluble phases, respectively. The <span class="html-italic">r</span>M-PDCoV was found in the insoluble phase as inclusion bodies. (<b>c</b>) The <span class="html-italic">r</span>M-PDCoV confirmation after purification with Ni-NTA agarose column with His-tag affinity. Purified recombinant proteins were observed from elutions 5 to 7. The <span class="html-italic">r</span>M-PDCoV is 37.7 kDa of the expected molecular weight.</p>
Full article ">Figure 4
<p>Antigenicity assay. (<b>a</b>) Immunization scheme used to evaluate the antibody production. (<b>b</b>) Antibody production against <span class="html-italic">r</span>M-PDCoV in three BALB/c mice. The blue line indicates the <span class="html-italic">r</span>M-PDCoV + Matrix-M<sup>TM</sup> adjuvant group. The red line indicates the <span class="html-italic">r</span>M-PDCoV group. The black line indicates the mouse group immunized with only PBS. Dots and triangles indicated the days of sera collection: 7, 14, 21 and 28. The bars indicated the standard error of mean. The statistical significance was calculated using KlusKal Wallis and Dunn’s test. <span class="html-italic">p</span> &lt; 0.05 and <span class="html-italic">p</span> &lt; 0.005 is the statistical significance with a 95% confidence interval represented by * and **, respectively.</p>
Full article ">Figure 5
<p>Antigenicity evaluation of the <span class="html-italic">r</span>M-PDCoV by an iELISA assay using 44 sera samples from Guanajuato, Aguascalientes and Jalisco. Positive and negative samples are represented by black triangles and blue squares, respectively. The cut-off value is shown by the red line which corresponds to an absorbance of 0.3732 ± 3 SD and 14.3137% PP. Guanajuato: 13 positives (30.23%) and 11 negatives (25.58%); Aguascalientes: 2 positives (4.65%) and 7 negatives (16.28%); Jalisco: 8 positives (18.6%) and 3 negatives (6.98%).</p>
Full article ">
14 pages, 6160 KiB  
Article
Interactions of Tomato Chlorosis Virus p27 Protein with Tomato Catalase Are Involved in Viral Infection
by Xiaohui Sun, Lianyi Zang, Xiaoying Liu, Shanshan Jiang, Xianping Zhang, Dan Zhao, Kaijie Shang, Tao Zhou, Changxiang Zhu and Xiaoping Zhu
Viruses 2023, 15(4), 990; https://doi.org/10.3390/v15040990 - 18 Apr 2023
Cited by 2 | Viewed by 1396
Abstract
Tomato chlorosis virus (ToCV) severely threatens tomato production worldwide. P27 is known to be involved in virion assembly, but its other roles in ToCV infection are unclear. In this study, we found that removal of p27 reduced systemic infection, while ectopic expression of [...] Read more.
Tomato chlorosis virus (ToCV) severely threatens tomato production worldwide. P27 is known to be involved in virion assembly, but its other roles in ToCV infection are unclear. In this study, we found that removal of p27 reduced systemic infection, while ectopic expression of p27 promoted systemic infection of potato virus X in Nicotiana benthamiana. We determined that Solanum lycopersicum catalases (SlCAT) can interact with p27 in vitro and in vivo and that amino acids 73 to 77 of the N-terminus of SlCAT represent the critical region for their interaction. p27 is distributed in the cytoplasm and nucleus, and its coexpression with SlCAT1 or SlCAT2 changes its distribution in the nucleus. Furthermore, we found that silencing of SlCAT1 and SlCAT2 can promote ToCV infection. In conclusion, p27 can promote viral infection by binding directly to inhibit anti-ToCV processes mediated by SlCAT1 or SlCAT2. Full article
(This article belongs to the Special Issue State-of-the-Art Plant Viruses Research in Asia)
Show Figures

Figure 1

Figure 1
<p>P27 promotes ToCV systemic infection in <span class="html-italic">N. benthamiana</span> plants. (<b>a</b>) Schematic diagram of the structure of the ToCV p27 base mutation. (<b>b</b>) Diagram of results of the sequencing of the p27 termination. (<b>c</b>) Schematic diagram of the genomic structure of ToCV-WT and ToCV-p27X. (<b>d</b>) Schematic map of the genomic structure of <span class="html-italic">N. benthamiana</span> infected with the ToCV-p27X system. (<b>e</b>) Four weeks after inoculation, WT, mutant, and anaplerosis ToCV caused symptoms in <span class="html-italic">N. benthamiana</span>. (<b>f</b>) Four weeks after ToCV immunosuppression, the expression of the ToCV CP protein in the leaves of the <span class="html-italic">N. benthamiana</span> system was detected by Western blot analysis. (<b>g</b>) Four weeks after inoculation, the accumulation of ToCV CP in <span class="html-italic">N. benthamiana</span> was detected by qRT-PCR.</p>
Full article ">Figure 2
<p>Expression of the ToCV p27 protein in <span class="html-italic">N. benthamiana</span> after inoculation with pGR106-p27. (<b>a</b>) On the 14th day after inoculation, the symptoms of <span class="html-italic">N. benthamiana</span> were inoculated with pGR106 and the pGR106-p27 virus. (<b>b</b>) Expression of the ToCV p27 protein in <span class="html-italic">N. benthamiana</span> 14 days after inoculation with pGR106-p27. (<b>c</b>) The accumulation of PVX <span class="html-italic">CP</span> in <span class="html-italic">N. benthamiana</span> was determined by qRT-PCR at 14 dpi. Error bars indicate standard deviation and statistical significance was calculated with Student’s <span class="html-italic">t</span>-test (*** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 3
<p>ToCV p27 interacts with itself and with SlCAT1 and SlCAT2. (<b>a</b>) Y2H verified the interaction of ToCV p27 with SlCAT1. (<b>b</b>) Y2H verified the interaction of ToCV p27 with SlCAT2. (<b>c</b>) BiFC verification shows that ToCV p27 interacts with SlCAT1 and SlCAT2. (<b>d</b>) Co-IP verified that ToCV p27 interacts with SlCAT1 and SlCAT2.</p>
Full article ">Figure 4
<p>Verification of the functional region of the yeast cotransformation and the p27 interaction. (<b>a</b>) Prediction of the secondary structure of SlCAT1. (<b>b</b>) Prediction of the SlCAT1 domain, and Y2H was used to verify the interaction between p27 and the SlCAT1 deletion mutants.</p>
Full article ">Figure 5
<p>The expression of SlCAT1 and SlCAT2 changes the localisation of p27.</p>
Full article ">Figure 6
<p>TRV-mediated silencing of the <span class="html-italic">NbCAT1</span> gene promotes ToCV infection of host plants. (<b>a</b>) Phenotype of <span class="html-italic">N. benthamiana</span> with TRV-mediated silencing of the <span class="html-italic">NbCAT1</span> and <span class="html-italic">NbCAT2</span> gene. (<b>b</b>) Detection of the silencing efficiency of the <span class="html-italic">NbCAT1</span> gene in <span class="html-italic">N. benthamiana</span> by qRT-PCR at 10 dpi. (<b>c</b>) Detection of the silencing efficiency of the <span class="html-italic">NbCAT2</span> gene in <span class="html-italic">N. benthamiana</span> by qRT-PCR at 10 dpi. (<b>d</b>) The ToCV symptom in <span class="html-italic">N. benthamiana</span> silent in <span class="html-italic">NbCAT1.</span> (<b>e</b>) The ToCV symptom in <span class="html-italic">N. benthamiana</span> silent of <span class="html-italic">NbCAT2.</span> (<b>f</b>) The relative accumulation of ToCV <span class="html-italic">CP</span> in <span class="html-italic">N. benthamiana</span> without <span class="html-italic">NbCAT1</span> by qRT-PCR. (<b>g</b>) The relative accumulation of ToCV <span class="html-italic">CP</span> in <span class="html-italic">N. benthamiana</span> silent of <span class="html-italic">NbCAT2</span> by qRT-PCR. (<b>h</b>) Western blot detection showing the expression of the ToCV CP protein in the upper leaves of <span class="html-italic">N. benthamiana</span> silent of <span class="html-italic">NbCAT1</span> at 20 dpi. (<b>i</b>) Western blot detection showing the expression of the ToCV CP protein in the upper leaves of <span class="html-italic">N. benthamiana</span> silent of <span class="html-italic">NbCAT1</span> at 20 dpi. Error bars indicate standard deviation and statistical significance was calculated with Student’s <span class="html-italic">t</span>-test (*** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">
13 pages, 2918 KiB  
Article
Development of a Pan-Filoviridae SYBR Green qPCR Assay for Biosurveillance Studies in Bats
by Jessica Coertse, Marinda Mortlock, Antoinette Grobbelaar, Naazneen Moolla, Wanda Markotter and Jacqueline Weyer
Viruses 2023, 15(4), 987; https://doi.org/10.3390/v15040987 - 17 Apr 2023
Viewed by 1837
Abstract
Recent studies have indicated that bats are hosts to diverse filoviruses. Currently, no pan-filovirus molecular assays are available that have been evaluated for the detection of all mammalian filoviruses. In this study, a two-step pan-filovirus SYBR Green real-time PCR assay targeting the nucleoprotein [...] Read more.
Recent studies have indicated that bats are hosts to diverse filoviruses. Currently, no pan-filovirus molecular assays are available that have been evaluated for the detection of all mammalian filoviruses. In this study, a two-step pan-filovirus SYBR Green real-time PCR assay targeting the nucleoprotein gene was developed for filovirus surveillance in bats. Synthetic constructs were designed as representatives of nine filovirus species and used to evaluate the assay. This assay detected all synthetic constructs included with an analytical sensitivity of 3–31.7 copies/reaction and was evaluated against the field collected samples. The assay’s performance was similar to a previously published probe based assay for detecting Ebola- and Marburgvirus. The developed pan-filovirus SYBR Green assay will allow for more affordable and sensitive detection of mammalian filoviruses in bat samples. Full article
(This article belongs to the Special Issue Bat-Borne Viruses Research)
Show Figures

Figure 1

Figure 1
<p>Graphic representation of the synthetic constructs designed for use as assay controls. Each construct consists of a 537-nucleotide partial nucleoprotein sequence of the selected viral species (<a href="#viruses-15-00987-t001" class="html-table">Table 1</a>). The expected amplicon size for the filovirus products was 157 bp. The control tag sequence was situated at position 34–51 within the assay amplification region. The SP6 promotor was located at the end of the synthetic construct to allow for the synthesis of RNA transcripts.</p>
Full article ">Figure 2
<p>Filovirus SYBR Green qPCR data analysis workflow. The diagram depicts the workflow to assess individual samples at several steps to determine the additional analyses required for suspected positive samples. Results indicated in light blue represent negative samples and no further analyses were needed. Findings indicated in light green require further data analyses indicated with arrows. *: Numerical values of the C<sub>q</sub>/Tm provided; #: Visual representation of the amplification/melt curve plots; <span>$</span>: Analyses of nucleotide sequences using the BLAST function of the National Center for Biotechnology Information (NCBI) database (accessible online at <a href="https://blast.ncbi.nlm.nih.gov/Blast.cgi" target="_blank">https://blast.ncbi.nlm.nih.gov/Blast.cgi</a> (accessed on 10 January 2023)).</p>
Full article ">Figure 3
<p>Average melting temperatures of the amplified filovirus standard templates (10<sup>0</sup>–10<sup>9</sup> copies/reaction). Error bars indicate the standard deviation.</p>
Full article ">Figure 4
<p>Standard curves for the serially diluted (10<sup>5</sup>–10<sup>9</sup> copies/reaction) standard templates.</p>
Full article ">Figure 5
<p>Comparison of the Cq values for the probe-based assay and the average Cq values for the SYBR Green assay for the titrated (TCID<sub>50</sub>/mL) serially diluted Ebola virus.</p>
Full article ">Figure 6
<p>Comparison of the Cq values for the probe-based assay and the average Cq values for the SYBR Green assay for the titrated (TCID<sub>50</sub>/mL) serially diluted Marburg virus.</p>
Full article ">Figure 7
<p>Comparison of the Cq values for the probe-based assay and the average Cq values for the SYBR Green assay for the negative bat sera spiked with the titrated (TCID<sub>50</sub>/mL) serially diluted Ebola virus.</p>
Full article ">Figure 8
<p>Comparison of the Cq values for the probe-based assay and the average Cq values for the SYBR Green assay for negative bat sera spiked with the titrated (TCID<sub>50</sub>/mL) serially diluted Margburg virus.</p>
Full article ">
17 pages, 4182 KiB  
Article
Receptor Binding-Induced Conformational Changes in Herpes Simplex Virus Glycoprotein D Permit Interaction with the gH/gL Complex to Activate Fusion
by Doina Atanasiu, Wan Ting Saw, Tina M. Cairns, Harvey M. Friedman, Roselyn J. Eisenberg and Gary H. Cohen
Viruses 2023, 15(4), 895; https://doi.org/10.3390/v15040895 - 30 Mar 2023
Cited by 2 | Viewed by 1931
Abstract
Herpes simplex virus (HSV) requires four essential virion glycoproteins—gD, gH, gL, and gB—for virus entry and cell fusion. To initiate fusion, the receptor binding protein gD interacts with one of two major cell receptors, HVEM or nectin-1. Once gD binds to a receptor, [...] Read more.
Herpes simplex virus (HSV) requires four essential virion glycoproteins—gD, gH, gL, and gB—for virus entry and cell fusion. To initiate fusion, the receptor binding protein gD interacts with one of two major cell receptors, HVEM or nectin-1. Once gD binds to a receptor, fusion is carried out by the gH/gL heterodimer and gB. A comparison of free and receptor-bound gD crystal structures revealed that receptor binding domains are located within residues in the N-terminus and core of gD. Problematically, the C-terminus lies across and occludes these binding sites. Consequentially, the C-terminus must relocate to allow for both receptor binding and the subsequent gD interaction with the regulatory complex gH/gL. We previously constructed a disulfide bonded (K190C/A277C) protein that locked the C-terminus to the gD core. Importantly, this mutant protein bound receptor but failed to trigger fusion, effectively separating receptor binding and gH/gL interaction. Here, we show that “unlocking” gD by reducing the disulfide bond restored not only gH/gL interaction but fusion activity as well, confirming the importance of C-terminal movement in triggering the fusion cascade. We characterize these changes, showing that the C-terminus region exposed by unlocking is: (1) a gH/gL binding site; (2) contains epitopes for a group (competition community) of monoclonal antibodies (Mabs) that block gH/gL binding to gD and cell–cell fusion. Here, we generated 14 mutations within the gD C-terminus to identify residues important for the interaction with gH/gL and the key conformational changes involved in fusion. As one example, we found that gD L268N was antigenically correct in that it bound most Mabs but was impaired in fusion, exhibited compromised binding of MC14 (a Mab that blocks both gD–gH/gL interaction and fusion), and failed to bind truncated gH/gL, all events that are associated with the inhibition of C-terminus movement. We conclude that, within the C-terminus, residue 268 is essential for gH/gL binding and induction of conformational changes and serves as a flexible inflection point in the critical movement of the gD C-terminus. Full article
(This article belongs to the Special Issue Research on Herpes Virus Fusion and Entry)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Surface representation of gD306 crystal structure (PDB 2c36) showing the C-term (dark gray), the nectin-1 biding site (pink), and two of the three gH/gL binding sites defined by MC5 (site 2, blue) and MC14 (site 3, brown). gH/gL binding site 1 overlaps with the nectin-1 binding site. Region 257-267 which is missing from all crystal structures is shown as a dotted line and the C-term, lying over the receptor binding site. Point mutations generated in the gH/gL binding site 3 are shown as labelled side chains. In green are the point mutations that introduced the cysteine residues (at positions 190 and 277), which allow for the formation of a disulfide bond. (<b>b</b>) Based on competition and biological function, gD Mabs were grouped into communities: blue, brown, green, and red. The red community is subdivided into red and pink. The green community is subdivided into green and yellow. Circles indicate that competition was measured as both a ligand and an analyte; squares indicate that competition was measured as either a ligand or an analyte only. Solid connecting lines specify that competition between the two Mabs was measured in both directions (each as a ligand and analyte). Dashed connecting lines identify that the competition between Mabs was measured in one direction only.</p>
Full article ">Figure 2
<p>Binding of gHgL to gD by SPR. (<b>a</b>). Diagram outlining the SPR protocol. (<b>b</b>)<b>.</b> 1D3 (anti-gD Mab) was covalently coupled to a CM5 biosensor chip. An equal amount of gD<sub>285t</sub> (gray) and gD<sub>306t</sub> (black) were immobilized to different spots on the Biacore chip. Soluble gD<sub>306t</sub> or gD<sub>285t</sub> and gH/gL were sequentially injected (injection start indicated by diamond arrow). An increase in response units (RUs) after injection indicates binding. gH/gL binds to gD<sub>285t</sub> only (black curve). No gH/gL binding was observed for gD<sub>306t</sub> (gray curve). <span class="html-italic">x</span> axis, time (seconds). <span class="html-italic">y</span> axis, resonant units (RU). (<b>c</b>) Diagram of the SPR protocol for sequential injections of gD<sub>306t</sub> and nectin-1, followed by injection of soluble gH/gL. (<b>d</b>) gD306 was attached to two flow cells on the biosensor chip via 1D3 Mab. Nectin-1 was flowed over cell 1 only. Flow cell 2 did not receive nectin. gH/gL was flowed over both flow cells. In the absence of nectin-1, gD<sub>306t</sub> does not bind gH/gL (black curves). Pre-binding of nectin-1 to gD<sub>306t</sub> induced conformational changes that allow for gH/gL binding (purple). Experiments were performed a minimum of three times.</p>
Full article ">Figure 3
<p>Characterization of wt and cys2 proteins. (<b>a</b>) CELISA of B78 cells transfected with full-length wt and cys2 gD and probed with select Mabs. Expression of cys2 (brown bars) is presented as % of binding to wt with the same antibody. The average of three experiments, each done in duplicate. (<b>b</b>) Fusion assay. Ability of full-length gD wt (black curve) and cys2 (brown curve) to trigger fusion in a live cell fusion assay. Data are presented as percent fusion by wt gD at the 2 h time point. Representative curve from three independent experiments, each done in duplicate. (<b>c</b>) CELISA after TCEP treatment. Binding of sentinel Mabs from each community to B78 cells transfected with full-length wt (gray) or cys2 (yellow) gD after cells were treated with TCEP for 10 min. Average of three experiments, each done in duplicate. Data for each antibody was normalized to no treatment wt samples. (<b>d</b>) Fusion assay. Effector cells expressing gB, gH/gL, and wt or cys2 gD were treated for 10 min with 10 mM TCEP before fusion was triggered by the addition of donor nectin-1 expressing cells. Data are presented as percent fusion by untreated wt gD at the 2-h time point, when the rate of fusion was at its maximum. Representative curve from three independent experiments, each done in duplicate.</p>
Full article ">Figure 4
<p>Characterization of wt and cys2 proteins. (<b>a</b>) Diagram of the Biacore 3000 protocol for analyzing gH/gL binding to gD, wt or cys2. (<b>b</b>) Soluble wt<sub>306t</sub> and cys2<sub>306t</sub> proteins were attached to the biosensor chip via 1D3 Mab. Soluble gH/gL was injected across the chip surface. Only the gH/gL binding curves are shown. Black curve, binding of gH/gL to wt gD; brown curve, binding of gH/gL to cys2. (<b>c</b>) Diagram of the Biacore 3000 protocol for analyzing the binding of gH/gL to a gD-nectin-1 receptor complex. gD molecules were attached to the biosensor chip via 1D3 Mab. Sequential injection of nectin-1 followed by injection of soluble gH/gL. (<b>d</b>) Nectin-1 binds to both wt (dark purple) and cys2 (light purple) gD proteins. (<b>e</b>) gH/gL bound to wt gD (blue) but not to cys2 mutant gD (cyan). (<b>f</b>) Diagram of the Biacore 3000 protocol for analyzing gD–gH/gL binding in the presence of TCEP. gD molecules were attached to the biosensor chip via 1D3 Mab. Nectin-1 was pre-mixed with 1 mM TCEP and flowed over the captured gD molecules, followed by the injection of soluble gH/gL. (<b>g</b>) Nectin-1 binds to both wt (purple) and cys2 (light purple) gD proteins. (<b>h</b>) As expected, gH/gL bound to wt gD (blue). Due to the reduction of the disulfide bond, gH/gL is now able to bind cys2 mutant gD as well (cyan).</p>
Full article ">Figure 5
<p>Characterization of point mutations in region 262–280 of full-length gD. (<b>a</b>) CELISA. Detection of the surface-expressed wt and point gD mutations with sentinel Mabs from each community. Data normalized to wt with each Mab (dotted line). Average of three experiments, each done in duplicate. (<b>b</b>) Fusion function of full-length wt and mutant gDs when combined with gB, gH, and gL constructs. Data normalized to wt gD (dotted line). Average of three independent experiments, each in duplicate.</p>
Full article ">Figure 6
<p>Ability of gD L268N to bind gH/gL. (<b>a</b>) Diagram of the Biacore 3000 protocol for analyzing the binding of gH/gL to a gD-receptor complex. gD molecules, wt, or L268N, both as 306t truncations were attached to the biosensor chip via 1D3 Mab. Sequential injection of nectin-1 followed by injection of soluble gH/gL. (<b>b</b>) Nectin-1 binds to both wt (purple) and L268N (pink) gD proteins. (<b>c</b>) gH/gL was flowed over the gD-nectin-1 complexes. gH/gL bound to wt gD (blue) but not to L268N mutant gD (cyan). (<b>d</b>) Diagram of the SPR protocol for analyzing gH/gL binding to gD<sub>285t</sub>, wt, or the L268N mutant. Only the gH/gL binding curves are shown. gH/L binds to wt gD (black curve). No gH/gL binding was observed for the L268N mutant (orange curve).</p>
Full article ">Figure 7
<p>Proposed intermediates for the transition of gD from a pre-receptor (closed) to post-receptor binding state (open). (<b>a</b>) Before binding to a receptor, gD is in a closed conformation, with the receptor binding site (purple) covered by the C-term (black). The three gH/gL binding sites are also shown in red (site 1), blue (site 2), and brown (site 3). (<b>b</b>) In the presence of receptor or antibodies that compete with the receptor for binding, the C-term is displaced to expose the receptor binding site. Residue 277 is one inflection point in the C-term that is sufficient for receptor binding, but not gH/gL interaction. (<b>c</b>) Access to gH/gL binding site 3 (brown) is controlled by a second inflection point at position 268. Mutagenesis of this residue results in a molecule that partially exposes the gH/gL site 3. (<b>d</b>) After opening of gD past residue 268, gD is in a fully open conformation that can bind gH/gL.</p>
Full article ">
23 pages, 11375 KiB  
Article
Development and Characterization of Efficient Cell Culture Systems for Genotype 1 Hepatitis E Virus and Its Infectious cDNA Clone
by Putu Prathiwi Primadharsini, Shigeo Nagashima, Toshinori Tanaka, Suljid Jirintai, Masaharu Takahashi, Kazumoto Murata and Hiroaki Okamoto
Viruses 2023, 15(4), 845; https://doi.org/10.3390/v15040845 - 26 Mar 2023
Cited by 4 | Viewed by 2257
Abstract
Hepatitis E virus (HEV) is a major cause of acute viral hepatitis globally. Genotype 1 HEV (HEV-1) is responsible for multiple outbreaks in developing countries, causing high mortality rates in pregnant women. However, studies on HEV-1 have been hindered by its poor replication [...] Read more.
Hepatitis E virus (HEV) is a major cause of acute viral hepatitis globally. Genotype 1 HEV (HEV-1) is responsible for multiple outbreaks in developing countries, causing high mortality rates in pregnant women. However, studies on HEV-1 have been hindered by its poor replication in cultured cells. The JE04-1601S strain recovered from a Japanese patient with fulminant hepatitis E who contracted HEV-1 while traveling to India was serially passaged 12 times in human cell lines. The cell-culture-generated viruses (passage 12; p12) grew efficiently in human cell lines, but the replication was not fully supported in porcine cells. A full-length cDNA clone was constructed using JE04-1601S_p12 as a template. It was able to produce an infectious virus, and viral protein expression was detectable in the transfected PLC/PRF/5 cells and culture supernatants. Consistently, HEV-1 growth was also not fully supported in the cell culture of cDNA-derived JE04-1601S_p12 progenies, potentially recapitulating the narrow tropism of HEV-1 observed in vivo. The availability of an efficient cell culture system for HEV-1 and its infectious cDNA clone will be useful for studying HEV species tropism and mechanisms underlying severe hepatitis in HEV-1-infected pregnant women as well as for discovering and developing safer treatment options for this condition. Full article
(This article belongs to the Special Issue Molecular Biology of RNA Viruses)
Show Figures

Figure 1

Figure 1
<p>Quantification of HEV RNA in culture supernatants of PLC/PRF/5 (for passage 0, p0) inoculated with serum sample of JE04-1601S/wild-type and in culture supernatants of A549_1-1H8 cells inoculated with culture supernatants of p0, p1, p2, p3, p4, p5, p6, p7, p8, p9, p10, or p11 that were harvested on the final day of each passage (see <a href="#viruses-15-00845-t001" class="html-table">Table 1</a>). The harvested culture supernatant of each passage was purified by passing through a microfilter with a pore size of 0.22 µm (see Materials and Methods) and then inoculated onto A549_1-1H8 cells. The dotted horizontal line represents the limit of detection by real-time RT-PCR used in the current study, at 2.0 × 10<sup>1</sup> RNA copies/mL. Each passage was performed for three wells and one representative well showing median viral titer at 10 and 18 days postinoculation was selected, and culture media collected serially from the selected well were subjected to quantification of HEV RNA.</p>
Full article ">Figure 2
<p>A phylogenetic tree constructed using the neighbor-joining tree of Jukes–Cantor distances based on the entire genomic sequences of three JE04-1601S strains obtained in the present study (JE04-1601S_wild-type (wt), JE04-1601S_p10, and JE04-1601S_p12), all known genotype 1 strains (1a, n = 13; 1b, n = 10; 1c, n = 2; 1d, n = 1; 1e, n = 1; 1f, n = 32; 1g, n = 18; and unclassified subtypes, n = 2), and each one of the prototype strains of genotypes 2–8. The three JE04-1601S strains obtained in the present study are highlighted with closed circles for clarity. Each reference sequence is shown with the genotype/subtype, followed by the accession number and the name of the country in which it was detected. The bootstrap values (≥70%) of the nodes are indicated as a percentage of data obtained from 1000 resamplings. Tips are collapsed for 21 Bangladeshi 1f strains with similar sequences. The scale bar (0.05) represents the number of nucleotide substitutions per site.</p>
Full article ">Figure 3
<p>Quantification of HEV RNA in culture supernatants of human-derived cell lines (PLC/PRF/5, A549_1-1H8, and HepG2/C3A cells; indicated with continuous lines) and in culture supernatants of porcine-kidney-derived cell lines (PK15, IBRS-2, and LLC-PK1 cells; indicated with dotted lines) inoculated with the HEV-3 (JE03-1760F_p26) strain (<b>A</b>), the HEV-4 (HE-JF5/15F_p24) strain (<b>B</b>), or the HEV-1 (JE04-1601S_p12) strain (<b>C</b>) at a titer of 1.0 × 10<sup>5</sup> copies/well (left panels) or 1.0 × 10<sup>6</sup> copies/well (right panels) in six-well plates. The virus growth was observed for 60 days. The dotted horizontal line represents the limit of detection by real-time RT-PCR used in the current study, at 2.0 × 10<sup>1</sup> RNA copies/mL. Each inoculation was performed for three wells, one representative well showing median viral titer at 20 and 60 days postinoculation was selected, and culture media collected serially from the selected well were subjected to quantification of HEV RNA.</p>
Full article ">Figure 4
<p>A schematic representation of the full-length genome of the JE04-1601S_p12 strain (<b>A</b>) and the strategy to construct its full-length cDNA clone (pJE04-1601S_p12) (<b>B</b>). Three fragments covering its whole genome were generated by reverse transcription-polymerase chain reaction (RT-PCR) and then cloned into the pUC19 vector in a stepwise manner using the In-Fusion cloning method. The 15-bp overlaps at their ends are highlighted with same colors. MeT, methyltransferase; Y, Y domain; PCP, papain-like cysteine protease; HVR, hypervariable region; X, X or macro domain; Hel, helicase; and RdRp, RNA-dependent RNA polymerase.</p>
Full article ">Figure 5
<p>Capability of the cDNA clone of JE04-1760S_p12 to produce infectious progeny viruses. (<b>A</b>) Quantification of HEV RNA in culture supernatants. RNA transcript of pJE04-1601S_p12 was transfected to PLC/PRF/5 cells, along with RNA transcript of its replication-defective mutant (pJE04-1601S_p12-GAA), which served as a negative control. HEV growth was observed for 28 days. The data are presented as the mean ± standard deviation (SD) for two wells each. The dotted horizontal line represents the limit of detection by real-time RT-PCR used in the current study, at 2.0 × 10<sup>1</sup> RNA copies/mL. RNA transfection experiment was performed twice for two wells each, and representative result was shown. (<b>B</b>) A Western blot analysis of the culture supernatants transfected with RNA transcript of pJE04-1601S_p12 or that of pJE04-1601S_p12-GAA to examine the expression of HEV ORF2 (upper panel) and ORF3 proteins (lower panel) at day 28 posttransfection. (<b>C</b>) Immunofluorescence staining of the cells transfected with the RNA transcript of pJE04-1601S_p12 (upper panel) or that of pJE04-1601S_p12-GAA (lower panel) to examine the HEV ORF2 protein expression at day 28 posttransfection. For Western blotting and immunofluorescence assay, results representative of two experiments are shown.</p>
Full article ">Figure 6
<p>Species tropism of HEV-1 to humans in cell culture of the cDNA-derived JE04-1601S_p12 progeny viruses. Quantification of HEV RNA in culture supernatants of the human-derived cell lines (left panels) PLC/PRF/5 (<b>A</b>) and A549_1-1H8 (<b>B</b>) cells as well as the porcine-kidney-derived cell lines (right panels) PK15 (<b>C</b>), IBRS-2 (<b>D</b>), and LLC-PK1 (<b>E</b>) cells inoculated with cDNA-derived JE04-1601S_p12 progeny viruses. Inoculum titers were 1.0 × 10<sup>5</sup> copies/well or 1.0 × 10<sup>6</sup> copies/well in six-well plates. HEV growth was observed for 28 days. The data are presented as the mean ± SD for three wells each. The dotted horizontal line represents the limit of detection by real-time RT-PCR used in the current study, at 2.0 × 10<sup>1</sup> RNA copies/mL. Each inoculation was of single experiment for three wells.</p>
Full article ">Figure 7
<p>Sensitivity of HEV-1 to ribavirin in the cell culture system. (<b>A</b>) Quantification of HEV RNA in culture supernatants of PLC/PRF/5 cells inoculated with cDNA-derived JE04-1601S_p12 progeny viruses (1.0 × 10<sup>5</sup> copies/well) in the presence of 40 or 160 µM ribavirin in DMSO (final concentration, 1%). HEV kinetics were observed for 28 days. The data are presented as the mean ± SD for three wells each. The dotted horizontal line represents the limit of detection by real-time RT-PCR used in the current study, at 2.0 × 10<sup>1</sup> RNA copies/mL. The inoculation was of single experiment for three wells. (<b>B</b>) Immunofluorescence staining of PLC/PRF/5 cells inoculated with cDNA-derived JE04-1601S_p12 progenies in the presence of 160 µM ribavirin (right panel) to examine the HEV ORF2 protein expression at day 28 postinoculation in comparison to the ORF2 protein expression in untreated control cells (left panel). Results representative of two experiments are shown.</p>
Full article ">
23 pages, 6554 KiB  
Article
Identifying Putative Resistance Genes for Barley Yellow Dwarf Virus-PAV in Wheat and Barley
by Glenda Alquicer, Emad Ibrahim, Midatharahally N. Maruthi and Jiban Kumar Kundu
Viruses 2023, 15(3), 716; https://doi.org/10.3390/v15030716 - 9 Mar 2023
Cited by 1 | Viewed by 2358
Abstract
Barley yellow dwarf viruses (BYDVs) are one of the most widespread and economically important plant viruses affecting many cereal crops. Growing resistant varieties remains the most promising approach to reduce the impact of BYDVs. A Recent RNA sequencing analysis has revealed potential genes [...] Read more.
Barley yellow dwarf viruses (BYDVs) are one of the most widespread and economically important plant viruses affecting many cereal crops. Growing resistant varieties remains the most promising approach to reduce the impact of BYDVs. A Recent RNA sequencing analysis has revealed potential genes that respond to BYDV infection in resistant barley genotypes. Together with a comprehensive review of the current knowledge on disease resistance in plants, we selected nine putative barley and wheat genes to investigate their involvement in resistance to BYDV-PAV infection. The target classes of genes were (i) nucleotide binding site (NBS) leucine-rich repeat (LRR), (ii) coiled-coil nucleotide-binding leucine-rich repeat (CC-NB-LRR), (iii) LRR receptor-like kinase (RLK), (iv) casein kinase, (v) protein kinase, (vi) protein phosphatase subunits and the transcription factors (TF) (vii) MYB TF, (viii) GRAS (gibberellic acid-insensitive (GAI), repressor of GAI (RGA) and scarecrow (SCR)), and (ix) the MADS-box TF family. Expression of genes was analysed for six genotypes with different levels of resistance. As in previous reports, the highest BYDV-PAV titre was found in the susceptible genotypes Graciosa in barley and Semper and SGS 27-02 in wheat, which contrast with the resistant genotypes PRS-3628 and Wysor of wheat and barley, respectively. Statistically significant changes in wheat show up-regulation of NBS-LRR, CC-NBS-LRR and RLK in the susceptible genotypes and down-regulation in the resistant genotypes in response to BYDV-PAV. Similar up-regulation of NBS-LRR, CC-NBS-LRR, RLK and MYB TF in response to BYDV-PAV was also observed in the susceptible barley genotypes. However, no significant changes in the expression of these genes were generally observed in the resistant barley genotypes, except for the down-regulation of RLK. Casein kinase and Protein phosphatase were up-regulated early, 10 days after inoculation (dai) in the susceptible wheat genotypes, while the latter was down-regulated at 30 dai in resistant genotypes. Protein kinase was down-regulated both earlier (10 dai) and later (30 dai) in the susceptible wheat genotypes, but only in the later dai in the resistant genotypes. In contrast, GRAS TF and MYB TF were up-regulated in the susceptible wheat genotypes while no significant differences in MADS TF expression was observed. Protein kinase, Casein kinase (30 dai), MYB TF and GRAS TF (10 dai) were all up-regulated in the susceptible barley genotypes. However, no significant differences were found between the resistant and susceptible barley genotypes for the Protein phosphatase and MADS FT genes. Overall, our results showed a clear differentiation of gene expression patterns in both resistant and susceptible genotypes of wheat and barley. Therefore, further research on RLK, NBS-LRR, CC-NBS-LRR, GRAS TF and MYB TF can lead to BYDV-PAV resistance in cereals. Full article
(This article belongs to the Section Viruses of Plants, Fungi and Protozoa)
Show Figures

Figure 1

Figure 1
<p>BYDV-PAV titre (virus copy number) in barley genotypes (<b>A</b>) and wheat genotypes (<b>B</b>). Significant differences are shown after one-way ANOVA and Tukey’s multiple comparison test, (* = <span class="html-italic">p</span> &lt; 0.05 ** = <span class="html-italic">p</span> &lt; 0.01 and *** = <span class="html-italic">p</span> &lt; 0.001). Bars represent the means and standard errors of three biological replicates.</p>
Full article ">Figure 2
<p>Heat map showing expression profile of genes associated with resistance to BYDV-PAV infection in barley genotypes (<b>A</b>) and wheat genotypes (<b>B</b>). The blue colour stands for upregulation, the green for downregulation and the grey for control values (1). Up-regulation is mainly observed in susceptible genotypes (<b>left</b>), while down-regulation is more common in resistant genotypes (<b>right</b>). These patterns are particularly consistent for NBS, CC-NBS-LRR and Rec Kin in the inoculated samples of barley and wheat. Slight differences between 10 dai and 30 dai are also observed. The maps were generated with Graphpad Prism software using expression fold change values obtained from qPCR, followed by statistical analysis using two-way ANOVA.</p>
Full article ">Figure 3
<p>Box plots showing fold change in NBS-LRR resistance genes expression in different barley and wheat genotypes at 10 and 30 dai. The horizontal line in each box represents the mean expression fold change, and the lower and upper box limits the minimum and maximum values, respectively. Two-way ANOVA and Tukey’s multiple comparison test revealed significant differences * = <span class="html-italic">p</span> &lt; 0.05, ** = <span class="html-italic">p</span> &lt; 0.01, *** = <span class="html-italic">p</span> &lt; 0.001, and **** = <span class="html-italic">p</span> &lt; 0.0001. The straight arrow from Semper to PSR 3628 indicates the same<span class="html-italic">p</span>value when Semper is compared to all genotypes.</p>
Full article ">Figure 4
<p>Box plots showing the expression fold change of CC-NBS-LRR in different barley wheat genotypes at 10 and 30 dai. The horizontal line in each box represents the mean expression fold change, and the lower and upper box limits the minimum and maximum values, respectively. Two-way ANOVA and Tukey’s multiple comparison test revealed significant differences * = <span class="html-italic">p</span> &lt; 0.05, ** = <span class="html-italic">p</span> &lt; 0.01, *** = <span class="html-italic">p</span> &lt; 0.001, and **** = <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 5
<p>Box plots showing the expression fold change of LRR receptor-like kinase in different barley and wheat genotypes at 10 and 30 dai. The horizontal line in each box represents the mean expression fold change, and the lower and upper box limits the minimum and maximum values, respectively. Two-way ANOVA and Tukey’s multiple comparison test revealed significant differences * = <span class="html-italic">p</span> &lt; 0.05, ** = <span class="html-italic">p</span> &lt; 0.01, *** = <span class="html-italic">p</span> &lt; 0.001, and **** = <span class="html-italic">p</span> &lt; 0.0001. The straight line at the top indicates the same<span class="html-italic">p</span>value when that genotype is compared to the rest in line.</p>
Full article ">Figure 6
<p>Box plots showing the expression fold change of Casein kinase in different barley and wheat genotypes at 10 and 30 dai. The horizontal line in each box represents the mean expression fold change, and the lower and upper box limits the minimum and maximum values, respectively. Two-way ANOVA and Tukey’s multiple comparison test revealed significant differences * = <span class="html-italic">p</span> &lt; 0.05, ** = <span class="html-italic">p</span> &lt; 0.01, *** = <span class="html-italic">p</span> &lt; 0.001, and **** = <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 7
<p>Box plots showing the expression fold change of <span class="html-italic">Protein kinase</span> in different barley and wheat genotypes at 10 and 30 dai. The horizontal line in each box represents the mean expression fold change, and the lower and upper box limits the minimum and maximum values, respectively. Two-way ANOVA and Tukey’s multiple comparison test revealed significant differences * = <span class="html-italic">p</span> &lt; 0.05, ** = <span class="html-italic">p</span> &lt; 0.01, *** = <span class="html-italic">p</span> &lt; 0.001, and **** = <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 8
<p>Box plots showing the expression fold change of Protein phosphatase in different barley and wheat genotypes at 10 and 30 dai. The horizontal line in each box represents the mean expression fold change, and the lower and upper limits the minimum and maximum values, respectively. Two-way ANOVA and Tukey’s multiple comparison test revealed significant differences * = <span class="html-italic">p</span> &lt; 0.05 ** = <span class="html-italic">p</span> &lt; 0.01 and **** = <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 9
<p>Box plots showing the expression fold change of <span class="html-italic">MYB TF</span> in different barley wheat genotypes at 10 and 30 dai. The horizontal line in each box represents the mean expression fold change, and the lower and upper box limits the minimum and maximum values, respectively. Two-way ANOVA and Tukey’s multiple comparison test revealed significant differences * = <span class="html-italic">p</span> &lt; 0.05, ** = <span class="html-italic">p</span> &lt; 0.01, *** = <span class="html-italic">p</span> &lt; 0.001, and **** = <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 10
<p>Box plots showing the expression fold change of <span class="html-italic">GRAS TF</span> in different barley and wheat genotypes at 10 and 30 dai. The horizontal line in each box represents the mean expression fold change, and the lower and upper box limits the minimum and maximum values, respectively. Two-way ANOVA and Tukey’s multiple comparison test revealed significant differences * = <span class="html-italic">p</span> &lt; 0.05, ** = <span class="html-italic">p</span> &lt; 0.01, *** = <span class="html-italic">p</span> &lt; 0.001, and **** = <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 11
<p>Box plots showing the expression fold change of MADS box transcription factor in different barley and wheat genotypes at 10 and 30 dai. The horizontal line in each box represents the mean expression fold change, and the lower and upper box limits the minimum and maximum values, respectively. Two-way ANOVA and Tukey’s multiple comparison test did not revealed any significant differences.</p>
Full article ">
24 pages, 5357 KiB  
Article
Efficient Purification of Cowpea Chlorotic Mottle Virus by a Novel Peptide Aptamer
by Georg Tscheuschner, Marco Ponader, Christopher Raab, Prisca S. Weider, Reni Hartfiel, Jan Ole Kaufmann, Jule L. Völzke, Gaby Bosc-Bierne, Carsten Prinz, Timm Schwaar, Paul Andrle, Henriette Bäßler, Khoa Nguyen, Yanchen Zhu, Antonia S. J. S. Mey, Amr Mostafa, Ilko Bald and Michael G. Weller
Viruses 2023, 15(3), 697; https://doi.org/10.3390/v15030697 - 7 Mar 2023
Cited by 2 | Viewed by 3157
Abstract
The cowpea chlorotic mottle virus (CCMV) is a plant virus explored as a nanotechnological platform. The robust self-assembly mechanism of its capsid protein allows for drug encapsulation and targeted delivery. Additionally, the capsid nanoparticle can be used as a programmable platform to display [...] Read more.
The cowpea chlorotic mottle virus (CCMV) is a plant virus explored as a nanotechnological platform. The robust self-assembly mechanism of its capsid protein allows for drug encapsulation and targeted delivery. Additionally, the capsid nanoparticle can be used as a programmable platform to display different molecular moieties. In view of future applications, efficient production and purification of plant viruses are key steps. In established protocols, the need for ultracentrifugation is a significant limitation due to cost, difficult scalability, and safety issues. In addition, the purity of the final virus isolate often remains unclear. Here, an advanced protocol for the purification of the CCMV from infected plant tissue was developed, focusing on efficiency, economy, and final purity. The protocol involves precipitation with PEG 8000, followed by affinity extraction using a novel peptide aptamer. The efficiency of the protocol was validated using size exclusion chromatography, MALDI-TOF mass spectrometry, reversed-phase HPLC, and sandwich immunoassay. Furthermore, it was demonstrated that the final eluate of the affinity column is of exceptional purity (98.4%) determined by HPLC and detection at 220 nm. The scale-up of our proposed method seems to be straightforward, which opens the way to the large-scale production of such nanomaterials. This highly improved protocol may facilitate the use and implementation of plant viruses as nanotechnological platforms for in vitro and in vivo applications. Full article
(This article belongs to the Special Issue Applications of Plant Virus in Biotechnology)
Show Figures

Figure 1

Figure 1
<p>3D model of the capsid of CCMV (PDB ID: 1ZA7) generated with X-ray crystallography data from the RCSB protein data bank [<a href="#B31-viruses-15-00697" class="html-bibr">31</a>], created with Mol* Viewer [<a href="#B22-viruses-15-00697" class="html-bibr">22</a>]. The 180 subunits of the virus consist of chemically identical proteins with three types of different symmetries, highlighted in green, red, and purple, respectively. Please note that the model is created using high-resolution X-ray crystallography data of a K42R CCMV mutant.</p>
Full article ">Figure 2
<p>Simplified workflow for the purification of CCMV using a novel peptide aptamer.</p>
Full article ">Figure 3
<p><b>Left</b>: A small image detail of the combinatoric peptide library immobilized on a glass slide after incubation with fluorescently labeled CCMV in a diluted extract of <span class="html-italic">Vigna unguiculata</span>. In total, approximately 25,000 different peptides were screened on one chip. The positive bead is highlighted with a white circle. <b>Right</b>: MALDI-TOF mass spectrum of the peptide after cleavage from the bead. The ladder sequence introduced during peptide synthesis allows for the simple readout of the peptide sequence without the need for fragmentation.</p>
Full article ">Figure 4
<p><b>Left</b>: Representative structure of the equilibrium-folded structure of the peptide; the glycine–serine linker is shown in purple, and the peptide aptamer in green, with the silane linker shown in licorice representation. <b>Right</b>: Cartoon representation of the peptide for comparison using the same color scheme.</p>
Full article ">Figure 5
<p>Scheme for preparing the affinity column for CCMV purification: Surface functionalization of a sintered glass monolith with epoxy silane and coupling of the thiol-containing CCMV-binding peptide aptamer via ring-opening of the epoxide. Please note: In this figure, the <span class="html-italic">N</span>-terminus of the peptide NH<sub>2</sub>-Tyr-Ile-Gln-Ile-Tyr-Phe-Gly-Tyr-Gly-Gly-Ser-Gly-Gly-Ser-Cys-NH<sub>2</sub> is located on the right-hand side, and the <span class="html-italic">C</span>-terminus is amidated. Due to the specific synthesis and subsequent screening procedure, the peptide must be oriented so that the C-terminus is directed toward the solid phase.</p>
Full article ">Figure 6
<p>Column capacity for the binding of CCMV. Three 10 mL injections of a resuspended PEG-precipitate solution containing 0.1 mg/mL CCMV were made in succession. The flowthrough of the 1st and 2nd injections is free of any CCMV as determined with ELISA. It contains impurities from the plant matrix and PEG 8000. After the third injection, the column capacity is exceeded, and flowthrough peak is increased, as indicated by the dashed line. The column capacity was determined as approximately 2 mg of CCMV. After washing with binding buffer and elution at pH 3.6, pure CCMV was obtained as a narrow peak. The eluting CCMV is immediately neutralized to approximately 50 mM sodium acetate pH 4.8 with a small volume of concentrated neutralization buffer present in the fraction collector.</p>
Full article ">Figure 7
<p>Optimized workflow for the purification of CCMV from leaves of <span class="html-italic">Vigna unguiculata</span>: Samples ① to ⑤ are taken from the different purification steps. ① Leaves are homogenized to a crude extract. The extract is centrifuged, and 10% PEG 8000 is added. After incubation overnight, the solution is centrifuged, and the supernatant ② is removed. The pellet containing CCMV is completely resuspended with buffer. Afterward, the suspension is centrifuged again to obtain the essentially CCMV-free pellet. The supernatant ③ containing the CCMV, among other proteins and PEG 8000, is filtered. Pure CCMV ⑤ is isolated from the supernatant by affinity extraction with the CCMV binding peptide immobilized on a sintered glass monolith surface; the flowthrough of column ④ is discarded.</p>
Full article ">Figure 8
<p>Silver-stained SDS-PAGE analysis of each purification step and the final eluate. Lane 1: Protein ladder; Lane 2: Crude extract; Lane 3: Supernatant after precipitation with PEG 8000; Lane 4: Supernatant of resuspended pellet; Lane 5: Flowthrough of affinity column; Lane 6: Binding buffer as blank; Lane 7: CCMV eluate obtained by affinity purification (final product, diluted to approx. 20 µg/mL).</p>
Full article ">Figure 9
<p>Size exclusion chromatograms (SEC) of samples taken at different steps of the purification protocol are shown in <a href="#viruses-15-00697-f007" class="html-fig">Figure 7</a>: ① crude extract; ③ filtered supernatant after pelleting with PEG 8000; ④ flowthrough of the affinity extraction; ⑤ eluate of the affinity extraction. SEC was performed with a Hi Prep 26/60 Sephacryl S-300-HR, Cytiva (17119601); the running buffer was 0.05 M sodium acetate buffer, 0.15 M NaCl and 1 mM Na<sub>2</sub>EDTA, pH 4.8, with a flow rate of 0.5 mL/min.</p>
Full article ">Figure 10
<p>MALDI-TOF mass spectra of samples taken at different steps of the purification protocol are shown in <a href="#viruses-15-00697-f007" class="html-fig">Figure 7</a>: ① Crude extract; ③ filtered supernatant after pelleting with PEG 8000; ④ flowthrough of the affinity extraction; ⑤ eluate of the affinity extraction.</p>
Full article ">Figure 11
<p>Reversed-phase high-performance liquid chromatography (RP-HPLC) of CCMV after purification by affinity chromatography. 9 µg CCMV was injected and subsequently separated using the following gradients: 0–3 min 99% A (H<sub>2</sub>O with 0.2% TFA) and 1% B (ACN with 0.16% TFA); 3–23 min 30% A and 70% B; 23–40 min 1% A and 99% B, at a constant flow rate of 0.8 mL/min.</p>
Full article ">Figure 12
<p>Representation of CCMV nanoparticles after affinity extraction. <b>Top</b>: Negative staining TEM image showing virions with a diameter of 28 nm. <b>Bottom</b>: AFM image.</p>
Full article ">Figure 13
<p>Calibration curve for the determination of CCMV with a sandwich enzyme-linked immunosorbent assay (ELISA). Error bars correspond to the standard deviation of quadruplicates. LOD and LOQ were calculated to be 0.25 µg/L, and 0.79 µg/L, respectively. No matrix effect was observed by diluting the samples to the working range of this assay (3–200 µg/L).</p>
Full article ">
28 pages, 4094 KiB  
Article
SIV Infection Regulates Compartmentalization of Circulating Blood Plasma miRNAs within Extracellular Vesicles (EVs) and Extracellular Condensates (ECs) and Decreases EV-Associated miRNA-128
by Steven Kopcho, Marina McDew-White, Wasifa Naushad, Mahesh Mohan and Chioma M. Okeoma
Viruses 2023, 15(3), 622; https://doi.org/10.3390/v15030622 - 24 Feb 2023
Cited by 2 | Viewed by 3219
Abstract
Background: This is Manuscript 1 of a two-part Manuscript of the same series. Here, we present findings from our first set of studies on the abundance and compartmentalization of blood plasma extracellular microRNAs (exmiRNAs) into extracellular particles, including blood plasma extracellular vesicles [...] Read more.
Background: This is Manuscript 1 of a two-part Manuscript of the same series. Here, we present findings from our first set of studies on the abundance and compartmentalization of blood plasma extracellular microRNAs (exmiRNAs) into extracellular particles, including blood plasma extracellular vesicles (EVs) and extracellular condensates (ECs) in the setting of untreated HIV/SIV infection. The goals of the study presented in this Manuscript 1 are to (i) assess the abundance and compartmentalization of exmiRNAs in EVs versus ECs in the healthy uninfected state, and (ii) evaluate how SIV infection may affect exmiRNA abundance and compartmentalization in these particles. Considerable effort has been devoted to studying the epigenetic control of viral infection, particularly in understanding the role of exmiRNAs as key regulators of viral pathogenesis. MicroRNA (miRNAs) are small (~20–22 nts) non-coding RNAs that regulate cellular processes through targeted mRNA degradation and/or repression of protein translation. Originally associated with the cellular microenvironment, circulating miRNAs are now known to be present in various extracellular environments, including blood serum and plasma. While in circulation, miRNAs are protected from degradation by ribonucleases through their association with lipid and protein carriers, such as lipoproteins and other extracellular particles—EVs and ECs. Functionally, miRNAs play important roles in diverse biological processes and diseases (cell proliferation, differentiation, apoptosis, stress responses, inflammation, cardiovascular diseases, cancer, aging, neurological diseases, and HIV/SIV pathogenesis). While lipoproteins and EV-associated exmiRNAs have been characterized and linked to various disease processes, the association of exmiRNAs with ECs is yet to be made. Likewise, the effect of SIV infection on the abundance and compartmentalization of exmiRNAs within extracellular particles is unclear. Literature in the EV field has suggested that most circulating miRNAs may not be associated with EVs. However, a systematic analysis of the carriers of exmiRNAs has not been conducted due to the inefficient separation of EVs from other extracellular particles, including ECs. Methods: Paired EVs and ECs were separated from EDTA blood plasma of SIV-uninfected male Indian rhesus macaques (RMs, n = 15). Additionally, paired EVs and ECs were isolated from EDTA blood plasma of combination anti-retroviral therapy (cART) naïve SIV-infected (SIV+, n = 3) RMs at two time points (1- and 5-months post infection, 1 MPI and 5 MPI). Separation of EVs and ECs was achieved with PPLC, a state-of-the-art, innovative technology equipped with gradient agarose bead sizes and a fast fraction collector that allows high-resolution separation and retrieval of preparative quantities of sub-populations of extracellular particles. Global miRNA profiles of the paired EVs and ECs were determined with RealSeq Biosciences (Santa Cruz, CA) custom sequencing platform by conducting small RNA (sRNA)-seq. The sRNA-seq data were analyzed using various bioinformatic tools. Validation of key exmiRNAs was performed using specific TaqMan microRNA stem-loop RT-qPCR assays. Results: We showed that exmiRNAs in blood plasma are not restricted to any type of extracellular particles but are associated with lipid-based carriers—EVs and non-lipid-based carriers—ECs, with a significant (~30%) proportion of the exmiRNAs being associated with ECs. In the blood plasma of uninfected RMs, a total of 315 miRNAs were associated with EVs, while 410 miRNAs were associated with ECs. A comparison of detectable miRNAs within paired EVs and ECs revealed 19 and 114 common miRNAs, respectively, detected in all 15 RMs. Let-7a-5p, Let-7c-5p, miR-26a-5p, miR-191-5p, and let-7f-5p were among the top 5 detectable miRNAs associated with EVs in that order. In ECs, miR-16-5p, miR-451, miR-191-5p, miR-27a-3p, and miR-27b-3p, in that order, were the top detectable miRNAs in ECs. miRNA-target enrichment analysis of the top 10 detected common EV and EC miRNAs identified MYC and TNPO1 as top target genes, respectively. Functional enrichment analysis of top EV- and EC-associated miRNAs identified common and distinct gene-network signatures associated with various biological and disease processes. Top EV-associated miRNAs were implicated in cytokine–cytokine receptor interactions, Th17 cell differentiation, IL-17 signaling, inflammatory bowel disease, and glioma. On the other hand, top EC-associated miRNAs were implicated in lipid and atherosclerosis, Th1 and Th2 cell differentiation, Th17 cell differentiation, and glioma. Interestingly, infection of RMs with SIV revealed that the brain-enriched miR-128-3p was longitudinally and significantly downregulated in EVs, but not ECs. This SIV-mediated decrease in miR-128-3p counts was validated by specific TaqMan microRNA stem-loop RT-qPCR assay. Remarkably, the observed SIV-mediated decrease in miR-128-3p levels in EVs from RMs agrees with publicly available EV miRNAome data by Kaddour et al., 2021, which showed that miR-128-3p levels were significantly lower in semen-derived EVs from HIV-infected men who used or did not use cocaine compared to HIV-uninfected individuals. These findings confirmed our previously reported finding and suggested that miR-128 may be a target of HIV/SIV. Conclusions: In the present study, we used sRNA sequencing to provide a holistic understanding of the repertoire of circulating exmiRNAs and their association with extracellular particles, such as EVs and ECs. Our data also showed that SIV infection altered the profile of the miRNAome of EVs and revealed that miR-128-3p may be a potential target of HIV/SIV. The significant decrease in miR-128-3p in HIV-infected humans and in SIV-infected RMs may indicate disease progression. Our study has important implications for the development of biomarker approaches for various types of cancer, cardiovascular diseases, organ injury, and HIV based on the capture and analysis of circulating exmiRNAs. Full article
(This article belongs to the Special Issue Viruses and Extracellular Vesicles 2023)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Study workflow, EV and EC isolation and characterization. (<b>A</b>) Description of experimental model; 15 male Indian Rhesus Macaques were randomly assigned to 5 groups of 3. Pre-infection and pre-treatment blood plasma samples were collected and processed. (<b>B</b>) Methodological workflow for isolation of EVs and ECs and their characterization. (<b>C</b>) Representative PPLC spectra of EVs and ECs. Blue box: indicates EV-containing fraction. Green box: indicates EC-containing fraction. (<b>D</b>) Representative negative-stain TEM images of purified EVs and ECs from pooled (<span class="html-italic">n</span> = 15) RMs. Blue arrows indicate gold-labeled CD9 on the surface of EVs. Green arrows indicate ECs. Scale bars: 200 nm for EVs and ECs (Top), 50 nm EVs (bottom), and 100 nm ECs (bottom). (<b>E</b>–<b>G</b>) Nanoparticle tracking analysis (NTA) measurements of different BEV properties, including (<b>E</b>) mean EV size, (<b>F</b>) mean EV concentration, (<b>G</b>) mean EV zeta-potential.</p>
Full article ">Figure 2
<p>Identification of common BEV and BEC miRNAs. (<b>A</b>) Number of miRNAs detected (miRNA distribution count ≥1) for each RM (<span class="html-italic">n</span> = 15), for both EVs and ECs. (<b>B</b>) Venn diagram comparing total detectable miRNAs for EVs and ECs (<span class="html-italic">n</span> = 15). To be included in the list, miRNA count needed to be ≥1 at least 1 RM. (<b>C</b>,<b>D</b>) Venn diagram showing common and unique miRNAs among the 5 groups for (<b>C</b>) EVs and (<b>D</b>) ECs. Dotted red circle indicates miRNAs detected in monkeys (<span class="html-italic">n</span> = 15) for EVs (19) and ECs (114). (<b>E</b>,<b>F</b>) Top 10 detected commonly expressed miRNAs as measured by miRNA distribution counts for (<b>E</b>) EVs and (<b>F</b>) ECs. Unpaired T-test with Welch’s correction was used to assess statistical differences between EVs and ECs in panel (<b>A</b>). Error bars represent S.E.M. ****, <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 3
<p>The top 10 miRNAs identified in EVs and ECs regulate distinctive pathways. (<b>A</b>,<b>B</b>) miRNA-target enrichment analysis showing top target genes by number of interactions for A) EV-associated miRNAs and (<b>B</b>) EC-associated miRNAs. The color of the bars represents adjusted <span class="html-italic">p</span>-values (FDR). (<b>C</b>,<b>D</b>) Visualization of miRNA-target interaction network for (<b>C</b>) EV-associated miRNAs and (<b>D</b>) EC-associated miRNAs. Blue circles indicate miRNAs, yellow circles indicate their target genes. (<b>E</b>,<b>F</b>) Dot plot of functional enrichment analysis for target genes of top 10 miRNAs resulting from miRNA-target enrichment analysis for (<b>E</b>) EV-associated miRNAs and (<b>F</b>) EC-associated miRNAs. Color of dots represents adjusted <span class="html-italic">p</span>-values (FDR), and size of dots represents gene ratio (number of miRNA targets found enriched in each category/number of total genes associated with that category). (<b>G</b>) Venn diagram comparing differences and similarities in KEGG pathways of EV- and EC-associated miRNAs.</p>
Full article ">Figure 4
<p>Identification and pathway analysis of common and unique miRNAs associated with EVs and ECs. (<b>A</b>) Venn diagram showing common and unique miRNAs among the common EV and EC miRNAs (<span class="html-italic">n</span> = 15). (<b>B</b>) miRNA distribution counts of EV-associated unique miRNAs (1) for <span class="html-italic">n</span> = 15 RMs. (<b>C</b>) miRNA distribution counts of top 10 EC-associated miRNAs. (<b>D</b>) miRNA-target enrichment analysis showing top target genes by number of interactions for the 1 unique EV-associated miRNA. (<b>E</b>) Visualization of miRNA-target interaction network for the 1 unique EV-associated miRNA. (<b>F</b>) miRNA-target enrichment analysis showing top target genes by number of interactions for the top 10 unique EC-associated miRNAs. (<b>G</b>) Visualization of miRNA-target interaction network for the top 10 unique EC-associated miRNAs. (<b>H</b>) Dot plot of functional enrichment analysis for the top 10 unique EC-associated miRNAs. Color of dots represents adjusted <span class="html-italic">p</span>-values (FDR), and size of dots represents gene ratio (number of miRNA targets found enriched in each category/number of total genes associated with that category). (<b>I</b>) PCA plot of the 18 (arrow from panel (<b>A</b>)) common EV and EC miRNAs. Unit variance scaling is applied to rows; SVD with imputation is used to calculate principal components. X and Y axis show principal component 1 and principal component 2, which explain 74.4% and 19.1% of the total variance, respectively. Predication ellipses are such that with a probability of 0.95, a new observation from the same group will fall inside the ellipse. <span class="html-italic">N</span> = 15 data points. (<b>J</b>) Hierarchical clustering heatmap of the 18 common EV and EC miRNAs. Rows are centered; unit variance scaling is applied to rows. Both rows and columns are clustered using correlation distance and average linkage. (<b>K</b>) miRNA-target enrichment analysis showing top target genes by number of interactions for the 18 common EV- and EC-associated miRNAs. (<b>L</b>) Visualization of miRNA-target interaction network for 18 common EV- and EC-associated miRNAs. (<b>M</b>) Dot plot of functional enrichment analysis for target genes of 18 common EV- and EC-associated miRNAs. Color of dots represents adjusted <span class="html-italic">p</span>-values (FDR), and size of dots represents gene ratio (number of miRNA targets found enriched in each category/number of total genes associated with that category.</p>
Full article ">Figure 5
<p>SIV infection of RMs longitudinally downregulates EV-associated miR-128-3p. (<b>A</b>) Schematic of SIV infection of RMs; 12 male Indian RMs were infected with SIV. One month post-infection (1 MPI), blood plasma was collected from <span class="html-italic">n</span> = 12 RMS. Five months post-infection (5 MPI), blood plasma was collected from <span class="html-italic">n</span> = 3 RMS. (<b>B</b>) Number of miRNAs detected (miRNA distribution count ≥ 1) for each RM, for both EVs and ECs. Pre (<span class="html-italic">n</span> = 15), SIV 1 MPI (<span class="html-italic">n</span> = 12), SIV 5 MPI (<span class="html-italic">n</span> = 3). (<b>C</b>–<b>F</b>) Volcano plots showing down-regulated (blue) and up-regulated (red) miRNAs in (<b>C</b>) EVs 1 MPI, (<b>D</b>) ECs 1 MPI, (<b>E</b>) EVs 5 MPI, and (<b>F</b>) BCs 5 MPI compared to healthy uninfected RMs (Pre). (<b>G</b>–<b>I</b>) miRNA-target enrichment analysis (<b>G</b>), visualization of miRNA-target interaction network (<b>H</b>), and dot plot of functional enrichment analysis (<b>I</b>) for the longitudinally downregulated EV-associated miRNAs (miR-206, miR-99a-5p, miR-128-3p). Color of dots in panel (<b>I</b>) represents adjusted <span class="html-italic">p</span>-values (FDR), and size of dots represents gene ratio (number of miRNA targets found enriched in each category/number of total genes associated with that category. (<b>J</b>) TaqMan PCR validation using 128a-3p specific assays. Statistical differences were assessed by ordinary one-way ANOVA test with Tukey’s correction (<span class="html-italic">n</span> = 3). *, <span class="html-italic">p</span> &lt; 0.05. (<b>K</b>) miRNA-target enrichment analysis showing top target genes by number of interactions for miR-128-3p. (<b>L</b>) Visualization of miRNA-target interaction network for miR-128-3p. (<b>M</b>,<b>N</b>) Dot plots of functional enrichment analysis (<b>M</b>) KEGG and (<b>N</b>) disease Ontology for target genes of miR-128-3p. Color of dots represents adjusted <span class="html-italic">p</span>-values (FDR), and size of dots represents gene ratio (number of miRNA targets found enriched in each category/number of total genes associated with that category). Unpaired T-test with Welch’s correction was used to assess statistical differences between EVs and ECs in panels (<b>B</b>) and (<b>J</b>) (left). Error bars represent S.E.M. *, <span class="html-italic">p</span> &lt; 0.05; ****, <span class="html-italic">p</span> &lt; 0.0001; ns, not significant. In Panel J, Ordinary One-way ANOVA multiple comparison test (Tukey’s test) was used to assess statistical differences, with ns denoting non-significant.</p>
Full article ">Figure 6
<p>Circulating blood plasma miRNAs and their association with EVs and ECs in uninfected and SIV-infected rhesus macaques. Part of this illustration was created with BioRender.com.</p>
Full article ">
21 pages, 2488 KiB  
Article
A Capsid Protein Fragment of a Fusagra-like Virus Found in Carica papaya Latex Interacts with the 50S Ribosomal Protein L17
by Marlonni Maurastoni, Tathiana F. Sá Antunes, Emanuel F. M. Abreu, Simone G. Ribeiro, Angela Mehta, Marcio M. Sanches, Wagner Fontes, Elliot W. Kitajima, Fabiano T. Cruz, Alexandre M. C. Santos, Jose A. Ventura, Ana C. M. M. Gomes, F. Murilo Zerbini, Patricia Sosa-Acosta, Fábio C. S. Nogueira, Silas P. Rodrigues, Francisco J. L. Aragão, Anna E. Whitfield and Patricia M. B. Fernandes
Viruses 2023, 15(2), 541; https://doi.org/10.3390/v15020541 - 15 Feb 2023
Cited by 2 | Viewed by 2434
Abstract
Papaya sticky disease is caused by the association of a fusagra-like and an umbra-like virus, named papaya meleira virus (PMeV) and papaya meleira virus 2 (PMeV2), respectively. Both viral genomes are encapsidated in particles formed by the PMeV ORF1 product, which has the [...] Read more.
Papaya sticky disease is caused by the association of a fusagra-like and an umbra-like virus, named papaya meleira virus (PMeV) and papaya meleira virus 2 (PMeV2), respectively. Both viral genomes are encapsidated in particles formed by the PMeV ORF1 product, which has the potential to encode a protein with 1563 amino acids (aa). However, the structural components of the viral capsid are unknown. To characterize the structural proteins of PMeV and PMeV2, virions were purified from Carica papaya latex. SDS-PAGE analysis of purified virus revealed two major proteins of ~40 kDa and ~55 kDa. Amino-terminal sequencing of the ~55 kDa protein and LC-MS/MS of purified virions indicated that this protein starts at aa 263 of the deduced ORF1 product as a result of either degradation or proteolytic processing. A yeast two-hybrid assay was used to identify Arabidopsis proteins interacting with two PMeV ORF1 product fragments (aa 321–670 and 961–1200). The 50S ribosomal protein L17 (AtRPL17) was identified as potentially associated with modulated translation-related proteins. In plant cells, AtRPL17 co-localized and interacted with the PMeV ORF1 fragments. These findings support the hypothesis that the interaction between PMeV/PMeV2 structural proteins and RPL17 is important for virus–host interactions. Full article
(This article belongs to the Section Viruses of Plants, Fungi and Protozoa)
Show Figures

Figure 1

Figure 1
<p>Characterization of the papaya meleira virus (PMeV) complex capsid protein composition. (<b>A</b>) The three opalescent fractions (T- Top, M- middle, and B- Bottom fraction) were obtained after ultracentrifugation at 145,000× <span class="html-italic">g</span> for 18 h at 4 °C in a 50% (<span class="html-italic">w/v</span>) cesium chloride isopycnic gradient. (<b>B</b>) Transmission electron microscopy images of viral particles from the M and B fractions. Viral particles are approximately 50 nm in diameter and are circled in the figure. (<b>C</b>) Coomassie blue-stained SDS-PAGE of fractions, lanes B and M are fractions from ultracentrifugation and L is the protein ladder. (<b>D</b>) Mapping positions of peptides identified in cesium chloride-purified virions collected from B and M fractions matching with the PMeV deduced ORF1 protein (GenBank accession OP834191) (light-blue rectangles) and protein fragments used for binary interactions by yeast two-hybrid (Y2H) and structural proteins. The black rectangle represents the N-terminal sequencing of the ~55 kDa band obtained after the separation of proteins in the virion preparation; the orange rectangle indicates an ~15 kDa gap with no peptides identified. The yellow rectangle represents a degraded or cleaved ~30 kDa N-terminal protein. The green rectangle represents p55. The light-yellow rectangle represents the putative coding region of uncharacterized structural protein(s).</p>
Full article ">Figure 2
<p>Summary of yeast two-hybrid assays mapping dimerization fragments in the papaya meleira virus (PMeV) ORF1 product, and detection of five fragments of the PMeV ORF1 product. (<b>A</b>) Each ORF1 fragment was fused either to the GAL4 binding domain (BD) or GAL4 activation domain (AD) in pDESTGBKT7 and pDESTGADT7, respectively, and transformed in the yeast strain Y2HGold. Positive interactors were selected in QDO/X/A media. (<b>B</b>) The c-myc-fused protein expression was verified by SDS-PAGE of yeast crude protein extracts and Western blotting using an anti-c-myc antibody. The 1- to 5-BD represents each fragment of the PMeV ORF1 product fused to GAL4 BD. 53-BD (GAL4 DNA-BD fused with murine p53) and Lam-BD (GAL4 BD fused with Lamin) were used as controls. Untransformed yeast and yeast transformed with pDEST-BGKT7 plasmids were used as the negative control. M: PageRuler Plus prestained protein ladder.</p>
Full article ">Figure 3
<p>Protein–protein interaction network (PPI) of <span class="html-italic">Carica papaya</span> differentially expressed proteins during PMeV complex infection and PMeV CP2 and CP4-interacting proteins. PPI network of PMeV complex-infected <span class="html-italic">C. papaya</span> at (<b>A</b>) pre-flowering stage (4 months post-germination) and (<b>B</b>) post-flowering stage (7 months post-germination). The PPI network was filtered to show proteins involved in translation (see <a href="#app1-viruses-15-00541" class="html-app">Figure S6</a> for a complete network). Red nodes are up-regulated proteins; green nodes are down-regulated proteins; blue nodes are PMeV CP2 and CP4-interacting proteins.</p>
Full article ">Figure 4
<p>Transient expression and co-localization of <span class="html-italic">At</span>RPL17 and CP2 in <span class="html-italic">Nicotiana benthamiana</span>. (<b>A</b>) Localization of green fluorescent protein (GFP)-fused <span class="html-italic">At</span>RPL17 and CP2 proteins (CP2::GFP, <span class="html-italic">At</span>RPL17::GFP, and GFP::AtRPL17) in leaf epidermal cells of <span class="html-italic">N. benthamiana</span> visualized at 2 days post-infiltration. (<b>B</b>) Co-localization of <span class="html-italic">At</span>RPL17 and CP2 expressed as fusions to (GFP) or red fluorescent protein (RFP) in <span class="html-italic">N. benthamiana</span> epidermal leaf cells. Right column image is GFP and RFP overlayed channels. The fusion proteins RFP::<span class="html-italic">At</span>RPL17 were expressed along with GFP::CP2. White arrows: co-localization signals.</p>
Full article ">Figure 5
<p>Interaction of <span class="html-italic">At</span>RPL17 and CP2 in <span class="html-italic">Nicotiana benthamiana</span>. (<b>A</b>,<b>B</b>) PMeV CP2-<span class="html-italic">At</span>RPL17 interaction in vivo by bimolecular fluorescence complementation (BiFC) in wild type (<b>A</b>) and transgenic <span class="html-italic">N. benthamiana</span> expressing RFP::H2B as a nuclear marker (<b>B</b>). Fusion proteins nYFP::<span class="html-italic">At</span>RPL17 or GST::nYFP were expressed along with CP2::cYFP by agroinfiltration of the encoding plasmids into leaves of <span class="html-italic">N. benthamiana</span>. The reconstitution of yellow fluorescence was visualized 2 days post-infiltration. Fusion protein combinations expressed in each sample are indicated at the left of the corresponding row of images. Inset shows a detailed region of the nuclear marker and the reconstitution of yellow fluorescent protein. White arrows: interaction signals.</p>
Full article ">
16 pages, 3074 KiB  
Article
Associations between NK Cells in Different Immune Organs and Cellular SIV DNA and RNA in Regional HLADR CD4+ T Cells in Chronically SIVmac239-Infected, Treatment-Naïve Rhesus Macaques
by Xinjie Li, Liyan Zhu, Yue Yin, Xueying Fan, Linting Lv, Yuqi Zhang, Yijin Pan, Yangxuanyu Yan, Hua Liang, Jing Xue and Tao Shen
Viruses 2022, 14(11), 2513; https://doi.org/10.3390/v14112513 - 13 Nov 2022
Viewed by 1673
Abstract
With the development of NK cell-directed therapeutic strategies, the actual effect of NK cells on the cellular SIV DNA levels of the virus in SIV-infected macaques in vivo remains unclear. In this study, five chronically SIVmac239-infected, treatment-naïve rhesus macaques were euthanized, [...] Read more.
With the development of NK cell-directed therapeutic strategies, the actual effect of NK cells on the cellular SIV DNA levels of the virus in SIV-infected macaques in vivo remains unclear. In this study, five chronically SIVmac239-infected, treatment-naïve rhesus macaques were euthanized, and the blood, spleen, pararectal/paracolonic lymph nodes (PaLNs), and axillary lymph nodes (ALNs) were collected. The distributional, phenotypic, and functional profiles of NK cells were detected by flow cytometry. The highest frequency of NK cells was found in PBMC, followed by the spleen, while only 0~0.5% were found in LNs. Peripheral NK cells also exhibited higher cytotoxic potential (CD56 CD16+ NK subsets) and IFN-γ-producing capacity but low PD-1 and Tim-3 levels than those in the spleen and LNs. Our results demonstrated a significant positive correlation between the frequency of NK cells and the ratios of cellular SIV DNA/RNA in HLADR CD4+ T cells (r = 0.6806, p < 0.001) in SIV-infected macaques, despite no discrepancies in the cellular SIV DNA or RNA levels that were found among the blood, spleen, and LNs. These findings showed a profile of NK cell frequencies and NK cytotoxicity levels in different immune organs from chronically SIVmac239-infected, treatment-naïve rhesus macaques. It was suggested that NK cell frequencies could be closely related to SIV DNA/RNA levels, which could affect the transcriptional activity of SIV proviruses. However, the cytotoxicity effect of NK cells on the latent SIV viral load in LNs could be limited due to the sparse abundance of NK cells in LNs. The development of NK cell-directed treatment approaches aiming for HIV clearance remains challenging. Full article
(This article belongs to the Special Issue Viral-Host Cell Interactions of Animal Viruses)
Show Figures

Figure 1

Figure 1
<p>The distribution of simian NK cells and their subsets in peripheral blood and secondary lymphoid organs in SIV-infected macaques. (<b>A</b>) Flow cytometric gating defined NK cell subsets (CD3<sup>−</sup> CD8α<sup>+</sup> NKG2A<sup>+</sup>) in PBMCs from healthy macaques (<span class="html-italic">n</span> = 5) and in single cells of peripheral blood, PaLNs, ALNs, and spleen from SIV-infected macaques (<span class="html-italic">n</span> = 5). This illustration shows a representative plot of NK cell subsets form a healthy macaque and an infected macaque. (<b>B</b>) The frequencies of NK cells in total lymphocytes from different tissues. (<b>C</b>) The distributional characteristics of three NK subsets (CD56<sup>−</sup> CD16<sup>+</sup>, CD56<sup>+</sup> CD16<sup>−</sup>, and DN, CD56<sup>−</sup> CD16<sup>−</sup>). Lines represent median values; little circles in different colors mean five SIV-infected macaques’ percentage levels in different groups. As the data did not obey a normal distribution, after comparing multiple group differences using the Kruskal–Wallis test, the Mann–Whitney U test was performed to compare differences between the two groups. “#” indicates significant differences using the Kruskal–Wallis test (#, <span class="html-italic">p</span> &lt; 0.05; ###, <span class="html-italic">p</span> &lt; 0.001). “*” indicates significant differences by Mann–Whitney U test (*, <span class="html-italic">p</span> &lt; 0.05; **, <span class="html-italic">p</span> &lt; 0.01). PBMC—peripheral blood mononuclear cell; PaLN—paracolonic lymph node; ALN—axillary lymph node; DN—double negative.</p>
Full article ">Figure 2
<p>The ratio of NK cells to CD4<sup>+</sup> T cells and HLADR<sup>−</sup> CD4<sup>+</sup> T cells of SIV-infected macaques. (<b>A</b>) Levels of CD4<sup>+</sup> T cells and HLADR<sup>−</sup> CD4<sup>+</sup> T cells in single-cell suspension. (<b>B</b>) Ratio of NK cells to CD4<sup>+</sup> T cells and HLADR<sup>−</sup> CD4<sup>+</sup> T cells. Lines represent median values; little circles in different colors mean five SIV-infected macaques’ percentage levels in different groups. As the data did not obey a normal distribution, after comparing multiple group differences using the Kruskal–Wallis test, the Mann–Whitney U test was performed to compare differences between the two groups. “#” indicates significant differences using the Kruskal–Wallis test (#, <span class="html-italic">p</span> &lt; 0.05; ###, <span class="html-italic">p</span> &lt; 0.001). “*” indicates significant differences by Mann–Whitney U test (*, <span class="html-italic">p</span> &lt; 0.05; **, <span class="html-italic">p</span> &lt; 0.01). PBMC—peripheral blood mononuclear cell; PaLN—paracolonic lymph node; ALN—axillary lymph node.</p>
Full article ">Figure 3
<p>Expression of PD-1, Tim-3, and NK cells receptors on simian NK cells from healthy and SIV-infected macaques. (<b>A</b>) and (<b>C</b>) are representative cytofluorometric plots for one healthy macaque (a) and SIV-infected macaques (b, c, d, e). Representation of groups a, b, c, d, and e were on the top right corner. Group a was PBMC form healthy macaques; Group b, c, d, e, respectively represented PBMC, PaLN, ALN, spleen from SIV-infected macaques. (<b>A</b>) Representative cytofluorometric analysis of PD-1 and Tim-3 expression on simian NK cells. NK cells were gated according to the lymphocyte forward and side scatter pattern and then CD3<sup>−</sup> CD8α<sup>+</sup> NKG2A<sup>+</sup> cells were gated to analyze PD-1 and Tim-3 expression. (<b>B</b>) Box-plot analysis for the frequencies of PD-1<sup>+</sup> and Tim-3<sup>+</sup>NK cells. Lines represent median values, boxes show the 25th and 75th percentiles, and bars show minimum and maximum values. As the data did not obey a normal distribution, after comparing multiple group differences using the Kruskal–Wallis test, the Mann–Whitney U test was performed to compare differences between the two groups. “#” indicates significant differences using the Kruskal–Wallis test (#, <span class="html-italic">p</span> &lt; 0.05; ##, <span class="html-italic">p</span> &lt; 0.01). “*” indicates significant differences by Mann–Whitney U test (*, <span class="html-italic">p</span> &lt; 0.05; **, <span class="html-italic">p</span> &lt; 0.01). (<b>C</b>) Representative flow cytometric plots indicating expressions of NKG2D, NKp44 and NKp46 on simian NK cells. NK cells were gated according to the lymphocyte forward and side scatter pattern and then CD3<sup>−</sup> CD8α<sup>+</sup> NKG2A<sup>+</sup> cells were gated for analysis. (<b>D</b>) Box-plot analysis for the frequencies of NKG2D<sup>+</sup>, NKp44<sup>+</sup>, and NKp46<sup>+</sup> NK cells. Lines represent median values, boxes show the 25th and 75th percentiles, and bars show minimum and maximum values. As the data did not obey a normal distribution, after comparing multiple group differences using the Kruskal–Wallis test, the Mann–Whitney U test was performed to compare differences between the two groups. “#” indicates significant differences using the Kruskal–Wallis test (#, <span class="html-italic">p</span> &lt; 0.05; ##, <span class="html-italic">p</span> &lt; 0.01). “*” indicates significant differences by Mann–Whitney U test (*, <span class="html-italic">p</span> &lt; 0.05; **, <span class="html-italic">p</span> &lt; 0.01). PBMC—peripheral blood mononuclear cell; PaLN—paracolonic lymph node; ALN—axillary lymph node.</p>
Full article ">Figure 4
<p>Functional activity of NK cells from simian PBMCs and splenic single-cell suspension. (<b>A</b>) Lymphocytes were isolated from the peripheral blood (healthy and SIV-infected macaques) and the splenic single-cell suspension (SIV-infected macaques only) and were stimulated with PMA and ionomycin, K562 cells, and CD16 cross-linking. IFN-γ production and CD107a expression by NK cells were shown. (<b>B</b>) The frequencies of IFN-γ<sup>+</sup> and CD107a<sup>+</sup> NK cells among total NK cells. “*” indicates significant differences by Mann–Whitney U test (*, <span class="html-italic">p</span> &lt; 0.05; **, <span class="html-italic">p</span> &lt; 0.01). PBMC—peripheral blood mononuclear cell; PaLN—paracolonic lymph node; ALN—axillary lymph node.</p>
Full article ">Figure 5
<p>Correlation analysis between frequencies of NK cells in lymphocytes, percentages of cytotoxic NK subsets, phenotypic indicators of NK cells, and cellular SIV DNA/RNA ratios in CD3<sup>+</sup> CD4<sup>+</sup> CD8<sup>−</sup> HLADR<sup>−</sup> T cells. The correlation was performed using the Spearman rank correlation coefficient test. <span class="html-italic">p</span> values &lt; 0.05 were considered a statistical difference (***, <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 6
<p>The diagram of NK cells and HLADR<sup>−</sup> CD4<sup>+</sup> T cells in the blood, LNs, and spleen.</p>
Full article ">
15 pages, 2583 KiB  
Article
IL-33 Induces an Antiviral Signature in Mast Cells but Enhances Their Permissiveness for Human Rhinovirus Infection
by Charlene Akoto, Anna Willis, Chiara F. Banas, Joseph A. Bell, Dean Bryant, Cornelia Blume, Donna E. Davies and Emily J. Swindle
Viruses 2022, 14(11), 2430; https://doi.org/10.3390/v14112430 - 1 Nov 2022
Cited by 3 | Viewed by 2394
Abstract
Mast cells (MCs) are classically associated with allergic asthma but their role in antiviral immunity is unclear. Human rhinoviruses (HRVs) are a major cause of asthma exacerbations and can infect and replicate within MCs. The primary site of HRV infection is the airway [...] Read more.
Mast cells (MCs) are classically associated with allergic asthma but their role in antiviral immunity is unclear. Human rhinoviruses (HRVs) are a major cause of asthma exacerbations and can infect and replicate within MCs. The primary site of HRV infection is the airway epithelium and MCs localise to this site with increasing asthma severity. The asthma susceptibility gene, IL-33, encodes an epithelial-derived cytokine released following HRV infection but its impact on MC antiviral responses has yet to be determined. In this study we investigated the global response of LAD2 MCs to IL-33 stimulation using RNA sequencing and identified genes involved in antiviral immunity. In spite of this, IL-33 treatment increased permissiveness of MCs to HRV16 infection which, from the RNA-Seq data, we attributed to upregulation of ICAM1. Flow cytometric analysis confirmed an IL-33-dependent increase in ICAM1 surface expression as well as LDLR, the receptors used by major and minor group HRVs for cellular entry. Neutralisation of ICAM1 reduced the IL-33-dependent enhancement in HRV16 replication and release in both LAD2 MCs and cord blood derived MCs. These findings demonstrate that although IL-33 induces an antiviral signature in MCs, it also upregulates the receptors for HRV entry to enhance infection. This highlights the potential for a gene-environment interaction involving IL33 and HRV in MCs to contribute to virus-induced asthma exacerbations. Full article
(This article belongs to the Special Issue Rhinovirus Infections 2.0)
Show Figures

Figure 1

Figure 1
<p>IL-33 induced 414 DEGs in MCs. (<b>A</b>)<b>,</b> volcano plot of DEGs 6 h post IL-33 (10 ng/mL) treatment, <span class="html-italic">n</span> = 7. (<b>B</b>), Heatmap of all 414 DEGs 6 h post IL-33 (10 ng/mL) treatment, <span class="html-italic">n</span> = 7. (<b>C</b>), Hierarchal clustering heatmap of top 100 most significant DEGs 6 h post IL-33 (10 ng/mL) treatment, <span class="html-italic">n</span> = 7.</p>
Full article ">Figure 2
<p>IL-33 induced DEGs associated with viral immunity in MCs. (<b>A</b>), gene enrichment map of GO terms analysed using g.profiler. (<b>B</b>), top KEGG pathways 6 h post IL-33 (10 ng/mL) treatment, <span class="html-italic">n</span> = 7. (<b>C</b>), genes associated with the influenza Pathway using KEGG on DEGs 6 h post IL-33 (10 ng/mL) treatment, <span class="html-italic">n</span> = 7. (<b>D</b>), mRNA expression of genes (<span class="html-italic">IFIH1, TNFA</span>) in influenza pathway 6 h post IL-33 stimulation (10 ng/mL), <span class="html-italic">n</span> = 7. ** <span class="html-italic">p</span> ≤ 0.01, *** <span class="html-italic">p</span> ≤ 0.001 for control versus IL-33.</p>
Full article ">Figure 3
<p>IL-33 enhanced HRV16-dependent anti-viral responses in MCs. mRNA expression of IFNs <span class="html-italic">IFNB1</span> (<b>A</b>), <span class="html-italic">IFNL1</span> (<b>B</b>) and IFN-stimulated genes <span class="html-italic">IFIH1</span> (<b>C</b>), <span class="html-italic">OAS1</span> (<b>D</b>) and IFN-β protein (<b>E</b>) in LAD2 MCs pretreated with or without IL-33 (1–10 ng/mL) for 24 h prior to HRV16 or UV-HRV16 (control) infection (MOI 7.5) for a further 24 h, <span class="html-italic">n</span> = 3–6. * <span class="html-italic">p</span> ≤ 0.05, ** <span class="html-italic">p</span> ≤ 0.01 for no cytokine versus IL-33 (1 or 10 ng/mL). <sup>#</sup> <span class="html-italic">p</span> ≤ 0.05, <sup>##</sup> <span class="html-italic">p</span> ≤ 0.05 for UV-HRV16 versus HRV16.</p>
Full article ">Figure 4
<p>IL-33 enhances HRV16 replication and virion release in MCs. Viral RNA (<b>A</b>) and virion release (<b>B</b>) in LAD2 MCs pretreated with or without IL-33 (10 ng/mL) for 24 h prior to HRV16 infection (MOI 7.5) for a further 24 h, <span class="html-italic">n</span> = 15. ** <span class="html-italic">p</span> ≤ 0.01, **** <span class="html-italic">p</span> ≤ 0.0001 for HRV16 versus IL-33+HRV16.</p>
Full article ">Figure 5
<p>Role of ICAM1 in mediating IL-33-dependent enhancement of HRV16 replication in LAD2 MCs. A representative flow cytometric trace (<b>A</b>) and averaged geometric mean (<b>B</b>) for ICAM1 cell surface expression in MCs treated with IL-33 (10 ng/mL) for 24 h, <span class="html-italic">n</span> = 8. HRV16 replication (<b>C</b>) and virion release (<b>D</b>) and IFN-β protein release (<b>E</b>) following IL-33 (10 ng/mL) stimulation for 24 h prior to HRV16 infection (MOI 7.5) in the presence or absence of anti-ICAM1 antibody or IgG2a isotype for a further 24 h, <span class="html-italic">n</span> = 5. ** <span class="html-italic">p</span> ≤ 0.01 for no cytokine versus IL-33 and * <span class="html-italic">p</span> ≤ 0.05 for IL-33 IgG2a+HRV16 versus IL-33 anti-ICAM1+HRV16.</p>
Full article ">Figure 6
<p>Role of ICAM1 in mediating IL-33-dependent enhancement of HRV16 replication in CBMCs. Viral RNA in CBMCs pretreated with or without IL-33 (10 ng/mL) for 24 h prior to HRV16 infection (MOI 7.5) for a further 24 h, <span class="html-italic">n</span> = 5 (<b>A</b>). A representative flow cytometric trace (<b>B</b>) and averaged geometric mean (<b>C</b>) for ICAM1 cell surface expression in MCs treated with IL-33 (10 ng/mL) for 24 h, <span class="html-italic">n</span> = 2. HRV16 replication (<b>D</b>), virion release (<b>E</b>) and IFN-β release (<b>F</b>) following IL-33 (10 ng/mL) stimulation for 24 h prior to HRV16 infection (MOI 7.5) in the presence or absence of anti-ICAM1 antibody or IgG2a isotype for a further 24 h, <span class="html-italic">n</span> = 4. * <span class="html-italic">p</span> ≤ 0.05 for no cytokine versus IL-33 and * <span class="html-italic">p</span> ≤ 0.05, **** <span class="html-italic">p</span> ≤ 0.0001 for IL-33 IgG2a+HRV16 versus IL-33 anti-ICAM1+HRV16.</p>
Full article ">
12 pages, 1480 KiB  
Article
Increased Polymerase Activity of Zoonotic H7N9 Allows Partial Escape from MxA
by Philipp P. Petric, Jacqueline King, Laura Graf, Anne Pohlmann, Martin Beer and Martin Schwemmle
Viruses 2022, 14(11), 2331; https://doi.org/10.3390/v14112331 - 24 Oct 2022
Cited by 3 | Viewed by 2127
Abstract
The interferon-induced myxovirus resistance protein A (MxA) is a potent restriction factor that prevents zoonotic infection from influenza A virus (IAV) subtype H7N9. Individuals expressing antivirally inactive MxA variants are highly susceptible to these infections. However, human-adapted IAVs have acquired specific mutations in [...] Read more.
The interferon-induced myxovirus resistance protein A (MxA) is a potent restriction factor that prevents zoonotic infection from influenza A virus (IAV) subtype H7N9. Individuals expressing antivirally inactive MxA variants are highly susceptible to these infections. However, human-adapted IAVs have acquired specific mutations in the viral nucleoprotein (NP) that allow escape from MxA-mediated restriction but that have not been observed in MxA-sensitive, human H7N9 isolates. To date, it is unknown whether H7N9 can adapt to escape MxA-mediated restriction. To study this, we infected Rag2-knockout (Rag2−/−) mice with a defect in T and B cell maturation carrying a human MxA transgene (MxAtg/−Rag2−/−). In these mice, the virus could replicate for several weeks facilitating host adaptation. In MxAtg/−Rag2−/−, but not in Rag2−/− mice, the well-described mammalian adaptation E627K in the viral polymerase subunit PB2 was acquired, but no variants with MxA escape mutations in NP were detected. Utilizing reverse genetics, we could show that acquisition of PB2 E627K allowed partial evasion from MxA restriction in MxAtg/tg mice. However, pretreatment with type I interferon decreased viral replication in these mice, suggesting that PB2 E627K is not a true MxA escape mutation. Based on these results, we speculate that it might be difficult for H7N9 to acquire MxA escape mutations in the viral NP. This is consistent with previous findings showing that MxA escape mutations cause severe attenuation of IAVs of avian origin. Full article
(This article belongs to the Special Issue Transcription and Replication of the Negative-Strand RNA Viruses)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>HPAIV H7N9 shows high sensitivity to the antiviral activity of human MxA. (<b>A</b>) MDCK-SIAT1 cells (ctrl) and MDCK-SIAT1 cells either expressing antivirally active MxA or inactive MxA-T103A were infected with H7N9 at an MOI of 0.001. Viral titers were determined at the indicated time points via plaque assay. Significance levels indicate differences between MDCK-SIAT1-MxA and -MxA-T103A cells. (<b>B</b>) HEK293T cells were co-transfected with expression plasmids coding for the viral polymerase subunits H7N9-PB2, -PB1 and -PA (10 ng each), H7N9-NP (100 ng), an artificial minigenome encoding firefly luciferase as a reporter under the control of the viral promoter (100 ng), and a plasmid encoding Renilla luciferase (30 ng). Constitutive co-expression of Renilla luciferase allows for normalization of transfection efficiency. In addition, expression plasmids encoding either MxA or MxA-T103A (200 ng) or the empty vector (EV) as a control were co-transfected. Luciferase activities were measured 24 h post-transfection. (<b>C</b>,<b>D</b>) C57BL/6 (B6) and (<b>E</b>,<b>F</b>) MxA<sup>tg/tg</sup> (MxA) mice (n = 5 each) were intranasally infected with the indicated doses of H7N9 in 40 µL PBS. Bodyweight changes (<b>C</b>,<b>E</b>) and survival rates (<b>D</b>,<b>F</b>) were monitored for 14 days. Mice were sacrificed once their weight fell below 75% of the initial body weight. Student‘s <span class="html-italic">t</span> test was performed to determine statistical differences. **, <span class="html-italic">p</span> &lt; 0.01; ***, <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 2
<p>Rag2<sup>−/−</sup> mice support prolonged IAV infection. (<b>A</b>,<b>B</b>) MxA<sup>tg/−</sup>Rag2<sup>−/−</sup> mice and MxA<sup>−/−</sup> mice with and without functional Rag2 (n = 4 each) were infected intranasally with 40 µL of 0.2 × LD<sub>50</sub> depending on their MxA genotype (10<sup>4</sup> PFU for MxA<sup>tg/−</sup> and 4 PFU for MxA<sup>−/−</sup>). Body weight changes (<b>A</b>) and survival rates (<b>B</b>) were monitored for 21 days. Mice were sacrificed once their weight fell below 75% of the initial body weight. At the day of death (or on day 21 for the surviving mice) lungs (<b>C</b>) and snouts (<b>D</b>) were harvested to determine viral titers and to isolate virus for further analysis.</p>
Full article ">Figure 3
<p>Virus isolates from MxA<sup>tg/−</sup>Rag2<sup>−/−</sup> mice show enhanced replication in MxA<sup>tg/tg</sup> mice. (<b>A</b>) MxA<sup>tg/tg</sup> mice were infected intranasally with 10<sup>4</sup> PFU of the viral lung isolates of day 11, day 13, and day 16 in <a href="#viruses-14-02331-f002" class="html-fig">Figure 2</a> or wt H7N9 as a control. The lungs of the infected mice were harvested 3 dpi to determine viral titers. (<b>B</b>) MxA<sup>tg/tg</sup> mice were infected intranasally with 10<sup>4</sup> PFU of the viral isolates with the highest titers in each group from the experiment shown in (<b>A</b>) (indicated with ′). The lungs of the infected mice were harvested 3 dpi to determine viral titers via plaque assay. (<b>C</b>) The twelve isolates from (<b>B</b>) were subjected to full-genome sequencing for frequency analysis of adaptive mutations. All mutations that were found with a frequency of at least 10% are listed. HA mutations are displayed according to H7N7 numbering [<a href="#B23-viruses-14-02331" class="html-bibr">23</a>].</p>
Full article ">Figure 4
<p>The PB2 E627K mutation enhances viral replication of HPAIV H7N9. (<b>A</b>) MDCK-SIAT1 cells (ctrl) and MDCK-SIAT1 cells either expressing antivirally active MxA or inactive MxA-T103A were infected with H7N9 at an MOI of 0.001. Viral titers were determined at the indicated time points via plaque assay. Significance levels indicate differences between MDCK-SIAT1-MxA and -MxA-T103A cells. (<b>B</b>) Viral polymerase reconstitution assay were performed as described in the legend to <a href="#viruses-14-02331-f001" class="html-fig">Figure 1</a>B using a plasmid encoding PB2-E627K instead of wildtype PB2. (<b>C</b>,<b>D</b>) MxA<sup>tg/tg</sup> mice (MxA) were infected intranasally with the indicated doses of H7N9-PB2-E627K in 40 µL PBS. Body weight changes (<b>C</b>) and survival rates (<b>D</b>) were monitored for 14 days. Mice were sacrificed once their weight fell below 75% of the initial body weight. (<b>E</b>) MxA<sup>tg/tg</sup> mice were pretreated with IFN-α (105 units) or PBS 24 h prior to infection with 10<sup>4</sup> PFU of the indicated viruses. The lungs of the infected mice were harvested 3 dpi to determine viral titers. Student‘s <span class="html-italic">t</span> test was performed to determine statistical differences. *, <span class="html-italic">p</span> &lt; 0.05; **, <span class="html-italic">p</span> &lt; 0.01; ***, <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">
15 pages, 2310 KiB  
Article
Cytokine Profiling of Amniotic Fluid from Congenital Cytomegalovirus Infection
by Nicolas Bourgon, Wendy Fitzgerald, Hugues Aschard, Jean-François Magny, Tiffany Guilleminot, Julien Stirnemann, Roberto Romero, Yves Ville, Leonid Margolis and Marianne Leruez-Ville
Viruses 2022, 14(10), 2145; https://doi.org/10.3390/v14102145 - 28 Sep 2022
Cited by 5 | Viewed by 2086
Abstract
Background: Congenital cytomegalovirus (cCMV) infection is frequent and potentially severe. The immunobiology of cCMV infection is poorly understood, involving cytokines that could be carried within or on the surface of extracellular vesicles (EV). We investigated intra-amniotic cytokines, mediated or not by EV, in [...] Read more.
Background: Congenital cytomegalovirus (cCMV) infection is frequent and potentially severe. The immunobiology of cCMV infection is poorly understood, involving cytokines that could be carried within or on the surface of extracellular vesicles (EV). We investigated intra-amniotic cytokines, mediated or not by EV, in cCMV infection. Methods: Forty infected fetuses following early maternal primary infection and forty negative controls were included. Infected fetuses were classified according to severity at birth: asymptomatic, moderately or severely symptomatic. Following the capture of EV in amniotic fluid (AF), the concentrations of 38 cytokines were quantified. The association with infection and its severity was determined using univariate and multivariate analysis. A prediction analysis based on principal component analysis was conducted. Results: cCMV infection was nominally associated with an increase in six cytokines, mainly soluble (IP-10, IL-18, ITAC, and TRAIL). EV-associated IP-10 was also increased in cases of fetal infection. Severity of fetal infection was nominally associated with an increase in twelve cytokines, including five also associated with fetal infection. A pattern of specific increase in six proteins fitted severely symptomatic infection, including IL-18soluble, TRAILsoluble, CRPsoluble, TRAILsurface, MIGinternal, and RANTESinternal. Conclusion: Fetal infection and its severity are associated with an increase in pro-inflammatory cytokines involved in Th1 immune response. Full article
(This article belongs to the Special Issue Congenital Cytomegalovirus Infection)
Show Figures

Figure 1

Figure 1
<p>Concentrations of relevant cytokines according to fetal infection. Boxplots represent variations in cytokines’ concentrations (median). UF: uninfected fetuses, IF: infected fetuses.</p>
Full article ">Figure 2
<p>Concentrations of relevant cytokines according to symptomatic state at birth and severity. Boxplots representing represent variations in cytokines’ concentrations (median). UF: unifected fetuses, AIF: asymptomatic infected fetuses, SNSIF: symptomatic and non-severe infected fetuses, SSIF: symptomatic and severe infected fetuses.</p>
Full article ">Figure 2 Cont.
<p>Concentrations of relevant cytokines according to symptomatic state at birth and severity. Boxplots representing represent variations in cytokines’ concentrations (median). UF: unifected fetuses, AIF: asymptomatic infected fetuses, SNSIF: symptomatic and non-severe infected fetuses, SSIF: symptomatic and severe infected fetuses.</p>
Full article ">Figure 3
<p>Prediction accuracy of fetal infection (<b>a</b>) and of severity (<b>b</b>).</p>
Full article ">Figure 4
<p>Immunobiology of cCMV: focus on cytokines identified in this study and interactions with infected and immune cells. NKC: natural-killer cell, MP: macrophage, Lc Th1: T-cell involved in Th1 immune response, Lc Th2: T-cell involved in Th2 immune response, DC: dendritic cell, TLR: Toll-like receptor, PAMPS/DAMPs: pathogen-associated molecular patterns/damage-associated molecular patterns.</p>
Full article ">
13 pages, 3408 KiB  
Article
A Secreted Form of the Hepatitis E Virus ORF2 Protein: Design Strategy, Antigenicity and Immunogenicity
by Zihao Chen, Shaoqi Guo, Guanghui Li, Dong Ying, Guiping Wen, Mujin Fang, Yingbin Wang, Zimin Tang, Zizheng Zheng and Ningshao Xia
Viruses 2022, 14(10), 2122; https://doi.org/10.3390/v14102122 - 26 Sep 2022
Cited by 2 | Viewed by 2310
Abstract
Hepatitis E virus (HEV) is an important public health burden worldwide, causing approximately 20 million infections and 70,000 deaths annually. The viral capsid protein is encoded by open reading frame 2 (ORF2) of the HEV genome. Most ORF2 protein present in body fluids [...] Read more.
Hepatitis E virus (HEV) is an important public health burden worldwide, causing approximately 20 million infections and 70,000 deaths annually. The viral capsid protein is encoded by open reading frame 2 (ORF2) of the HEV genome. Most ORF2 protein present in body fluids is the glycosylated secreted form of the protein (ORF2S). A recent study suggested that ORF2S is not necessary for the HEV life cycle. A previously reported efficient HEV cell culture system can be used to understand the origin and life cycle of ORF2S but is not sufficient for functional research. A more rapid and productive method for yielding ORF2S could help to study its antigenicity and immunogenicity. In this study, the ORF2S (tPA) expression construct was designed as a candidate tool. A set of representative anti-HEV monoclonal antibodies was further used to map the functional antigenic sites in the candidates. ORF2S (tPA) was used to study antigenicity and immunogenicity. Indirect ELISA revealed that ORF2S (tPA) was not antigenically identical to HEV 239 antigen (p239). The ORF2S-specific antibodies were successfully induced in one-dose-vaccinated BALB/c mice. The ORF2S-specific antibody response was detected in plasma from HEV-infected patients. Recombinant ORF2S (tPA) can act as a decoy to against B cells. Altogether, our study presents a design strategy for ORF2S expression and indicates that ORF2S (tPA) can be used for functional and structural studies of the HEV life cycle. Full article
(This article belongs to the Special Issue Hepatitis E Virus (HEV))
Show Figures

Figure 1

Figure 1
<p>Design strategy, expression, identification, and purification in constructed ORF2<sup>S</sup> clone. (<b>A</b>) Expi-293F cells were transfected with the ORF2<sup>S</sup> (tPA) constrict. Mass spectrometry indicated that a.a. 34–660 sequence of ORF2 corresponds to the ORF2<sup>S</sup> protein. Then, 10× His-tags were added to the C-termini of the construct genes. Cell supernatants and pellets were collected for further analysis. (<b>B</b>) Given the high sensitivity and specificity of commercial HEV Ag ELISA (Wantai, Beijing), the kit was used to detect the ORF2 protein levels in supernatants and pellets. HEV Ag levels are shown as the signal to cutoff (S/CO) ratio. The original #4 secondary antibody (Wantai, Beijing, China) in the kit was replaced with an anti-His tag secondary antibody (Proteintech) to detect the 10× His tag. Anti-His tag levels are shown as the median effective dose (ED<sub>50</sub>). (<b>C</b>) Supernatants and pellets were collected and analyzed by Western blotting (WB) with mouse monoclonal antibody 1B7, which recognizes the linear epitope HEV ORF2 a.a. 443–457. (<b>D</b>) Glycosylation analysis of ORF2<sup>S</sup> (tPA) proteins in supernatants and pellets. ORF2<sup>S</sup> (tPA) proteins in supernatants and pellets were denatured and digested with indicated glycosidases (+) or without (−). The dashed line shows the mobility shift on nitrocellulose filter membrane of ORF2<sup>S</sup> (tPA) proteins to assess the extent of de-glycosylation. (<b>E</b>) Components of the imidazole-eluted concentration gradient were analyzed by SDS/PAGE and WB. The majority of ORF2<sup>S</sup> was eluted by 250 mM imidazole solution. (<b>F</b>) Gel filtration high-performance liquid chromatography (HPLC). Purified HEV ORF2<sup>S</sup> (tPA) was equilibrated in 20 mM PBS and detected in PBS at OD<sub>280nm</sub>. The column flow rate was maintained at 0.5 mL/min, and the run time was 30 min. N, no heat under nonreducing conditions; H, heat under reducing conditions.</p>
Full article ">Figure 2
<p>Comparison of immunoreactivity and epitope mapping in different HEV capsid proteins (ORF2<sup>S</sup>, p495, p239, and E2). A panel of anti-ORF2 monoclonal antibodies (mAbs) that recognize nine different antigenic epitopes (C1-C6 and L1-L3) was used for detection. (<b>A</b>) Information on the four proteins and epitope regions recognized by a panel of mAbs (adapted with permission from Ref. [A Comprehensive Study of Neutralizing Antigenic Sites on the Hepatitis E Virus (HEV) Capsid by Constructing, Clustering, and Characterizing a Tool Box]. 2015, Zhao et al.). S domain, shell domain; M domain, middle domain; P domain, protruding domain; VLP, virus-like particles. (<b>B</b>) The panel of mAbs used was generated in a previous study [<a href="#B19-viruses-14-02122" class="html-bibr">19</a>]. Indirect ELISA was performed with a goat anti-mouse antibody (Thermo Scientific) as the visualized secondary antibody. The data shown are the concentrations that elicit 50% of maximal effect (EC<sub>50</sub>). C1–C6, Cluster 1–Cluster 6; L1–L3, linear epitope 1–linear epitope 3. (<b>C</b>) Alignment of amino acid sequences (394–606) of four HEV capsid proteins (ORF2<sup>S</sup>, p495, p239, and E2). The blue frame indicates globally similar sites. The red box indicates antigenic sites conserved among the four proteins. Red solid circles, red hollow circles, and red solid triangles represent sites recognized by C3 antibodies, C4 antibodies, and both C3 and C4 antibodies, respectively. Red solid rectangles represent potential glycosylation sites affecting C1 antibody binding.</p>
Full article ">Figure 3
<p>ORF2<sup>S</sup> as an immunogen for assessing immunogenicity. (<b>A</b>) Immunization schemes of the animal experiments. BALB/c mice were administered one dose of 20 µg ORF2<sup>S</sup> (tPA), p239, or PBS, and serum samples were collected weekly. (<b>B</b>) Serum anti-HEV IgG titers over time. The ORF2<sup>S</sup> (tPA) and p239 proteins of HEV gt4 were tested separately. The data are shown as the median effective dose (ED<sub>50</sub>). (<b>C</b>) Comparison of anti-HEV IgG serum titers in mice at the second week. Black bars indicate the standard error of the mean (SEM). The column data show the mean with SEM.</p>
Full article ">Figure 4
<p>Analyses of the antibody response to ORF2<sup>S</sup> in plasma and B cells. (<b>A</b>) Serial dilutions of plasma samples separately bound to ORF2<sup>S</sup> (tPA) and p239. The WHO standard serum was used as a reference to quantify the titer of anti-HEV IgG in naturally infected patients. The black lines are drawn between the same plasma samples. ****, <span class="html-italic">p</span> &lt; 0.0001. (<b>B</b>) Correlation of anti-HEV IgG titers measured for ORF2<sup>S</sup> (tPA) (Y axis) and p239 (X axis). The black line indicates the fitted linear regression. The 95% confidence intervals (CI) are shown in silver. ****, <span class="html-italic">p</span> &lt; 0.0001. (<b>C</b>) Comparison of anti-HEV IgG titers between ORF2<sup>S</sup> and p239. Solid red dots indicate increased anti-HEV IgG titers, and solid blue dots indicate decreased anti-HEV IgG titers binding to ORF2<sup>S</sup> (tPA), compared with p239. Data for anti-HEV IgG titers binding to ORF2<sup>S</sup> (tPA)/p239, converted to log2. (<b>D</b>) Calibration with nonspecific memory B cells using control cells from healthy donors (top right). The proportions of ORF2<sup>S</sup>-specific memory B cells are recognized by two fluorescently labeled ORF2<sup>S</sup> (tPA) probes (labeled with DyLight 488 and allophycocyanin (APC), respectively). Recognized memory B cells from HEV 239-vaccinated donors are shown at the top left. ORF2<sup>S</sup>-specific memory B cells from HEV-infected donors are highlighted in boxes (bottom). (<b>E</b>) The proportions of ORF2<sup>S</sup>-specific memory B cells are displayed as columns. Light blue, red, and dark blue columns indicate vaccinated, naturally infected, and control donors, respectively.</p>
Full article ">
20 pages, 7003 KiB  
Article
Biophysical Modeling of SARS-CoV-2 Assembly: Genome Condensation and Budding
by Siyu Li and Roya Zandi
Viruses 2022, 14(10), 2089; https://doi.org/10.3390/v14102089 - 20 Sep 2022
Cited by 10 | Viewed by 7354
Abstract
The COVID-19 pandemic caused by the Severe Acute Respiratory Syndrome Coronavirus 2 (SARS-CoV-2) has spurred unprecedented and concerted worldwide research to curtail and eradicate this pathogen. SARS-CoV-2 has four structural proteins: Envelope (E), Membrane (M), Nucleocapsid (N), and Spike (S), which self-assemble along [...] Read more.
The COVID-19 pandemic caused by the Severe Acute Respiratory Syndrome Coronavirus 2 (SARS-CoV-2) has spurred unprecedented and concerted worldwide research to curtail and eradicate this pathogen. SARS-CoV-2 has four structural proteins: Envelope (E), Membrane (M), Nucleocapsid (N), and Spike (S), which self-assemble along with its RNA into the infectious virus by budding from intracellular lipid membranes. In this paper, we develop a model to explore the mechanisms of RNA condensation by structural proteins, protein oligomerization and cellular membrane–protein interactions that control the budding process and the ultimate virus structure. Using molecular dynamics simulations, we have deciphered how the positively charged N proteins interact and condense the very long genomic RNA resulting in its packaging by a lipid envelope decorated with structural proteins inside a host cell. Furthermore, considering the length of RNA and the size of the virus, we find that the intrinsic curvature of M proteins is essential for virus budding. While most current research has focused on the S protein, which is responsible for viral entry, and it has been motivated by the need to develop efficacious vaccines, the development of resistance through mutations in this crucial protein makes it essential to elucidate the details of the viral life cycle to identify other drug targets for future therapy. Our simulations will provide insight into the viral life cycle through the assembly of viral particles de novo and potentially identify therapeutic targets for future drug development. Full article
(This article belongs to the Special Issue Physical Virology - Viruses at Multiple Levels of Complexity)
Show Figures

Figure 1

Figure 1
<p>Schematic view of the structure of a coronavirus. The figure shows the structural proteins: Spike (S) proteins (magenta), Membrane (M) proteins (green) and Envelope (E) proteins (yellow). The complex of genome and Nucleocapsid (N) proteins (beige) are enclosed in a lipid membrane called an envelope (red). The picture is adapted from <a href="https://www.scientificanimations.com/coronavirus-symptoms-and-prevention-explained-through-medical-animation" target="_blank">https://www.scientificanimations.com/coronavirus-symptoms-and-prevention-explained-through-medical-animation</a>.</p>
Full article ">Figure 2
<p>Schematic view of an N protein. The dimerization domain (CTD) of the N protein and the RNA binding domain (RBD) are obtained from PDB 6ZCO and 6YI3, respectively [<a href="#B50-viruses-14-02089" class="html-bibr">50</a>,<a href="#B51-viruses-14-02089" class="html-bibr">51</a>]. The intrinsic disorder regions (C<math display="inline"><semantics> <msub> <mrow/> <mi>IDR</mi> </msub> </semantics></math>, Linker<math display="inline"><semantics> <msub> <mrow/> <mi>IDR</mi> </msub> </semantics></math>, N<math display="inline"><semantics> <msub> <mrow/> <mi>IDR</mi> </msub> </semantics></math>) are obtained from UniProt P0DTC9. The visualization is performed through RCSB 3D-view, where the concentrated positive charges are colored light green; the rest of the regions have a zero net charge. The shaded circles indicate the coarse-grained model including three spherical particles: the N-terminal region (N1, red), the linker region (N2, yellow), and the C-terminal region (N3, black).</p>
Full article ">Figure 3
<p>(<b>A</b>) A schematic view of ERGIC. The membrane vertices (Mc) are green and the M protein particles are colored blue-purple-white. The right figure shows the bond connections in the ERGIC membrane built from a triangular lattice. (<b>B</b>) a schematic view of an M protein. The coarse-grained M protein model contains three hard particles: M1 (N-terminal domain, blue); M2 (Transmembrane domain, purple), and M3 (C-terminal or cytosolic domain, white). The arrows indicate the attractive M2–M2 and M3–M3 interactions. The diameters of M1 and M2 are equal while that of M3 is smaller than the other two, which is the source of the intrinsic curvature of M proteins; (<b>C</b>) illustration of the N–N (left) and N–M (right) interactions. The attractive interactions between domains are indicated with arrows, where the C-terminal of the N proteins (N3) attracts each other, and N3 also attracts the C-terminal of M proteins (M3).</p>
Full article ">Figure 4
<p>A schematic diagram of the Bond-Flip Monte Carlo method in which the shared edge of two adjacent triangles are clipped and reconnected to the diagonal vertices to change the local topology of the structure [<a href="#B60-viruses-14-02089" class="html-bibr">60</a>].</p>
Full article ">Figure 5
<p>(<b>A</b>) N proteins condense a linear chain with <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>800</mn> </mrow> </semantics></math> subunits and form RNP. Both cases of the weak (<math display="inline"><semantics> <mrow> <msub> <mi>Z</mi> <mi>g</mi> </msub> <mo>=</mo> <mo>−</mo> <mn>2</mn> </mrow> </semantics></math>) and strong (<math display="inline"><semantics> <mrow> <msub> <mi>Z</mi> <mi>g</mi> </msub> <mo>=</mo> <mo>−</mo> <mn>5</mn> </mrow> </semantics></math>) RNA–N protein interactions are presented. The N protein concentration is 25<math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">μ</mi> <mi mathvariant="normal">M</mi> </mrow> </semantics></math>. In each window, the upper snapshot shows the RNP with genome in blue and N proteins in red and black. The middle and lower snapshots display the RNA and N protein clusters (NPCs), respectively. Note that, if the distance between the N-terminals of N proteins (black) are less than a cut-off distance, they are considered to belong to the same cluster. (<b>B</b>) N proteins condense RNA. RNA is modeled as a branched polymer with the branch number <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>30</mn> </mrow> </semantics></math>. Both cases of the weak and strong RNA–N protein interactions are considered. The branched points and end points are colored in green and yellow, respectively. At the weak RNA–N protein interaction, the N proteins aggregate into clusters in the same way as they do in the case of linear chains, regardless of the branching structure of the chain. The situation differs for the case of the strong RNA–N protein interaction, where the N proteins condense along the branches first, which later aggregate to form a compact RNP.</p>
Full article ">Figure 6
<p>(<b>A</b>) Snapshots of budding through the ERGIC membrane as a function of the strength of the interaction between the M protein transmembrane domains (<span class="html-italic">y</span>-axis) and between the endodomains (<span class="html-italic">x</span>-axis). The interactions become stronger from top to bottom and from left to right. The lipid–lipid interaction strength is kept constant <math display="inline"><semantics> <mrow> <msub> <mi>ϵ</mi> <mrow> <mi>M</mi> <mi>c</mi> <mi>M</mi> <mi>c</mi> </mrow> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> for all simulations; (<b>B</b>) the perimeters (in the unit of <span class="html-italic">a</span>) of the cap built from the M proteins as a function of time in the budding process. The dotted, dashed, and solid lines correspond to the endodomain interaction strengths <math display="inline"><semantics> <msub> <mi>ϵ</mi> <mrow> <msub> <mi>M</mi> <mn>3</mn> </msub> <msub> <mi>M</mi> <mn>3</mn> </msub> </mrow> </msub> </semantics></math> = 0 (a, d, g), 1 (b, e, h) and 2 (c, f, i), respectively. As the strength of <math display="inline"><semantics> <msub> <mi>ϵ</mi> <mrow> <msub> <mi>M</mi> <mn>3</mn> </msub> <msub> <mi>M</mi> <mn>3</mn> </msub> </mrow> </msub> </semantics></math> becomes stronger (the solid line), the budding process becomes faster.</p>
Full article ">Figure 7
<p>Snapshots of simulations of the budding process for two different cases: Scenario 1. Condensation of the genome occurs at the same time as the budding process. In this case, at the beginning of the simulations, the naked branched genome is located near the M proteins embedded in the ERGIC membrane surrounded by N proteins (25 <math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">μ</mi> <mi mathvariant="normal">M</mi> </mrow> </semantics></math>), where the N-terminals of N proteins attract the genome and their C-terminals interact with M proteins, triggering both RNP condensation and budding simultaneously. Scenario 2: N proteins (25 <math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">μ</mi> <mi mathvariant="normal">M</mi> </mrow> </semantics></math>) have already condensed RNA before the transport of RNP to the ERGIC interface. In this scenario, the N proteins within RNP are arranged such that their C-terminals interact with the endodomain of M proteins and trigger the budding process. For clarity, only the N proteins (not RNA) are shown after the completion of the budding process in both cases. For both scenarios, the genome length is <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>800</mn> </mrow> </semantics></math> and genome charge is <math display="inline"><semantics> <mrow> <msub> <mi>Z</mi> <mi>g</mi> </msub> <mo>=</mo> <mo>−</mo> <mn>2</mn> </mrow> </semantics></math>, the interactions between proteins are <math display="inline"><semantics> <mrow> <msub> <mi>ϵ</mi> <mrow> <mi>M</mi> <mi>c</mi> <mi>M</mi> <mi>c</mi> </mrow> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>ϵ</mi> <mrow> <msub> <mi>M</mi> <mn>2</mn> </msub> <msub> <mi>M</mi> <mn>2</mn> </msub> </mrow> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>ϵ</mi> <mrow> <msub> <mi>M</mi> <mn>3</mn> </msub> <msub> <mi>M</mi> <mn>3</mn> </msub> </mrow> </msub> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>ϵ</mi> <mrow> <msub> <mi>N</mi> <mn>3</mn> </msub> <msub> <mi>N</mi> <mn>3</mn> </msub> </mrow> </msub> <mo>=</mo> <mn>20</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <msub> <mi>ϵ</mi> <mrow> <msub> <mi>N</mi> <mn>3</mn> </msub> <msub> <mi>M</mi> <mn>3</mn> </msub> </mrow> </msub> <mo>=</mo> <mn>10</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 8
<p>(<b>A</b>) Snapshots of the simulations of the distribution of the M proteins embedded in the ERGIC membrane as a function of time. Initially, the M proteins are randomly distributed in the ERGIC membrane. The first snapshot at the left illustrates the distribution of the M proteins in ERGIC where the proteins and membrane have been relaxed for 16,000 s with <math display="inline"><semantics> <mrow> <msub> <mi>ϵ</mi> <mrow> <msub> <mi>M</mi> <mn>2</mn> </msub> <msub> <mi>M</mi> <mn>2</mn> </msub> </mrow> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>ϵ</mi> <mrow> <msub> <mi>M</mi> <mn>3</mn> </msub> <msub> <mi>M</mi> <mn>3</mn> </msub> </mrow> </msub> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>. As a result of diffusion, the M proteins have formed a branched-like pattern. In the absence of the M3–M3 interaction, the branched-like pattern does not change (top row). However, with increasing the endodomain–endodomain attractive interaction <math display="inline"><semantics> <msub> <mi>ϵ</mi> <mrow> <msub> <mi>M</mi> <mn>3</mn> </msub> <msub> <mi>M</mi> <mn>3</mn> </msub> </mrow> </msub> </semantics></math>, the connected patches of M proteins begin to aggregate into circular domains (bottom row, indicated with red arrows), which will eventually bud into empty vesicles. Note that the M proteins cover half of the surface of ERGIC, i.e., the M protein fraction is <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>50</mn> <mo>%</mo> </mrow> </semantics></math> during the simulations; (<b>B</b>) snapshots of the simulations of the RNP budding from ERGIC. Initially, the randomly distributed M proteins form a branched-like pattern on the surface of ERGIC in the absence of the endodomain interaction. However, the interaction with RNP results in the aggregation of the proteins in the vicinity of RNP and the formation of the envelope enclosing RNP. The branched-like distribution is obtained by relaxing ERGIC and the M proteins for 16,000 s. The RNP structure is obtained by relaxing N proteins (25 <math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">μ</mi> <mi mathvariant="normal">M</mi> </mrow> </semantics></math>) and RNA for 8000 s. RNA is modeled as a branched polymers with the charge density <math display="inline"><semantics> <mrow> <msub> <mi>Z</mi> <mi>g</mi> </msub> <mo>=</mo> <mo>−</mo> <mn>2</mn> </mrow> </semantics></math> and the number of branch points <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>30</mn> </mrow> </semantics></math>. The comparison of (<b>A</b>,<b>B</b>) reveals that RNP can mediate the interaction between the endodomains giving rise to a preferred curvature between M proteins.</p>
Full article ">
36 pages, 6110 KiB  
Article
Apprehending the NAD+–ADPr-Dependent Systems in the Virus World
by Lakshminarayan M. Iyer, A. Maxwell Burroughs, Vivek Anantharaman and L. Aravind
Viruses 2022, 14(9), 1977; https://doi.org/10.3390/v14091977 - 7 Sep 2022
Cited by 7 | Viewed by 3556
Abstract
NAD+ and ADP-ribose (ADPr)-containing molecules are at the interface of virus–host conflicts across life encompassing RNA processing, restriction, lysogeny/dormancy and functional hijacking. We objectively defined the central components of the NAD+–ADPr networks involved in these conflicts and systematically surveyed 21,191 [...] Read more.
NAD+ and ADP-ribose (ADPr)-containing molecules are at the interface of virus–host conflicts across life encompassing RNA processing, restriction, lysogeny/dormancy and functional hijacking. We objectively defined the central components of the NAD+–ADPr networks involved in these conflicts and systematically surveyed 21,191 completely sequenced viral proteomes representative of all publicly available branches of the viral world to reconstruct a comprehensive picture of the viral NAD+–ADPr systems. These systems have been widely and repeatedly exploited by positive-strand RNA and DNA viruses, especially those with larger genomes and more intricate life-history strategies. We present evidence that ADP-ribosyltransferases (ARTs), ADPr-targeting Macro, NADAR and Nudix proteins are frequently packaged into virions, particularly in phages with contractile tails (Myoviruses), and deployed during infection to modify host macromolecules and counter NAD+-derived signals involved in viral restriction. Genes encoding NAD+–ADPr-utilizing domains were repeatedly exchanged between distantly related viruses, hosts and endo-parasites/symbionts, suggesting selection for them across the virus world. Contextual analysis indicates that the bacteriophage versions of ADPr-targeting domains are more likely to counter soluble ADPr derivatives, while the eukaryotic RNA viral versions might prefer macromolecular ADPr adducts. Finally, we also use comparative genomics to predict host systems involved in countering viral ADP ribosylation of host molecules. Full article
(This article belongs to the Special Issue Phage Assembly Pathways - to the Memory of Lindsay Black)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Structure of NAD<sup>+</sup> and (<b>b</b>) the substrates and products of various enzymes in the NAD<sup>+</sup>–ADPr network. Enzymes are color-coded based on the pathway in which they are involved. Bonds that are the target of particular enzymes are highlighted with colored circles.</p>
Full article ">Figure 2
<p>Distribution of NAD<sup>+</sup>–ADPr network domains in the virus world. (<b>a</b>) Fraction of RNA (left) and DNA (right) viruses with a genome ≥ a given length containing NAD<sup>+</sup>–ADPr network domains. (<b>b</b>) Distribution of various NAD<sup>+</sup>-ADPR domains in the Myoviruses, Siphoviruses and Podoviruses. The graphs depict the number of phages per 1000 containing the given domain in that morphological category. (<b>c</b>) Prevalence of various NAD<sup>+</sup>–ADPr processing domains depicted as a percentage of the total number.</p>
Full article ">Figure 3
<p>Phylogenetic trees of select viral NAD<sup>+</sup>–ADPr network components illustrating their origins as described in the text. (<b>a</b>) TIR, (<b>b</b>) Arc, (<b>c</b>) VIP2-like ART, (<b>d</b>) PART, (<b>e</b>) Ecto-ART, (<b>f</b>) Macro domain. Clades with a bootstrap support of &gt; 75% are marked by colored circles. Several clades are collapsed in the trees for brevity. Relevant exchanges of genes are indicated. The raw data for the phylogenetic trees can be obtained from <a href="#app1-viruses-14-01977" class="html-app">Supplementary S3</a>.</p>
Full article ">Figure 4
<p>(<b>a</b>) Topology diagram of the core of the Macro domain and the EF-Tu-like GTPases illustrating their structural relationship. Strands labeled with an ‘S’ prefix followed by their order number in the core structure are shown as yellow arrows, whereas helices which are similarly labeled with a ‘H’ prefix are shown as red cylinders. (<b>b</b>) The structure of the Nlig-Ia domain (cartoon rendering) that exists as a solo domain only in viruses. The figure illustrates the residues involved in binding NAD<sup>+</sup> and its relative position with respect to the C-terminal ATP-grasp and RAGNYA domain of the NAD<sup>+</sup>-dependent ligases (rendered as a tube).</p>
Full article ">Figure 5
<p>Representative contextual associations including domain architectures and gene neighborhoods of various domains of the NAD<sup>+</sup>–ADPr network. Gene neighborhoods are shown as box arrows with the arrowhead pointing to the 3′ gene. Domain architectures are shown by other shapes. The contextual associations are categorized based on their genomic contexts or their function including (<b>a</b>) domains associated with the Terminase-portal genes and those encoding other virion components; (<b>b</b>) domains involved in NAD<sup>+</sup> synthesis; (<b>c</b>) secreted toxin domains; (<b>d</b>) domains that are components of T–A and related conflict systems; (<b>e</b>) domains in RNA virus polyproteins; (<b>f</b>) domains involved in a predicted RNA repair system; (<b>g</b>) viral TIR systems; (<b>h</b>) domains involved in ADPr-processing and; (<b>i</b>) SLOG sensor-activated systems. Gene neighborhoods are labeled with the accession number and species name of the gene marked with an asterisk.</p>
Full article ">Figure 6
<p>Complete and pairwise co-occurrence patterns of NAD<sup>+</sup>–ADPr domains depicted as Euler diagrams. (<b>a</b>,<b>b</b>) Domains involved in NAD<sup>+</sup> biosynthesis/salvage in viruses. (<b>c</b>,<b>d</b>) Co-occurrence of the Macro, SLOG, Nudix and NADAR domains in DNA viruses. Co-occurrences are measured as a percentage of all the proteins that are being compared in a particular Euler diagram.</p>
Full article ">Figure 7
<p>Contextual associations including (<b>a</b>) domain architectures and (<b>b</b>) gene neighborhoods of the ARG-associated systems. (<b>c</b>) More gene neighborhoods of the ARPP domain. (<b>d</b>) Contextual network diagram and (<b>e</b>) co-occurrence frequencies of domains associated with the ARG domain. (<b>f</b>) Structural comparison of the newly identified members of the TY-chaperone superfamily found in these systems.</p>
Full article ">
18 pages, 3420 KiB  
Article
Structural Dynamics and Activity of B19V VP1u during the pHs of Cell Entry and Endosomal Trafficking
by Renuk V. Lakshmanan, Joshua A. Hull, Luke Berry, Matthew Burg, Brian Bothner, Robert McKenna and Mavis Agbandje-McKenna
Viruses 2022, 14(9), 1922; https://doi.org/10.3390/v14091922 - 30 Aug 2022
Cited by 4 | Viewed by 2573
Abstract
Parvovirus B19 (B19V) is a human pathogen that is the causative agent of fifth disease in children. It is also known to cause hydrops in fetuses, anemia in AIDS patients, and transient aplastic crisis in patients with sickle cell disease. The unique N-terminus [...] Read more.
Parvovirus B19 (B19V) is a human pathogen that is the causative agent of fifth disease in children. It is also known to cause hydrops in fetuses, anemia in AIDS patients, and transient aplastic crisis in patients with sickle cell disease. The unique N-terminus of Viral Protein 1 (VP1u) of parvoviruses, including B19V, exhibits phospholipase A2 (PLA2) activity, which is required for endosomal escape. Presented is the structural dynamics of B19V VP1u under conditions that mimic the pHs of cell entry and endosomal trafficking to the nucleus. Using circular dichroism spectroscopy, the receptor-binding domain of B19V VP1u is shown to exhibit an α-helical fold, whereas the PLA2 domain exhibits a probable molten globule state, both of which are pH invariant. Differential scanning calorimetry performed at endosomal pHs shows that the melting temperature (Tm) of VP1u PLA2 domain is tuned to body temperature (37 °C) at pH 7.4. In addition, PLA2 assays performed at temperatures ranging from 25–45 °C show both a temperature and pH-dependent change in activity. We hypothesize that VP1u PLA2 domain differences in Tm at differing pHs have enabled the virus to “switch on/off” the phospholipase activity during capsid trafficking. Furthermore, we propose the environment of the early endosome as the optimal condition for endosomal escape leading to B19V infection. Full article
(This article belongs to the Special Issue Viral Accessory Proteins)
Show Figures

Figure 1

Figure 1
<p>SDS PAGE and structure of B19V VP1u. (<b>A</b>) SDS PAGE profile of B19V VP1u, domain 1, and domain 2. (<b>B</b>) Representative CD spectrum of B19V VP1u, RBD (domain 1), and PLA<sub>2</sub> (domain 2) recorded in water at 20 °C. (<b>C)</b> Schematic of B19V VP1u with the RBD (dark brown) and PLA<sub>2</sub> domain (green). (<b>D</b>) In silico model of B19V VP1u generated using RoseTTAFold with the RBD (dark brown) and PLA<sub>2</sub> domain (green).</p>
Full article ">Figure 2
<p>PLA<sub>2</sub> activity and pH dependence of secondary structure. (<b>A</b>–<b>C</b>) Representative CD spectrum of B19V VP1u (left), RBD (domain 1, middle) and PLA<sub>2</sub> domain (domain 2, right) at pH 4.0, 5.5, 6.0, and 7.4 measured at 20 °C. (<b>D</b>) PLA<sub>2</sub> activity of 250 ng of B19V VP1u, domain 2 and domain 1. Calculated CD spectrum of RBD (<b>E</b>) and PLA<sub>2</sub> domain (<b>F</b>) at different pHs.</p>
Full article ">Figure 3
<p>Thermostability of B19V capsid protein. (<b>A</b>) Comparison of experimental baseline corrected DSC thermograms of VP1u at different pHs. (<b>B</b>) pH dependent thermostability profile of B19V VP1u based on the major endothermic peak. (<b>C</b>) pH dependent thermostability profile of B19V VLPs. All experimental values are shown as means ± standard deviation (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 4
<p>Effect of heat on secondary structure of B19V VP1u. The CD spectrum was recorded at pHs 4.0, 5.5, 6.0 and 7.4.</p>
Full article ">Figure 5
<p>PLA2 catalytic residues and activity. (<b>A</b>) Structure superposition of model of B19V PLA2 domain (blue) on Bee venom PLA2 crystal structure (color- beige, Cα rmsd = 1.1 Å, PDB- 1POC). VP1u PLA2 catalytic residues His153 and Asp175 are highlighted. Asp154 is predicted to bind calcium (Green sphere). (<b>B</b>) Plot shows pH dependence of PLA2 activity measured at 25 °C. All experimental values are shown as means ± standard deviation (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 6
<p>Effect of temperature and pH on VP1u PLA2 activity. (<b>A</b>) SDS PAGE of B19V VP1u. (<b>B</b>) PLA2 activity of B19V VP1u (250 ng) plotted as relative maximum activity. (<b>C</b>) PLA2 activity of B19V VP1u quantified at different temperatures and pHs. (<b>D</b>) SDS PAGE of MVM VP1u. (<b>E</b>) PLA2 activity of MVM VP1u (30 ng) plotted as relative maximum activity. (<b>F</b>) PLA2 activity of MVM VP1u quantified at different temperatures and pHs. All experimental values are shown as means ± standard deviation (<span class="html-italic">n</span> = 6).</p>
Full article ">Figure 7
<p>Effect of temperature and pH on PLA2 activity of B19V VLPs. (<b>A</b>) SDS PAGE of purified VLP showing VP1 and VP2. (<b>B</b>) Negative-stain EM showing assembled virions. (<b>C</b>) PLA2 activity of B19V VLPs (180 ng of VP1) measured at different temperatures and pHs plotted as relative maximum activity. (<b>D</b>) Quantification of PLA2 activity of B19V VLPs at different pHs. All experimental values are shown as means ± standard deviation (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 8
<p>Proposed model for internalization and cellular trafficking of parvovirus B19. Made using Servier medical ART (<a href="https://creativecommons.org/licenses/by/3.0/" target="_blank">https://creativecommons.org/licenses/by/3.0/</a>). Accessed on 20 May 2021.</p>
Full article ">
14 pages, 18553 KiB  
Article
Female Genital Fibroblasts Diminish the In Vitro Efficacy of PrEP against HIV
by Ashley F. George, Matthew McGregor, David Gingrich, Jason Neidleman, Rebecca S. Marquez, Kyrlia C. Young, Kaavya L. Thanigaivelan, Warner C. Greene, Phyllis C. Tien, Amelia N. Deitchman, Trimble L. Spitzer and Nadia R. Roan
Viruses 2022, 14(8), 1723; https://doi.org/10.3390/v14081723 - 4 Aug 2022
Viewed by 2303
Abstract
The efficacy of HIV pre-exposure prophylaxis (PrEP) is high in men who have sex with men, but much more variable in women, in a manner largely attributed to low adherence. This reduced efficacy, however, could also reflect biological factors. Transmission to women is [...] Read more.
The efficacy of HIV pre-exposure prophylaxis (PrEP) is high in men who have sex with men, but much more variable in women, in a manner largely attributed to low adherence. This reduced efficacy, however, could also reflect biological factors. Transmission to women is typically via the female reproductive tract (FRT), and vaginal dysbiosis, genital inflammation, and other factors specific to the FRT mucosa can all increase transmission risk. We have demonstrated that mucosal fibroblasts from the lower and upper FRT can markedly enhance HIV infection of CD4+ T cells. Given the current testing of tenofovir disoproxil fumarate, cabotegravir, and dapivirine regimens as candidate PrEP agents for women, we set out to determine using in vitro assays whether endometrial stromal fibroblasts (eSF) isolated from the FRT can affect the anti-HIV activity of these PrEP drugs. We found that PrEP drugs exhibit significantly reduced antiviral efficacy in the presence of eSFs, not because of decreased PrEP drug availability, but rather of eSF-mediated enhancement of HIV infection. These findings suggest that drug combinations that target both the virus and infection-promoting factors in the FRT—such as mucosal fibroblasts—may be more effective than PrEP alone at preventing sexual transmission of HIV to women. Full article
(This article belongs to the Special Issue Women in Virology)
Show Figures

Figure 1

Figure 1
<p>The tested concentrations of the PrEP drugs tenofovir disoproxil fumarate (TDF), cabotegravir, and dapivirine do not diminish PBMC viability in the absence or presence of eSFs. PHA/IL2-activated PBMCs were incubated with or without a HIV luciferase reporter virus (CXCR4-tropic NL4-3.Luc) and the indicated concentrations of (<bold>A</bold>) TDF, (<bold>B</bold>) cabotegravir, or (<bold>C</bold>) dapivirine in the absence or presence of eSFs. PBMC viability was monitored 3 days later by the CellTiter-Glo<sup>®</sup> Luminescent Cell Viability Assay. Mean luminescence ± SD derived from sextuplet experimental replicates are shown (RLU/s, relative light units per second). The limit of detection for luminescence ranged from 1 × 10<sup>2</sup> to 1 × 10<sup>10</sup> RLUs. Non-significant (<italic>p</italic> &lt; 0.05) in a group-wise comparison (one-way analysis of variance with a Tukey post-test).</p>
Full article ">Figure 2
<p>PrEP drugs tenofovir disoproxil fumarate (TDF), cabotegravir, and dapivirine lose antiviral activity in the presence of eSF. PHA/IL2-activated PBMCs were incubated with NL4-3.Luc at the indicated concentrations of (<bold>A</bold>) TDF, (<bold>B</bold>) cabotegravir, or (<bold>C</bold>) dapivirine in the absence or presence of eSF. Infection levels were monitored 3 days later by luminescence. Mean luciferase activities ± SD derived from sextuplet experimental replicates are shown (RLU/s, relative light units per second). The limit of detection for luminescence ranged from 1 × 10<sup>2</sup> to 1 × 10<sup>10</sup> RLUs. **** <italic>p</italic> &lt; 0001 relative to no eSF coculture in a group-wise comparison (two-way analysis of variance with a Tukey post-test).</p>
Full article ">Figure 3
<p>eSFs from multiple donors all diminish the in vitro efficacy of tenofovir disoproxil fumarate (TDF), cabotegravir, and dapivirine. Infection rates of PBMCs from a single donor exposed to (<bold>A</bold>) TDF, (<bold>B</bold>) cabotegravir, or (<bold>C</bold>) dapivirine in the presence or absence of eSFs from multiple donors. PHA/IL2-activated PBMCs treated with the indicated drug were mock-inoculated or inoculated with CXCR4-tropic NL4-3.Luc, in the absence or presence of eSFs (N = 4–5), and harvested 3 days later. Mean luciferase activities ± SD are shown. The limit of detection for luminescence ranged from 1 × 10<sup>2</sup> to 1 × 10<sup>10</sup> RLUs. Open circles indicate drug concentrations where infection was undetectable in PBMCs without eSFs. Filled circles indicate drug concentrations where infection was undetectable in PBMCs with or without eSFs. **** <italic>p</italic> &lt; 0001 relative to no coculture in a group-wise comparison (two-way analysis of variance with a Tukey post-test).</p>
Full article ">Figure 4
<p>eSFs diminish the in vitro efficacy of tenofovir disoproxil fumarate (TDF), cabotegravir, and dapivirine in PBMCs from multiple donors. Infection rates of PBMCs (N = 4) exposed to increasing concentrations of (<bold>A</bold>) TDF, (<bold>B</bold>) cabotegravir, or (<bold>C</bold>) dapivirine are shown. PHA/IL2-activated PBMCs treated with the indicated drug were mock-treated or inoculated with NL4-3.Luc, in the absence or presence of eSF, and harvested 3 days later. Mean luciferase activities ± SD are shown. The limit of detection for luminescence ranged from 1 × 10<sup>2</sup> to 1 × 10<sup>10</sup> RLUs. Open circles indicate drug concentrations where infection was undetectable in PBMCs without eSFs. Filled circles indicate drug concentrations where infection was undetectable in PBMCs with or without eSFs. **** <italic>p</italic> &lt; 0001 relative to no coculture in a group-wise comparison (two-way analysis of variance with a Tukey post-test).</p>
Full article ">Figure 5
<p>eSFs diminish in vitro efficacy of PrEP drugs against R5-tropic HIV infection. PHA/IL2-activated PBMCs exposed to increasing concentrations of tenofovir disoproxil fumarate (TDF), cabotegravir, or dapivirine cultured with or without eSF (N = 3), were infected with a CCR5-tropic transmitter/founder HSA-reporter virus, and monitored by flow cytometry for infection levels 3 days later. (<bold>A</bold>) Gating strategy for identification of infected cells. (<bold>B</bold>,<bold>C</bold>) eSFs diminish HIV infection rates in the presence of PrEP. Mean % of live, singlet CD3+CD8-HSA+ cells ± SD for (<bold>B</bold>) TDF, (<bold>C</bold>) cabotegravir, or (<bold>D</bold>) dapivirine are shown. Open circles indicate drug concentrations where infection was undetectable in PBMCs without eSFs. Filled circles indicate drug concentrations where infection was undetectable in PBMCs with or without eSFs. **** <italic>p</italic> &lt; 0001 relative to no coculture in a group-wise comparison (two-way analysis of variance with a Tukey post-test).</p>
Full article ">
21 pages, 3243 KiB  
Article
TRIM7 Restricts Coxsackievirus and Norovirus Infection by Detecting the C-Terminal Glutamine Generated by 3C Protease Processing
by Jakub Luptak, Donna L. Mallery, Aminu S. Jahun, Anna Albecka, Dean Clift, Osaid Ather, Greg Slodkowicz, Ian Goodfellow and Leo C. James
Viruses 2022, 14(8), 1610; https://doi.org/10.3390/v14081610 - 23 Jul 2022
Cited by 4 | Viewed by 3494
Abstract
TRIM7 catalyzes the ubiquitination of multiple substrates with unrelated biological functions. This cross-reactivity is at odds with the specificity usually displayed by enzymes, including ubiquitin ligases. Here we show that TRIM7′s extreme substrate promiscuity is due to a highly unusual binding mechanism, in [...] Read more.
TRIM7 catalyzes the ubiquitination of multiple substrates with unrelated biological functions. This cross-reactivity is at odds with the specificity usually displayed by enzymes, including ubiquitin ligases. Here we show that TRIM7′s extreme substrate promiscuity is due to a highly unusual binding mechanism, in which the PRYSPRY domain captures any ligand with a C-terminal helix that terminates in a hydrophobic residue followed by a glutamine. Many of the non-structural proteins found in RNA viruses contain C-terminal glutamines as a result of polyprotein cleavage by 3C protease. This viral processing strategy generates novel substrates for TRIM7 and explains its ability to inhibit Coxsackie virus and norovirus replication. In addition to viral proteins, cellular proteins such as glycogenin have evolved C-termini that make them a TRIM7 substrate. The ‘helix-ΦQ’ degron motif recognized by TRIM7 is reminiscent of the N-end degron system and is found in ~1% of cellular proteins. These features, together with TRIM7′s restricted tissue expression and lack of immune regulation, suggest that viral restriction may not be its physiological function. Full article
(This article belongs to the Section Human Virology and Viral Diseases)
Show Figures

Figure 1

Figure 1
<p>A linear epitope of GYG1 is sufficient to explain TRIM7 binding. (<b>A</b>) Sequence and domain organization of TRIM7 (grey) and Glycogenin1 (pale orange). TRIM7 is also depicted as a cartoon (circle—RING, rectangle—B-Box, line—Coiled-coil, cut-out circle—PRYSPRY) used throughout this paper. Highlighted are sequences determining TRIM7 binding. (<b>B</b>) ITC binding experiments with sequential truncations of rbGYG1 (shown above each titration) titrated into human hisTRIM7-PRYSPRY. Representative traces and their accompanying fitted KD’s are shown. (<b>C</b>) Cartoon overview of the crystal structure of hisTRIM7-PRYSPRY (grey) with the GYG1 peptide (orange) on left (PDB accession 7OVX). Comparison with TRIM21:Fc structure (2IWG, PRYSPRY wheat, Fc magenta) on the right.</p>
Full article ">Figure 2
<p>TRIM7 binds diverse substrates using a C-terminal helix-ΦQ motif. (<b>A</b>) MNV1 polyprotein (pale blue) and the relevant proteins in a darker shade (NS3 and NS6), CVB3 polyprotein in pale green and the relevant 2C protein in a darker shade. (<b>B</b>) Main peptide sequences used in the binding and structural experiments. Outlined are the sequences synthesized, whilst the shaded residues are those resolved in the crystal structures. Color coding is maintained throughout the figure. (<b>C</b>) hisTRIM7-PRYSPRY:Peptide complex structures superposed. Shows PRYSPRY as a transparent surface representation with a few key residues in stick. Peptides are color coded as described. PDB accession codes are 7OW2, 7OVX, 8A5L and 8A5M). (<b>D</b>) Detail of the PRYSPRY pocket with the recognition motif bound. Several key residues are highlighted. Dashes indicate charged or H-bonding between the peptide and the PRYSPRY residues. Underlined residues are essential for binding (see <a href="#app1-viruses-14-01610" class="html-app">Figure S3</a>). (<b>E</b>) Peptide substitutions based on the TIEALFQ peptide and possible polyprotein processing ends. Binding experiments with the minimal LQ motif, sequence is shown and the derived KD from ITC titrations. No binding* denotes where the peptide was only tested using nanoDSF. (<b>F</b>) Thermal denaturation data of hTRIM7-PRYSPRY derived using the Prometheus nanoDSF. Shows the Tm of the protein in the presence of peptide or DMSO. Peptides which bind stabilize the protein. The results of two independent measurements are shown. (<b>G</b>,<b>H</b>) Avidity enhancement of binding affinity with hisMBP-TRIM7-CC-PRYSPRY. (<b>G</b>) Shows the AlphaFold prediction of the TRIM7 dimer (grey and wheat) aligned with the crystal structure (pale blue and orange). (<b>H</b>) The full-length GYG1 protein (isoform GN1) shows clear 1:1 binding between the protein dimers, whilst no difference in affinity is observed when the peptide is the substrate. Representative traces and their accompanying fitted KD’s are shown.</p>
Full article ">Figure 3
<p>TRIM7 co-localizes with helix-ΦQ containing substrates inside cells and degrades them. (<b>A</b>–<b>E</b>) Schematic representation of in vitro ubiquitination experiments. (<b>A</b>) TRIM7-RING (grey circle) was mixed with combinations of Ubiquitin (Ub, blue circle), Ube2N/Ube2V2 (purple semi-circle) and Ube2W (green teardrop) as shown and the reactions followed by immunoblotting (<b>C</b>) with an anti-TRIM7 antibody. A representative blot from at least three independent experiments is shown. (<b>B</b>) E2 discharge experiment: Ube2N~Ub complex was incubated with TRIM7-RING, and the reaction followed over time as indicated in (<b>D</b>). Blots were probed with anti-ubiquitin antibody. A single asterisk denotes the charged Ube2N~Ub, whereas a double asterisk denotes the uncharged Ube2N. A representative blot from at least three independent experiments is shown. (<b>E</b>) Densitometry quantification of band intensities from 3D was plotted to show the kinetics of E2-Ub discharge. Error bars in all graphs depict the mean +/− SEM. Data represent three independent replicates. (<b>F</b>) Western blots from cells overexpressing indicated constructs of epitope-tagged TRIM7 (probed with anti-HA antibody). Ubiquitin-laddering is lost when the RING domain is deleted. A representative blot from two independent experiments is shown. (<b>G</b>) Schematic overview of the experiments presented in (<b>H</b>). EGFP-GYG1 is represented as a green-brown shape. TRIM7 is shown as usual but with N-terminal mCherry (magenta circle). Plasmids were co-expressed, and the fluorescence monitored using live imaging and the protein levels quantified using either the fluorescence intensity or using cell lysates. (<b>H</b>) Live-cell microscopy of U2OS cells expressing mCherry-TRIM7 and EGFP-GYG1 constructs. Left column shows the EGFP signal (green), middle column shows the mCherry signal (magenta) and the right column shows the false-colored merged image (EGFP—green; mCherry—magenta; merge—white). The scale is the same in all images, and the bar represents 10 µm. Rows represent different conditions: Top is both WT sequences. Middle has a Q333A mutation in the EGFP-GYG1 construct. Bottom has the R385A mutation in the mCh-TRIM7 con-struct. Example images are shown from at least two independent experiments. (<b>I</b>) Line profile analysis (ImageJ) of the fluorescent signal. Green trace shows the EGFP signal whilst the magenta trace shows the mCherry signal. The line used in the analysis is shown on the merged signal images. (<b>J</b>) Fluorescence-based quantification (using the IncuCyte) of EGFP-GYG1 protein levels, graph shows the total integrated intensity of the EGFP signal divided by the EGFP area from three biological replicates. ANOVA was used for statistical analysis and significant differences from T7 + GYG condition indicated (<span class="html-italic">p</span> &lt; 0.0005 (***), <span class="html-italic">p</span> &lt; 0.0001 (****)). (<b>K</b>,<b>L</b>) Protein quantification and blots using cell lysates and capillary-based Western Blot (Jess). Values from two biological replicates were normalized to loading control (actin) and the fraction of cells transfected (see methods). ANOVA was used for statistical analysis and significant differences from T7 + GYG condition indicated (<span class="html-italic">p</span> &lt; 0.05 (*), <span class="html-italic">p</span> &lt; 0.005 (**)). Black asterisks indicate significance in EGFP-GYG expression and pink asterisks in mCh-TRIM7 expression.</p>
Full article ">Figure 4
<p>TRIM7 restriction of MNV1 and CVB3 infection is driven by helix-ΦQ binding. (<b>A</b>–<b>C</b>) Co-transfection of TRIM constructs and reporter plasmids. Overexpression of TRIM7 induces strong signaling by AP-1 and NF-κB that is dependent on the RING domain. (<b>D</b>) Overexpression of TRIM7 induces signaling, as measured by qPCR; values are normalized to actin. (<b>E</b>) Point mutation that prevents binding to targets does not impact signaling. For all signaling experiments (<b>A</b>–<b>E</b>), representative examples of at least two independent experiments are shown, with values normalized to cells transfected with empty vector. ANOVA was used for statistical analysis and significant differences indicated (<span class="html-italic">p</span> &lt; 0.0005 (***), <span class="html-italic">p</span> &lt; 0.0001 (****)). (<b>F</b>,<b>G</b>) Schematic overview of the experiments shown in H and I. Lentivirus generation and stable transfection of HeLa-CD300lf cells (<b>F</b>), with the expression confirmed by Western blotting (<a href="#app1-viruses-14-01610" class="html-app">Figure S6</a>). Cells were infected with recombinant CVB3 virus-producing EGFP in host cells or MNV1 (<b>G</b>). (<b>H</b>) Quantification of CVB3 infection by measuring the fraction of EGFP-expressing cells. Data shown are from three independent experiments. (<b>I</b>) Quantification of MNV1 replication by TCID50 following infection of TRIM7-expressing cells. Virions from lysed cells were titrated onto susceptible BV-2 cells. Data shown are a representative result of three independent experiments. Error bars show the standard deviation. Non-parametric ANOVA was used for statistical analysis and significant differences indicated (<span class="html-italic">p</span> &lt; 0.05 (*), <span class="html-italic">p</span> &lt; 0.0005 (***), <span class="html-italic">p</span> &lt; 0.0001 (****)).</p>
Full article ">
17 pages, 2618 KiB  
Article
Phosphomimetic S207D Lysyl–tRNA Synthetase Binds HIV-1 5′UTR in an Open Conformation and Increases RNA Dynamics
by William A. Cantara, Chathuri Pathirage, Joshua Hatterschide, Erik D. Olson and Karin Musier-Forsyth
Viruses 2022, 14(7), 1556; https://doi.org/10.3390/v14071556 - 16 Jul 2022
Cited by 4 | Viewed by 2696
Abstract
Interactions between lysyl–tRNA synthetase (LysRS) and HIV-1 Gag facilitate selective packaging of the HIV-1 reverse transcription primer, tRNALys3. During HIV-1 infection, LysRS is phosphorylated at S207, released from a multi-aminoacyl–tRNA synthetase complex and packaged into progeny virions. LysRS is critical for [...] Read more.
Interactions between lysyl–tRNA synthetase (LysRS) and HIV-1 Gag facilitate selective packaging of the HIV-1 reverse transcription primer, tRNALys3. During HIV-1 infection, LysRS is phosphorylated at S207, released from a multi-aminoacyl–tRNA synthetase complex and packaged into progeny virions. LysRS is critical for proper targeting of tRNALys3 to the primer-binding site (PBS) by specifically binding a PBS-adjacent tRNA-like element (TLE), which promotes release of the tRNA proximal to the PBS. However, whether LysRS phosphorylation plays a role in this process remains unknown. Here, we used a combination of binding assays, RNA chemical probing, and small-angle X-ray scattering to show that both wild-type (WT) and a phosphomimetic S207D LysRS mutant bind similarly to the HIV-1 genomic RNA (gRNA) 5′UTR via direct interactions with the TLE and stem loop 1 (SL1) and have a modest preference for binding dimeric gRNA. Unlike WT, S207D LysRS bound in an open conformation and increased the dynamics of both the PBS region and SL1. A new working model is proposed wherein a dimeric phosphorylated LysRS/tRNA complex binds to a gRNA dimer to facilitate tRNA primer release and placement onto the PBS. Future anti-viral strategies that prevent this host factor-gRNA interaction are envisioned. Full article
(This article belongs to the Special Issue Regulatory Mechanisms of Viral UTRs)
Show Figures

Figure 1

Figure 1
<p>Proteins and RNAs used in this study. (<b>A</b>) The 532-residue human LysRS construct used for all experiments lacks the <span class="html-italic">N</span>-terminal 65 residues (LysRS(∆N65)). The phosphomimetic mutant replaces a serine at position 207 with an aspartate in the anticodon binding domain (S207D, red asterisk). (<b>B</b>) The UTR<sub>240</sub> construct contains the U5/AUG stem (black), PBS/TLE domain (red, nt 126–224), and Psi domain (blue, nt 117–125 and 225–332). The gray boxed regions indicate the mutation that was made to replace the DIS with a stable GNRA tetraloop (∆DIS). The PBS/TLE domain was also deleted for some studies (ΔPBS/TLE) and replaced with GAGA. Nt in gray circles indicate mutations that were made to promote transcription efficiency and stabilize the terminal stem, as described in the Methods. The secondary structure shown is based on Ref [<a href="#B38-viruses-14-01556" class="html-bibr">38</a>].</p>
Full article ">Figure 2
<p>SAXS analysis of LysRS-bound PBS/TLE complexes. The SAXS-derived ab initio envelope for PBS/TLE alone (blue) was overlaid with the envelopes calculated for PBS/TLE bound to either WT LysRS(∆N65) (grey mesh) or S207D LysRS(∆N65) (brown mesh).</p>
Full article ">Figure 3
<p>XL-SHAPE results for LysRS(∆N65) variants binding to WT UTR<sub>240</sub>. The baseline SHAPE reactivities are shown as colored circles behind each nt. Grey circles indicate the region of the structure that was not probed in our studies. SHAPE reactivity changes that occurred upon titration of WT LysRS(∆N65) (<b>left</b>) and S207D LysRS(∆N65) (<b>right</b>) are identified with colored arrows (red = increased reactivity, blue = decreased reactivity) and crosslinked sites are denoted with asterisks.</p>
Full article ">Figure 4
<p>SAXS analysis of the UTR<sub>240</sub>(ΔDIS). (<b>A</b>) SAXS envelope of the UTR<sub>240</sub>(ΔDIS) RNA (grey mesh). (<b>B</b>) Alignment of the UTR<sub>240</sub>(ΔDIS) with the envelope for UTR<sub>240</sub>(ΔDIS,ΔPBS/TLE) (blue spheres). (<b>C</b>) Each of the individual helices (PBS/TLE, red; U5-AUG, yellow; SL1, orange; SL3, purple) from a previously reported NMR structure [<a href="#B38-viruses-14-01556" class="html-bibr">38</a>] fit into specific regions of the UTR<sub>240</sub>(ΔDIS) SAXS envelope. (<b>D</b>) LysRS dimer crystal structure depicted at the same scale as the SAXS data with one monomer in red and the other in blue. Black brackets indicate approximate locations of LysRS binding sites determined by crosslinking.</p>
Full article ">Figure 5
<p>SAXS analysis of extended primer complexes. (<b>A</b>) Schematic of construct design for annealing of extended DNA primers (italic font) designed to disrupt the structure of the TLE stem. (<b>B</b>) SAXS-derived molecular envelope of the antiPBS<sub>18</sub>-annealed PBS/TLE construct. (<b>C</b>) SAXS-derived molecular envelopes for the PBS/TLE RNA with DNA primers of different lengths annealed. When aligned using the PBS loop as a reference, the location of the A-Bulge (cyan dot) rotates counterclockwise about the structure.</p>
Full article ">Figure 6
<p>Model of pS207–LysRS-directed primer placement to PBS and subsequent pS207–LysRS release from HIV-1 gRNA. (Step 1) Upon HIV-1 infection, LysRS is phosphorylated on S207 and is released from the multi-aminoacyl-tRNA synthetase complex (MSC) in a conformation inactive for tRNA aminoacylation. (Step 2) The HIV-1 gRNA can adopt various monomeric and dimeric conformations, including the “kissing loop” dimer shown. (Step 3) The catalytically inactive, tRNA-bound pS207–LysRS dimer preferentially binds to the PBS/TLE and SL1 regions of an HIV-1 gRNA dimer. This binding results in release of tRNA from LysRS and an increase in the dynamics of the PBS region and SL1 hairpin (blue squiggles). (Step 4) HIV-1 Gag facilitates annealing of the tRNA primer onto the PBS via chaperone activity of the nucleocapsid domain. (Step 5) After maturation, reverse transcriptase (RT) binds the primer:template complex and (Step 6) initiates proviral DNA synthesis (purple extension) leading to disruption of the TLE conformation and pS207–LysRS release. For clarity, the second gRNA monomer is not shown after Step 4.</p>
Full article ">
19 pages, 2213 KiB  
Article
Origins and Evolution of Seasonal Human Coronaviruses
by James R. Otieno, Joshua L. Cherry, David J. Spiro, Martha I. Nelson and Nídia S. Trovão
Viruses 2022, 14(7), 1551; https://doi.org/10.3390/v14071551 - 15 Jul 2022
Cited by 5 | Viewed by 3467
Abstract
Four seasonal human coronaviruses (sHCoVs) are endemic globally (229E, NL63, OC43, and HKU1), accounting for 5–30% of human respiratory infections. However, the epidemiology and evolution of these CoVs remain understudied due to their association with mild symptomatology. Using a multigene and complete genome [...] Read more.
Four seasonal human coronaviruses (sHCoVs) are endemic globally (229E, NL63, OC43, and HKU1), accounting for 5–30% of human respiratory infections. However, the epidemiology and evolution of these CoVs remain understudied due to their association with mild symptomatology. Using a multigene and complete genome analysis approach, we find the evolutionary histories of sHCoVs to be highly complex, owing to frequent recombination of CoVs including within and between sHCoVs, and uncertain, due to the under sampling of non-human viruses. The recombination rate was highest for 229E and OC43 whereas substitutions per recombination event were highest in NL63 and HKU1. Depending on the gene studied, OC43 may have ungulate, canine, or rabbit CoV ancestors. 229E may have origins in a bat, camel, or an unsampled intermediate host. HKU1 had the earliest common ancestor (1809–1899) but fell into two distinct clades (genotypes A and B), possibly representing two independent transmission events from murine-origin CoVs that appear to be a single introduction due to large gaps in the sampling of CoVs in animals. In fact, genotype B was genetically more diverse than all the other sHCoVs. Finally, we found shared amino acid substitutions in multiple proteins along the non-human to sHCoV host-jump branches. The complex evolution of CoVs and their frequent host switches could benefit from continued surveillance of CoVs across non-human hosts. Full article
(This article belongs to the Special Issue Drivers of Evolution of Animal RNA Viruses, Volume II)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>An illustration of the sHCoV genomes, not drawn to scale. The ORFs analyzed in this study are indicated in orange.</p>
Full article ">Figure 2
<p>Maximum clade credibility (MCC) trees inferred from dataset D5 for full genomes (WGS), and the spike, nucleocapsid, membrane, and envelope ORFs, with the branches color-coded by the inferred coronavirus host. The upper panel shows MCC trees from alphacoronaviruses while the lower panel shows MCC trees from betacoronaviruses. Human, camel, and porcine coronavirus clades have been collapsed to increase readability. Human * is a lone human CoV (FJ415324) that clusters with ungulate and canine CoVs. Individual and more detailed MCC trees can be found in <a href="#app1-viruses-14-01551" class="html-app">File S2</a>.</p>
Full article ">Figure 3
<p>Estimates of the evolutionary rate (<b>A</b>) and MRCA age (<b>B</b>) for full genomes and four open reading frames (dataset D5) of the seasonal human coronavirus species. The black horizontal lines in (<b>B</b>) are the dates of first isolation for the 229E (1966), OC43 (1967), NL63 (2004), and HKU1 (2005). The WGS is missing data points for HKU1_all (collective for both genotypes) and HKU1_genotype B as sequences for HKU1_genotype B were all removed in the generation of recombination-free WGS dataset D5.</p>
Full article ">Figure 4
<p>Summarized within and between host/species recombination patterns identified by RDP4, for alphaCoVs (<b>A</b>) and betaCoVs (<b>B</b>). For each sHCoV species, recombining CoVs are shown; non-human and sHCoV (black arrows), within sHCoV species (blue arrows), and between sHCoV species (green arrows). In orange is a lone human CoV (FJ415324) that clusters with ungulate and canine CoVs. Figure generated using Biorender.</p>
Full article ">Figure 5
<p>The number of inferred amino acid changes (AA) associated with the sHCoVs for AA positions in the envelope, membrane, nucleocapsid, and spike proteins from datasets D4 and D5. Panel (<b>A</b>) represents the aggregated AA changes from the alphaCoVs 229E and NL63 while (<b>B</b>) represents the aggregated changes from the betaCoVs OC43 and HKU1. At the top of each plot, the functional domains or regions of the respective proteins are shown; NTD = N-terminal domain, TM = transmembrane domain, CTD = C-terminal domain, RBD = receptor binding domain, LINK = central linker domain, LINK-Dimer = dimerization domain, S1 subunit, S2 subunit, FP = fusion peptide, IFP = internal fusion peptide, HR1 = heptad repeat 1, and HR2 = heptad repeat.</p>
Full article ">
Back to TopTop