[go: up one dir, main page]

Next Issue
Volume 15, April
Previous Issue
Volume 15, February
 
 

Viruses, Volume 15, Issue 3 (March 2023) – 224 articles

Cover Story (view full-size image): Extracellular miRNA (exmiRNA) are mediators of intercellular communication and potential non-invasive biomarkers. Identifying and evaluating exmiRNA carriers will shed light on HIV/SIV pathogenesis and response to treatment. We used our novel tool, PPLC, to isolate the carriers of exmiRNAs, including EVs and ECs from blood plasma of macaques that were infected or not infected with SIV and treated or not treated with delta-9-tetrahydrocannabinol (THC) and combination antiretroviral therapy (cART), followed by small RNA-Seq. We identified alterations in the exmiRNAs common or unique to EVs and ECs. The observed longitudinal changes in the exmiRNAs associated with EVs and ECs reveal key miRNA features of SIV pathogenesis and response to cART or THC. View this paper
  • Issues are regarded as officially published after their release is announced to the table of contents alert mailing list.
  • You may sign up for e-mail alerts to receive table of contents of newly released issues.
  • PDF is the official format for papers published in both, html and pdf forms. To view the papers in pdf format, click on the "PDF Full-text" link, and use the free Adobe Reader to open them.
Order results
Result details
Section
Select all
Export citation of selected articles as:
23 pages, 7013 KiB  
Article
Characterization of Human Norovirus Nonstructural Protein NS1.2 Involved in the Induction of the Filamentous Endoplasmic Reticulum, Enlarged Lipid Droplets, LC3 Recruitment, and Interaction with NTPase and NS4
by Chien-Hui Hung, Ju-Bei Yen, Pey-Jium Chang, Lee-Wen Chen, Tsung-Yu Huang, Wan-Ju Tsai and Yu-Chin Tsai
Viruses 2023, 15(3), 812; https://doi.org/10.3390/v15030812 - 22 Mar 2023
Cited by 2 | Viewed by 2446
Abstract
Human noroviruses (HuNVs) are the leading cause of gastroenteritis worldwide. NS1.2 is critical for HuNV pathogenesis, but the function is still unclear. The GII NS1.2 of HuNVs, unlike GI NS1.2, was localized to the endoplasmic reticulum (ER) and lipid droplets (LDs) and is [...] Read more.
Human noroviruses (HuNVs) are the leading cause of gastroenteritis worldwide. NS1.2 is critical for HuNV pathogenesis, but the function is still unclear. The GII NS1.2 of HuNVs, unlike GI NS1.2, was localized to the endoplasmic reticulum (ER) and lipid droplets (LDs) and is accompanied by a distorted-filamentous ER morphology and aggregated-enlarged LDs. LC3 was recruited to the NS1.2-localized membrane through an autophagy-independent pathway. NS1.2, expressed from a cDNA clone of GII.4 norovirus, formed complexes with NTPase and NS4, which exhibited aggregated vesicle-like structures that were also colocalized with LC3 and LDs. NS1.2 is structurally divided into three domains from the N terminus: an inherently disordered region (IDR), a region that contains a putative hydrolase with the H-box/NC catalytic center (H-box/NC), and a C-terminal 251–330 a.a. region containing membrane-targeting domain. All three functional domains of NS1.2 were required for the induction of the filamentous ER. The IDR was essential for LC3 recruitment by NS1.2. Both the H-Box/NC and membrane-targeting domains are required for the induction of aggregated-enlarged LDs, NS1.2 self-assembly, and interaction with NTPase. The membrane-targeting domain was sufficient to interact with NS4. The study characterized the NS1.2 domain required for membrane targeting and protein–protein interactions, which are crucial for forming a viral replication complex. Full article
(This article belongs to the Special Issue Viral Gastroenteritis 2022)
Show Figures

Figure 1

Figure 1
<p>The domain of NS1.2 required for ER targeting and reorganization. (<b>A</b>) The schematic diagram shows the name and structure of Flag-tagged NS1.2 and its deletion mutants. (<b>B</b>) A7 cells were transfected with plasmids expressing Flag-tagged NS1.2, which is denoted as F-NS1.2, or its derivative mutants. Cells were fixed at 24h after transfection. Indirect immunofluorescence staining was performed using anti-Flag and anti-PDI antibodies and shown as green and red color, respectively. (<b>C</b>) The amino acid sequences between 261 and 278 of NS1.2 (1–280) and substituted mutant NS1.2 (1–280LF/ED) are shown on the top of the panel. The hydrophobic and acidic amino acids were highlighted in yellow and red, respectively. The helical wheel plot between 261 and 278 of NS1.2 predicted from computational analysis (accessed on 13 July 2021, <a href="https://heliquest.ipmc.cnrs.fr/cgi-bin/ComputParams.py" target="_blank">https://heliquest.ipmc.cnrs.fr/cgi-bin/ComputParams.py</a>) was shown. The hydrophobic and basic residues were shown in yellow and blue, respectively. The acidic residues are shown in red. The serine and threonine residues are shown in purple. The small amino acids are shown in grey. Proline is shown in green. The arrow represents the direction and magnitude of the hydrophobic moment. A7 cell were transfected with F-tagged NS1.2 (1–280), denoted as F-1–280, and substituted mutant, F-1–280(LF/ED). Cells were fixed at 24 h after transfection. Indirect immunofluorescence staining was performed using anti-Flag and anti-PDI antibodies. (<b>D</b>) A7 cells were transfected with plasmids expressing GFP-tagged NS1.2, (251–330), (251–290), or (280–330), which are denoted as GFP-NS1.2, GFP-251–330, GFP-250–290 and GFP-280–330, respectively. Cells were fixed at 24h after transfection, and indirect immunofluorescence staining was performed using anti-PDI or anti-GM130 antibodies, as indicated and shown in red color. LD was stained with LipidTox red neutral lipid dye. (<b>E</b>) The A7, HeLa, and HEK293T cells expressed Flag-tagged NS1.2 (117–330) were examined by immunofluorescent analysis using anti-Flag and anti-GM130 antibodies. 4′, 6-diamidino-2-pheylindole (DAPI) staining revealed the nucleus. Scale bar: 10 μm.</p>
Full article ">Figure 2
<p>The domain of NS1.2 is required for LDs targeting and induced enlarged LDs. (<b>A</b>) The schematic diagram shows the name and structure of Flag-tagged NS1.2 and its deletion mutants. (<b>B</b>) A7 cells were transfected with plasmids expressing Flag-tagged NS1.2 and its deletion mutants. Cells were fixed at 24h after transfection. Indirect immunofluorescence staining was performed using anti-Flag and counterstain LD with LipidTox red neutral lipid dye. The transfected cells were marked with an asterisk; and the control cells were marked with a white arrowhead. The dashed boxes in the merged images were enlarged and shown on the right. 4′, 6-diamidino-2-pheylindole (DAPI) staining revealed the nucleus. Scale bar: 10 μm. (<b>C</b>) Quantification of the number and average size of LDs in panel B. The diameter of LD is represented as the size of LD. Each dot represents the quantitated data from a single cell (<span class="html-italic">n</span> &gt; 20). The results were analyzed statistically through a <span class="html-italic">t</span>-test. ** <span class="html-italic">p</span> &lt; 0.01; NS, not statistically significant. C: control cells that did not express F-tagged NS1.2 and its deletion mutants.</p>
Full article ">Figure 3
<p>LC3 is recruited to NS1.2 localized ER filamentous membrane. (<b>A</b>) A7 cells, (<b>B</b>) MEF Atg5+/+ and MEF Atg5−/−, (<b>C</b>) AD293(GFP-LC3) were transfected with control plasmid ((<b>C</b>), 1st row) or plasmid expressing Flag-tagged NS1.2 or NS1.2 (117–330), denoted as F-NS1.2 and F-117–330, as indicated. Cells were fixed at 24 h after transfection. Indirect immunofluorescence staining was performed using anti-Flag and anti-PDI ((<b>A</b>), 1st row) or anti-Flag and anti-LC3 ((<b>A</b>), 2nd and 3rd row and (<b>B</b>)) or anti-Flag (<b>C</b>) antibodies. *: cell without expressing F-NS1.2. 4′, 6-diamidino-2-pheylindole (DAPI) staining revealed the nucleus. Scale bar: 10 μm.</p>
Full article ">Figure 4
<p>The domain involved in the dimerization/oligomerization of NS.1.2. (<b>A</b>) The schematic diagram shows the structure of NS1.2 and deletion mutants of NS1.2 and the summary of coimmunoprecipitation results. The “+”, “++”, and “+++” represent the relative level of Myc-tagged protein coimmunoprecipitated with Flag-tagged protein by anti-Flag antibody in each experiment. (<b>B</b>) HEK293T cells were cotransfected with plasmids expressing GFP or GFP-NS1.2 and Flag-NS1.2. Flagtagged NS1.2 is denoted as F-NS1.2. Proteins in the lysates were immunoprecipitated (IP) at 48 h after transfection using GFP-Trap, or anti-Flag conjugated magnetic beads and analyzed by immunoblotting using anti-GFP or anti-Flag antibodies as indicated. (<b>C</b>–<b>E</b>) HEK293T cells were cotransfected with plasmids expression of Myc-tagged NS1.2 (<b>C</b>), Myc-tagged H-box/NC domain (117–250) (<b>D</b>), or Myc-tagged membrane-targeting domain (233–330) (<b>E</b>) with a control plasmid, Flag-NS1.2 or NS1.2 deletion mutants as indicated. The control plasmid is denoted as C. Proteins in the lysates were immunoprecipitated (IP) at 48 h after transfection using anti-Myc or anti-Flag conjugated magnetic beads and analyzed by immunoblotting using anti-Myc or anti-Flag antibodies as indicated. “*”: suggested dimer form of F-117–250.</p>
Full article ">Figure 5
<p>NS1.2 interacts and colocalizes with NS4 and NTPase. (<b>A</b>) 293T cell were co-transfected with plasmids expressing Flag-NS1.2 and Myc-tagged NS1.2, or NTPase, or NS4, or VPg, or Protease, or RdRP. Proteins in the lysates were immunoprecipitated (IP) 48 h after transfection using anti-Flag or anti-Myc antibodies and analyzed by immunoblotting using anti-Flag or anti-Myc antibodies as indicated. (<b>B</b>) Upper panel: The schematic figure of the Flag-NV plasmid: a cDNA expression plasmid of GII.4 HuNV. Lower panel: 293T cells were transfected with a control plasmid (C) or plasmid Flag-NV. The cell lysate was harvested at 48 h after transfection and examined by immunoblotting using antibodies as indicated. The arrows indicate the mature NS1.2, NTPase and NS4. *: nonspecific bands; arrowhead: truncated form of NS1.2. (<b>C</b>) 293T cells were transfected with plasmids p-F-NV for 48 h. Proteins in the lysates were immunoprecipitated (IP) using anti-Flag antibodies or control IgG. Proteins that bound on beads were analyzed by immunoblotting using antibodies as indicated. (<b>D</b>,<b>E</b>) A7 cells were transfected with plasmid Flag-NV. Cells were fixed at 48 h after transfection. Indirect immunofluorescence staining was performed using anti-Flag (NS1.2), anti-NTPase (<b>a1</b>,<b>a2</b>) or anti-NS4 (<b>b1</b>,<b>b2</b>), anti-LC3 (<b>c1</b>,<b>c2</b>) antibodies or counterstained LD with LipidTox red neutral lipid dye (<b>d1</b>,<b>d2</b>). (<b>F</b>) AD293(GFP-LC3) cells were transfected with control plasmids (1st row) or plasmids expressing Flag-tagged NS1.2 (2nd row) and were fixed at 24 h after transfection. Indirect immunofluorescence staining was performed using anti-Flag antibodies. *: cells expressing polyprotein of GII.4 HuNoV. DAPI staining revealed the nucleus.</p>
Full article ">Figure 6
<p>The domain of NS.1.2 interacts with NS4. (<b>A</b>) The schematic diagram shows the deletion mutants of NS1.2 and the summary of coimmunoprecipitation results. (<b>B</b>,<b>C</b>) 293T cells were cotransfected with plasmids expression Myc-tagged NS4 with control plasmid (C), Flag-tagged NS1.2, or NS1.2 deletion mutants as indicated. Proteins in the lysates were immunoprecipitated (IP) at 48 h after transfection using anti-Flag or anti-Myc conjugated magnetic beads and analyzed by immunoblotting using anti-Flag or anti-Myc antibodies. (<b>D</b>) 293T cells were cotransfected with plasmids expressing Myc-tagged NS4 and Flag-tagged NS1.2 or its deletion mutants as described. Indirect immunofluorescence staining was performed at 48 h after transfection using anti-Flag and anti-Myc antibodies. Cells coexpressing NS1.2 and NTPase were marked with white arrow; cells expressing NS4 were marked with an asterisk. 4′, 6-diamidino-2-pheylindole (DAPI) staining revealed the nucleus.</p>
Full article ">Figure 7
<p>The domain of NS1.2 interacts with NTPase. (<b>A</b>) The schematic diagram shows the deletion mutants of NS1.2 and the summary of coimmunoprecipitation results. The “+”, “++”, and “+++” represent the relative level of protein coimmunoprecipitated with Flag-NS1.2 or NTPase-Myc by anti-Flag or anti-Myc antibody, respectively. (<b>B</b>) 293T cells were cotransfected with plasmids expression Myc-tagged NTPase with control plasmid (C), Flag-tagged NS1.2 or NS1.2 deletion mutants as indicated. Proteins in the lysates were immunoprecipitated (IP) at 48 h after transfection by anti-Flag (middle panel) or anti-Myc (right panel) conjugated magnetic beads and analyzed by immunoblotting with anti-Flag or anti-Myc antibodies. The input lanes were loaded with 4% of lysate. White arrow: suggested dimer form of Flag-tagged protein. Arrow: NTPase. *: nonspecific band. (<b>C</b>) 293T cells were cotransfected with plasmids expressing Myc-tagged NS4 and Flag-tagged NS1.2 or its deletion mutants as indicated. Indirect immunofluorescence staining was performed 48 h after transfection using anti-Flag and anti-Myc antibodies. Cells coexpressing NS1.2 and NTPase were marked with white arrow; cells expressing NTPase alone were marked with an asterisk. 4′, 6-diamidino-2-pheylindole (DAPI) staining revealed the nucleus.</p>
Full article ">Figure 8
<p>Summary of the structural domains of HuNoV NS1.2 involved in specific functions. The schematic diagram on top of the figure shows the structural domain of NS1.2. The domain required for various functions of NS1.2 determined in this study was shown in the button of the schematic diagram.</p>
Full article ">
10 pages, 855 KiB  
Article
Real-World Experience of the Comparative Effectiveness and Safety of Molnupiravir and Nirmatrelvir/Ritonavir in High-Risk Patients with COVID-19 in a Community Setting
by Yoshikazu Mutoh, Takumi Umemura, Takeshi Nishikawa, Kaho Kondo, Yuta Nishina, Kazuaki Soejima, Yoichiro Noguchi, Tomohiro Bando, Sho Ota, Tatsuki Shimahara, Shuko Hirota, Satoshi Hagimoto, Reoto Takei, Jun Fukihara, Hajime Sasano, Yasuhiko Yamano, Toshiki Yokoyama, Kensuke Kataoka, Toshiaki Matsuda, Tomoki Kimura, Toshihiko Ichihara and Yasuhiro Kondohadd Show full author list remove Hide full author list
Viruses 2023, 15(3), 811; https://doi.org/10.3390/v15030811 - 22 Mar 2023
Cited by 10 | Viewed by 3608
Abstract
Molnupiravir (MOV) and nirmatrelvir/ritonavir (NMV/r) are efficacious oral antiviral agents for patients with the 2019 coronavirus (COVID-19). However, little is known about their effectiveness in older adults and those at high risk of disease progression. This retrospective single-center observational study assessed and compared [...] Read more.
Molnupiravir (MOV) and nirmatrelvir/ritonavir (NMV/r) are efficacious oral antiviral agents for patients with the 2019 coronavirus (COVID-19). However, little is known about their effectiveness in older adults and those at high risk of disease progression. This retrospective single-center observational study assessed and compared the outcomes of COVID-19 treated with MOV and NMV/r in a real-world community setting. We included patients with confirmed COVID-19 combined with one or more risk factors for disease progression from June to October 2022. Of 283 patients, 79.9% received MOV and 20.1% NMV/r. The mean patient age was 71.7 years, 56.5% were men, and 71.7% had received ≥3 doses of vaccine. COVID-19-related hospitalization (2.8% and 3.5%, respectively; p = 0.978) or death (0.4% and 3.5%, respectively; p = 0.104) did not differ significantly between the MOV and NMV/r groups. The incidence of adverse events was 2.7% and 5.3%, and the incidence of treatment discontinuation was 2.7% and 5.3% in the MOV and NMV/r groups, respectively. The real-world effectiveness of MOV and NMV/r was similar among older adults and those at high risk of disease progression. The incidence of hospitalization or death was low. Full article
(This article belongs to the Collection Coronaviruses)
Show Figures

Figure 1

Figure 1
<p>Cumulative hospitalization rate according to the drug type.</p>
Full article ">Figure 2
<p>Forest plots of risk factors for hospitalization. Odds ratios were calculated using multivariable logistic regression. C.I., confidence interval.</p>
Full article ">
10 pages, 386 KiB  
Article
Acute Bronchitis and Bronchiolitis Infection in Children with Asthma and Allergic Rhinitis: A Retrospective Cohort Study Based on 5,027,486 Children in Taiwan
by Fung-Chang Sung, Chang-Ching Wei, Chih-Hsin Muo, Shan P. Tsai, Chao W. Chen, Dennis P. H. Hsieh, Pei-Chun Chen and Chung-Yen Lu
Viruses 2023, 15(3), 810; https://doi.org/10.3390/v15030810 - 22 Mar 2023
Cited by 1 | Viewed by 2729
Abstract
This study evaluated the risks of childhood acute bronchitis and bronchiolitis (CABs) for children with asthma or allergic rhinitis (AR). Using insurance claims data of Taiwan, we identified, from children of ≤12 years old in 2000–2016, cohorts with and without asthma (N = [...] Read more.
This study evaluated the risks of childhood acute bronchitis and bronchiolitis (CABs) for children with asthma or allergic rhinitis (AR). Using insurance claims data of Taiwan, we identified, from children of ≤12 years old in 2000–2016, cohorts with and without asthma (N = 192,126, each) and cohorts with and without AR (N = 1,062,903, each) matched by sex and age. By the end of 2016, the asthma cohort had the highest bronchitis incidence, AR and non-asthma cohorts followed, and the lowest in the non-AR cohort (525.1, 322.4, 236.0 and 169.9 per 1000 person-years, respectively). The Cox method estimated adjusted hazard ratios (aHRs) of bronchitis were 1.82 (95% confidence interval (CI), 1.80–1.83) for the asthma cohort and 1.68 (95% CI, 1.68–1.69) for the AR cohort, relative to the respective comparisons. The bronchiolitis incidence rates for these cohorts were 42.7, 29.5, 28.5 and 20.1 per 1000 person-years, respectively. The aHRs of bronchiolitis were 1.50 (95% CI, 1.48–1.52) for the asthma cohort and 1.46 (95% CI, 1.45–1.47) for the AR cohort relative to their comparisons. The CABs incidence rates decreased substantially with increasing age, but were relatively similar for boys and girls. In conclusion, children with asthma are more likely to develop CABs than are children with AR. Full article
(This article belongs to the Special Issue Pediatric Respiratory Viral Infection)
Show Figures

Figure 1

Figure 1
<p>Flow chart for selecting cohorts with and without asthma and cohorts with and without allergic rhinitis from children of ≤12 years-old in 2000–2016.</p>
Full article ">
10 pages, 1183 KiB  
Communication
Molecular Screening for High-Risk Human Papillomaviruses in Patients with Periodontitis
by Kalina Shishkova, Raina Gergova, Elena Tasheva, Stoyan Shishkov and Ivo Sirakov
Viruses 2023, 15(3), 809; https://doi.org/10.3390/v15030809 - 22 Mar 2023
Cited by 2 | Viewed by 1687
Abstract
Members of the Papillomaviridae family account for 27.9–30% of all infectious agents associated with human cancer. The aim of our study was to investigate the presence of high-risk HPV (human papilloma virus) genotypes in patients with periodontitis and a pronounced clinical picture. To [...] Read more.
Members of the Papillomaviridae family account for 27.9–30% of all infectious agents associated with human cancer. The aim of our study was to investigate the presence of high-risk HPV (human papilloma virus) genotypes in patients with periodontitis and a pronounced clinical picture. To achieve this goal, after proving the bacterial etiology of periodontitis, the samples positive for bacteria were examined for the presence of HPV. The genotype of HPV is also determined in samples with the presence of the virus proven by PCR (polymerase chain reaction). All positive tests for bacteria associated with the development of periodontitis indicated the presence of HPV. There was a statistically significant difference in HPV positive results between the periodontitis positive target group and the control group. The higher presence of high-risk HPV genotypes in the target group, which was also positive for the presence of periodontitis-causing bacteria, has been proven. A statistically significant relationship was established between the presence of periodontitis-causing bacteria and high-risk strains of HPV. The most common HPV genotype that tests positive for bacteria associated with the development of periodontitis is HPV58. Full article
(This article belongs to the Special Issue Opportunistic Viral Infections)
Show Figures

Figure 1

Figure 1
<p>PCR with consensus primers GPE6/5B/6B and MY09/11of DNA materials from samples positive for periodontitis bacteria. M, 100 bp DNA marker; (<b>Upper row</b>) PCR performed with primers MY09/11. Reported fragment of about 450 bp; (<b>Bottom row</b>) PCR performed with primers GPE6/5B/6B. Reported fragment of about 700 bp; 10, negative control with DNA negative for HPV; M, 1 kb DNA marker. Samples from the target group were applied to the remaining starts.</p>
Full article ">Figure 2
<p>Presence of high-risk HIV genotypes in periodontitis bacteria-positive patients.</p>
Full article ">Figure 3
<p>Presence of high-risk HPV genotypes in coronal margin samples of target and control groups.</p>
Full article ">
13 pages, 3870 KiB  
Article
Establishment of an In Vitro Model of Pseudorabies Virus Latency and Reactivation and Identification of Key Viral Latency-Associated Genes
by Li Pan, Mingzhi Li, Xinyu Zhang, Yu Xia, Assad Moon Mian, Hongxia Wu, Yuan Sun and Hua-Ji Qiu
Viruses 2023, 15(3), 808; https://doi.org/10.3390/v15030808 - 22 Mar 2023
Cited by 2 | Viewed by 2117
Abstract
Alphaherpesviruses infect humans and most animals. They can cause severe morbidity and mortality. The pseudorabies virus (PRV) is a neurotropic alphaherpesvirus that can infect most mammals. The PRV persists in the host by establishing a latent infection, and stressful stimuli can induce the [...] Read more.
Alphaherpesviruses infect humans and most animals. They can cause severe morbidity and mortality. The pseudorabies virus (PRV) is a neurotropic alphaherpesvirus that can infect most mammals. The PRV persists in the host by establishing a latent infection, and stressful stimuli can induce the latent viruses to reactivate and cause recurrent diseases. The current strategies of antiviral drug therapy and vaccine immunization are ineffective in eliminating these viruses from the infected host. Moreover, overspecialized and complex models are also a major obstacle to the elucidation of the mechanisms involved in the latency and reactivation of the PRV. Here, we present a streamlined model of the latent infection and reactivation of the PRV. A latent infection established in N2a cells infected with the PRV at a low multiplicity of infection (MOI) and maintained at 42 °C. The latent PRV was reactivated when the infected cells were transferred to 37 °C for 12 to 72 h. When the above process was repeated with a UL54-deleted PRV mutant, it was observed that the UL54 deletion did not affect viral latency. However, viral reactivation was limited and delayed. This study establishes a powerful and streamlined model to simulate PRV latency and reveals the potential role of temperature in PRV reactivation and disease. Meanwhile, the key role of the early gene UL54 in the latency and reactivation of PRV was initially elucidated. Full article
(This article belongs to the Special Issue Pseudorabies Virus, Volume II)
Show Figures

Figure 1

Figure 1
<p>Generation and characterization of rPRV-EGFP. (<b>A</b>) Schematic diagram of the construction of the recombinant virus. Using the PRV genomic fosmid library, the EGFP sequence was inserted nondestructively downstream of the <span class="html-italic">US9</span> gene in the PRV genome. The relative position of the EGFP sequence in the WT-PRV genome is shown on top of the green box. (<b>B</b>) PCR amplification of the <span class="html-italic">EGFP</span> and <span class="html-italic">gB</span> genes from the genomes of rPRV-EGFP and WT-PRV. (<b>C</b>) The green fluorescence and cytopathic effects (CPEs) in the PK-15 cells infected with rPRV-EGFP and WT-PRV at 48 h postinfection (hpi). (<b>D</b>) Plaques of rPRV-EGFP and WT-PRV in the PK-15 cells. ns: not significant (<span class="html-italic">p</span> ≥ 0.05). The diameters of plaques were averaged for three independent experiments. ns: not significant. (<b>E</b>) Transmission electron microscopy photographs of rPRV-EGFP and WT-PRV. Scale bar = 200 nm. (<b>F</b>) One-step growth curves of rPRV-EGFP and WT-PRV in PK-15 cells.</p>
Full article ">Figure 2
<p>Screening for appropriate cell types and optimal infectious dose for in vitro PRV latent infection models. N2a, BHK-21, and PK-15 cells were seeded into 35 mm dishes at 5 × 10<sup>5</sup> per dish and were cultured with complete medium. rPRV-EGFP was added to the three cell types at an MOI of 1, 0.1, 0.01, or 0.001. The infected cells were incubated continuously at 37 or 42 °C for 72 h. Fluorescence was monitored at 72 h postinfection.</p>
Full article ">Figure 3
<p>Optimal incubation time for screening and detection of latency indicators. (<b>A</b>) Optimal incubation time screening. N2a cells were infected with rPRV-EGFP (MOI = 0.01) and were incubated at 37 °C for 2, 2.5, 3, 3.5, 4, 4.5, 5, and 5.5 h. The cells were then transferred to 42 °C for 120 h, and EGFP fluorescence was monitored. (<b>B</b>) Identification of the optimal conditions for latent PRV infection. N2a cells were infected at an MOI of 0.01, were incubated at 37 °C for 2 h, then transferred to 42 °C for 120 h. (<b>C</b>) Latent and lytic viral DNA detection. N2a cells were infected with rPRV-EGFP at an MOI of 0.01 and were incubated at 37 °C for 2 h. After rPRV-EGFP was cultured at 42 °C for 120 h, the samples were then collected, and viral DNA was extracted; additionally, the <span class="html-italic">gB</span> gene was detected through real-time PCR to quantify genomic DNA. (<b>D</b>) Quantification of transcripts of latent viral genes. After rPRV-EGFP was cultured at 42 °C for 120 h, the samples were then collected, and the mRNAs of <span class="html-italic">IE180</span>, <span class="html-italic">EP0</span>, <span class="html-italic">LAT</span>, and <span class="html-italic">gB</span> were detected through RT-qPCR. (<b>E</b>) Detection of late viral protein expression levels at two culture temperatures. N2a cells were infected with different doses of rPRV-EGFP (MOI = 1, 0.1, or 0.01) and were incubated at 37 or 42 °C for 120 h, respectively. At 120 h postinfection (hpi), the cells were harvested, and samples were prepared for Western blotting. The homemade mouse anti-gB and -gD monoclonal antibodies were used as primary antibodies, and goat antimouse FITC antibodies were used as secondary antibodies. Bars represent the means ± SDs of three independent experiments. ns: not significant (<span class="html-italic">p</span> ≥ 0.05). * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001, and **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 4
<p>Reactivation of latent rPRV-EGFP. (<b>A</b>) Fluorescence spots of rPRV-EGFP during reactivation. N2a cells were infected with rPRV-EGFP (MOI = 0.01), were incubated at 42 °C for 120 h, and were then transferred to 37 °C for additional incubation. EGFP fluorescence was monitored at 12, 24, 48, and 72 h postreactivation (hpr). (<b>B</b>) Detection of viral genomic DNA during reactivation. The viral DNA was extracted, and the <span class="html-italic">gB</span> gene was detected at 0, 12, 24, 48, and 72 hpr. (<b>C</b>) Identification of expression levels of viral late proteins following rPRV-EGFP reactivation. N2a cells were infected with rPRV-EGFP (MOI = 0.01), were incubated at 42 °C for 120 h, and were then transferred to 37 °C. At 0, 12, 24, 48, and 72 hpr, the cells were harvested, and samples were prepared for Western blotting. The homemade mouse anti-gB and -gD monoclonal antibodies were used as primary antibodies, and goat antimouse FITC antibodies were used as secondary antibodies. Bars represent the means ± SDs of three independent experiments. ns: not significant (<span class="html-italic">p</span> ≥ 0.05). * <span class="html-italic">p</span> &lt; 0.05, *** <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>Generation and characterization of rPRV-ΔUL54-EGFP. (<b>A</b>) Schematic diagram of the construction of rPRV-ΔUL54-EGFP. Using the PRV genomic fosmid library, the <span class="html-italic">UL54</span> deletion mutant was constructed using rPRV-EGFP as the parental strain. The relative position of the <span class="html-italic">UL54</span> gene in the WT-PRV genome is shown on top of the red box. (<b>B</b>) PCR amplification of the <span class="html-italic">UL54</span>, <span class="html-italic">EGFP</span>, and <span class="html-italic">gB</span> genes from the genomes of rPRV-ΔUL54-EGFP and WT-PRV. (<b>C</b>) The green fluorescence and cytopathic effects (CPEs) in the PK-15 cells infected with rPRV-ΔUL54-EGFP and WT-PRV at 60 h postinfection (hpi). (<b>D</b>) Plaques of rPRV-ΔUL54-EGFP and WT-PRV in the PK-15 cells. The diameters of plaques were averaged for three independent experiments. (<b>E</b>) Transmission electron microscopy imagingof rPRV-ΔUL54-EGFP and WT-PRV. Scale bar = 200 nm. (<b>F</b>) Multistep growth curves of rPRV-ΔUL54-EGFP and WT-PRV in PK-15 cells. PK-15 cells were infected with rPRV-ΔUL54-EGFP and WT-PRV (MOI = 0.01). Bars represent the means ± SDs of three independent experiments. ns: not significant (<span class="html-italic">p</span> ≥ 0.05). ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 6
<p>Latent infection and reactivation of rPRV-ΔUL54-EGFP. Based on the established model, rPRV-ΔUL54-EGFP was tested. (<b>A</b>) Identification of the effect of UL54 deletion on latent infection. (<b>B</b>) Latent and lytic viral DNA copy detection. N2a cells were infected with rPRV-ΔUL54-EGFP at an MOI of 0.01 and were incubated at 37 °C for 2 h. After rPRV-ΔUL54-EGFP was cultured at 42 °C for 120 h, the samples were then collected, and viral DNA was extracted; additionally, the <span class="html-italic">gB</span> gene was detected through real-time PCR to quantify genomic DNA. (<b>C</b>) Quantification of the transcripts of the latency-associated viral genes. After rPRV-ΔUL54-EGFP was cultured at 42 °C for 120 h, the samples were then collected, and total RNA was extracted; additionally, the mRNA of the <span class="html-italic">IE180</span>, <span class="html-italic">EP0</span>, <span class="html-italic">LAT</span>, and <span class="html-italic">gB</span> genes was detected through RT-qPCR. (<b>D</b>) Inability of latent rPRV-UL54-EGFP to express lytic viral proteins. (<b>E</b>) Fluorescence spots of rPRV-EGFP and rPRV-ΔUL54-EGFP during reactivation. N2a cells were infected with rPRV-EGFP and rPRV-ΔUL54-EGFP (MOI = 0.01), were incubated at 42 °C for 120 h, and were then transferred to 37 °C for culture. EGFP fluorescence was monitored at 12, 24, 48, and 72 h postreactivation (hpr). (<b>F</b>) Detection of viral genomic DNA during reactivation. N2a cells were infected with rPRV-ΔUL54-EGFP and rPRV-ΔUL54-EGFP (MOI = 0.01). The viral DNA was extracted, and the <span class="html-italic">gB</span> gene was quantified at 0, 12, 24, 48, and 72 hpr. (<b>G</b>) Identification of expression levels of viral envelope proteins following rPRV-ΔUL54-EGFP reactivation. Bars represent the means ± SDs of three independent experiments. ns: not significant (<span class="html-italic">p</span>  ≥ 0.05). * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001, and **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">
12 pages, 1904 KiB  
Article
Phage vs. Phage: Direct Selections of Sandwich Binding Pairs
by Emily C. Sanders, Alicia M. Santos, Eugene K. Nguyen, Aidan A. Gelston, Sudipta Majumdar and Gregory A. Weiss
Viruses 2023, 15(3), 807; https://doi.org/10.3390/v15030807 - 22 Mar 2023
Cited by 2 | Viewed by 2926
Abstract
The sandwich format immunoassay is generally more sensitive and specific than more common assay formats, including direct, indirect, or competitive. A sandwich assay, however, requires two receptors to bind non-competitively to the target analyte. Typically, pairs of antibodies (Abs) or antibody fragments (Fabs) [...] Read more.
The sandwich format immunoassay is generally more sensitive and specific than more common assay formats, including direct, indirect, or competitive. A sandwich assay, however, requires two receptors to bind non-competitively to the target analyte. Typically, pairs of antibodies (Abs) or antibody fragments (Fabs) that are capable of forming a sandwiching with the target are identified through a slow, guess-and-check method with panels of candidate binding partners. Additionally, sandwich assays that are reliant on commercial antibodies can suffer from changes to reagent quality outside the researchers’ control. This report presents a reimagined and simplified phage display selection protocol that directly identifies sandwich binding peptides and Fabs. The approach yielded two sandwich pairs, one peptide–peptide and one Fab–peptide sandwich for the cancer and Parkinson’s disease biomarker DJ-1. Requiring just a few weeks to identify, the sandwich pairs delivered apparent affinity that is comparable to other commercial peptide and antibody sandwiches. The results reported here could expand the availability of sandwich binding partners for a wide range of clinical biomarker assays. Full article
(This article belongs to the Special Issue Phage Display in Cancer Research)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Phage vs. phage selection strategy. (1) Biotinylated phage displaying the known binding partner are adsorbed to a microtiter plate and exposed to the target. Next, the phage library is added and allowed to bind. (2) Nonbinding phages are removed by washing. (3) Sandwich assemblies are eluted from the microtiter plate by treatment with low pH solution and sonication. (4) In the eluate, biotinylated phage display with the known binding partner can removed from the solution through binding to magnetic streptavidin beads. (5) The sandwich forming pair of phages are used to infect <span class="html-italic">E. coli</span> and amplified. (6) The process is repeated with the amplified selectants as the phage library for the next round of selection. (7) After three to four rounds of selection, the selectants are identified by DNA sequencing.</p>
Full article ">Figure 2
<p>Efficient phage biotinylation and its effects on target binding. (<b>A</b>) Biotinylation of phages (bDPep1Φ) was confirmed via direct detection ELISA with peroxidase-conjugated streptavidin (Strep-HRP) assay. (<b>B</b>) An indirect phage ELISA with DJ-1 verified that both biotinylated and non-biotinylated page (DPep1Φ) could bind the target antigen despite biotinylation. ELISA data were fit to a four-parameter logistic curve. Error bars indicate standard deviation (<span class="html-italic">n</span> = 3); each data point includes error bars, although most are quite small. HRP = horseradish peroxidase.</p>
Full article ">Figure 3
<p>Sensitivity and specificity determination of sandwich peptide selectants. (<b>A</b>) A dose-dependent direct phage assay with sandwich peptide selectants binding to DJ-1. Sub-micromolar phage EC<sub>50</sub> values were observed. NO = not observed. (<b>B</b>) An indirect phage assay with Hb, HEWL, HSA, <span class="html-italic">E. coli</span> lysate, and BSA determined the specificity of the sandwich selectants. Data are normalized to the DJ-1 signal. (<b>C</b>) Streptavidin-bound DPep1 peptide assay for sandwiching the target DJ-1 with DPep3Φ. The resultant dose-dependent response yielded an EC<sub>50</sub> for binding to DJ-1 of 249 pM. (<b>D</b>) The DPep1-DPep3Φ sandwich interaction also demonstrated the interaction’s improved selectivity for DJ-1. ANOVA with Sidak’s multiple comparisons yielded <span class="html-italic">p</span>-values of &lt;0.01 (**) and &lt;0.001 (***). ELISA data were fit to a four-parameter logistic curve. Error bars, included for every data point, depict standard error of the mean (SEM) (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 4
<p>Sensitivity and specificity determination of DFab1. (<b>A</b>) A dose-dependent direct phage assay with DFab1Φ and NegΦ. Data were fit to a four-parameter logistic curve and yielded picomolar EC<sub>50</sub> values for the DFab1Φ interaction with DJ-1. Error bars represent SEM (<span class="html-italic">n</span> = 3). NO = not observed. (<b>B</b>) An indirect phage assay with DJ-1, Hb, HEWL, HSA, <span class="html-italic">E. coli</span> lysate, and BSA determined the specificity of the sandwich selectants. ANOVA with Tukey’s multiple comparisons yielded <span class="html-italic">p</span>-values of &lt;0.0001 (****). (<b>C</b>) A similar dose-dependent response with an EC<sub>50</sub> of 1.53 μM was observed in an identical sandwich ELISA with DFab1Φ. (<b>D</b>) The DPep1-DFab1Φ also demonstrated strong selectivity for DJ-1. ELISA data were fit to a four-parameter logistic curve. Data are normalized to DFab1Φ binding to DJ-1. Error is represented as SEM (<span class="html-italic">n</span> = 3). NO = not observed. ANOVA with Tukey’s multiple comparisons yielded <span class="html-italic">p</span>-values of &lt;0.01 (**), &lt;0.001 (***), &lt;0.0001 (****).</p>
Full article ">
18 pages, 872 KiB  
Review
Interleukins, Chemokines, and Tumor Necrosis Factor Superfamily Ligands in the Pathogenesis of West Nile Virus Infection
by Emna Benzarti, Kristy O. Murray and Shannon E. Ronca
Viruses 2023, 15(3), 806; https://doi.org/10.3390/v15030806 - 22 Mar 2023
Cited by 6 | Viewed by 2667
Abstract
West Nile virus (WNV) is a mosquito-borne pathogen that can lead to encephalitis and death in susceptible hosts. Cytokines play a critical role in inflammation and immunity in response to WNV infection. Murine models provide evidence that some cytokines offer protection against acute [...] Read more.
West Nile virus (WNV) is a mosquito-borne pathogen that can lead to encephalitis and death in susceptible hosts. Cytokines play a critical role in inflammation and immunity in response to WNV infection. Murine models provide evidence that some cytokines offer protection against acute WNV infection and assist with viral clearance, while others play a multifaceted role WNV neuropathogenesis and immune-mediated tissue damage. This article aims to provide an up-to-date review of cytokine expression patterns in human and experimental animal models of WNV infections. Here, we outline the interleukins, chemokines, and tumor necrosis factor superfamily ligands associated with WNV infection and pathogenesis and describe the complex roles they play in mediating both protection and pathology of the central nervous system during or after virus clearance. By understanding of the role of these cytokines during WNV neuroinvasive infection, we can develop treatment options aimed at modulating these immune molecules in order to reduce neuroinflammation and improve patient outcomes. Full article
(This article belongs to the Special Issue Innate Immunity to Virus Infection 2023)
Show Figures

Figure 1

Figure 1
<p>Cellular targets of West Nile virus and corresponding cytokine response in mammals. The illustration was created in Biorender.com. Abbreviations: BAFF: B-cell activating factor; FasL: Fas ligand; TNF-α: tumor necrosis factor -α, TRAIL: TNF-related apoptosis-inducing ligand.</p>
Full article ">
9 pages, 271 KiB  
Opinion
Puumala Hantavirus Infections Show Extensive Variation in Clinical Outcome
by Antti Vaheri, Teemu Smura, Hanna Vauhkonen, Jussi Hepojoki, Tarja Sironen, Tomas Strandin, Johanna Tietäväinen, Tuula Outinen, Satu Mäkelä, Ilkka Pörsti and Jukka Mustonen
Viruses 2023, 15(3), 805; https://doi.org/10.3390/v15030805 - 22 Mar 2023
Cited by 5 | Viewed by 1979
Abstract
The clinical outcome of Puumala hantavirus (PUUV) infection shows extensive variation, ranging from inapparent subclinical infection (70–80%) to severe hemorrhagic fever with renal syndrome (HFRS), with about 0.1% of cases being fatal. Most hospitalized patients experience acute kidney injury (AKI), histologically known as [...] Read more.
The clinical outcome of Puumala hantavirus (PUUV) infection shows extensive variation, ranging from inapparent subclinical infection (70–80%) to severe hemorrhagic fever with renal syndrome (HFRS), with about 0.1% of cases being fatal. Most hospitalized patients experience acute kidney injury (AKI), histologically known as acute hemorrhagic tubulointerstitial nephritis. Why this variation? There is no evidence that there would be more virulent and less virulent variants infecting humans, although this has not been extensively studied. Individuals with the human leukocyte antigen (HLA) alleles B*08 and DRB1*0301 are likely to have a severe form of the PUUV infection, and those with B*27 are likely to have a benign clinical course. Other genetic factors, related to the tumor necrosis factor (TNF) gene and the C4A component of the complement system, may be involved. Various autoimmune phenomena and Epstein-Barr virus infection are associated with PUUV infection, but hantavirus-neutralizing antibodies are not associated with lower disease severity in PUUV HFRS. Wide individual differences occur in ocular and central nervous system (CNS) manifestations and in the long-term consequences of nephropathia epidemica (NE). Numerous biomarkers have been detected, and some are clinically used to assess and predict the severity of PUUV infection. A new addition is the plasma glucose concentration associated with the severity of both capillary leakage, thrombocytopenia, inflammation, and AKI in PUUV infection. Our question, “Why this variation?” remains largely unanswered. Full article
24 pages, 3647 KiB  
Article
The Antiviral Compound PSP Inhibits HIV-1 Entry via PKR-Dependent Activation in Monocytic Cells
by Eduardo Alvarez-Rivera, Madeline Rodríguez-Valentín and Nawal M. Boukli
Viruses 2023, 15(3), 804; https://doi.org/10.3390/v15030804 - 22 Mar 2023
Cited by 3 | Viewed by 3476
Abstract
Actin depolymerization factor (ADF) cofilin-1 is a key cytoskeleton component that serves to lessen cortical actin. HIV-1 manipulates cofilin-1 regulation as a pre- and post-entry requisite. Disruption of ADF signaling is associated with denial of entry. The unfolded protein response (UPR) marker Inositol-Requiring [...] Read more.
Actin depolymerization factor (ADF) cofilin-1 is a key cytoskeleton component that serves to lessen cortical actin. HIV-1 manipulates cofilin-1 regulation as a pre- and post-entry requisite. Disruption of ADF signaling is associated with denial of entry. The unfolded protein response (UPR) marker Inositol-Requiring Enzyme-1α (IRE1α) and interferon-induced protein (IFN-IP) double-stranded RNA- activated protein kinase (PKR) are reported to overlap with actin components. In our published findings, Coriolus versicolor bioactive extract polysaccharide peptide (PSP) has demonstrated anti-HIV replicative properties in THP1 monocytic cells. However, its involvement towards viral infectivity has not been elucidated before. In the present study, we examined the roles of PKR and IRE1α in cofilin-1 phosphorylation and its HIV-1 restrictive roles in THP1. HIV-1 p24 antigen was measured through infected supernatant to determine PSP’s restrictive potential. Quantitative proteomics was performed to analyze cytoskeletal and UPR regulators. PKR, IRE1α, and cofilin-1 biomarkers were measured through immunoblots. Validation of key proteome markers was done through RT-qPCR. PKR/IRE1α inhibitors were used to validate viral entry and cofilin-1 phosphorylation through Western blots. Our findings show that PSP treatment before infection leads to an overall lower infectivity. Additionally, PKR and IRE1α show to be key regulators in cofilin-1 phosphorylation and viral restriction. Full article
Show Figures

Figure 1

Figure 1
<p>Diagram depicting HIV-1 entry during infection in CD4+ cells. (<b>A</b>) Early interaction between HIV-1 and CD4 receptor initiates activation of guanine nucleotide exchange factors (GEFs) and subsequently downstream signaling through RhoA/ROCK/LIMK-1. This results in phosphorylation and inactivation of cofilin-1. Actin dynamics will shift towards the polymerization state leading to CXCR4/CCR5 co-receptor clustering as the first requirement for viral entry. Filamin A serves as a linkage factor between CD4 and co-receptors. (<b>B.1</b>) HIV-1 associates with CXCR4/CCR5 in proximity and triggers downstream dephosphorylation and activation of cofilin-1. This process is carried out by G-protein coupled receptor signaling, specifically through Gα subunit-mediated phosphatases such as SSH3. (<b>B.2</b>) Cofilin-1 will begin the breakdown of actin filaments and subsequently lead to viral fusion as the final entry requirement.</p>
Full article ">Figure 2
<p>PSP lowers HIV entry in THP1 cells. Percentages of HIV entry in PSP-treated cells analyzed through viral load of HIV p24 antigen and compared with unpaired <span class="html-italic">t</span> test. Data are represented as mean ± SEM. Statistically significant difference (**), <span class="html-italic">p</span> &lt; 0.01 is shown, <span class="html-italic">n</span> = 3.</p>
Full article ">Figure 3
<p>Inhibition of PKR and IRE1α activity facilitates HIV-1 entry. Percentages of HIV entry in PSP-treated cells supplemented with either 56.09 nM of C16-PKR or 221.8 nM of 4µ8C-IRE1α pharmaceutical blockers. Experiments were analyzed through viral loads of the HIV p24 antigen and compared with one-way ANOVA with Tukey multiple comparisons tests. Data are represented as mean ± SEM. Statistically significant difference (***), <span class="html-italic">p</span> &lt; 0.001 and (****), <span class="html-italic">p</span> &lt; 0.0001 are shown, <span class="html-italic">n</span> = 3.</p>
Full article ">Figure 4
<p>Western blot results related to UPR, IFN, and cytoskeletal biomarkers in PSP-treated THP1 monocytic cells. (<b>A</b>) Representative Western blot data for the protein expression levels of UPR: (<b>B</b>) GRP78; (<b>C</b>) IRE1α; IFN: (<b>D</b>) PKR; (<b>E</b>) p-PKR; ADF: (<b>F</b>) Cofilin1; (<b>G</b>) p-Cofilin1; and ABP: (<b>H</b>) Gelsolin signaling. β-Actin was used as a loading control and for normalization of data. Images were quantified using the ImageJ software (NIH, version 1.52a) by comparing the integrated density value with the control group. Mean ± SEM and significant difference (*), <span class="html-italic">p</span> ≤ 0.05, (**), <span class="html-italic">p</span> ≤ 0.01, (***), <span class="html-italic">p</span> ≤ 0.001, (****), <span class="html-italic">p</span> ≤ 0.0001 are shown and were determined using one-way ANOVA with Tukey multiple comparisons test, <span class="html-italic">n</span> = 3. All Western blot images can be found in <a href="#app1-viruses-15-00804" class="html-app">Figure S2</a>.</p>
Full article ">Figure 5
<p>PKR inhibition decreases cofilin-1 phosphorylation and the expression levels relating to gelsolin. The activation of PKR was suppressed after 56.09 nM of C16 inhibitor in THP1 monocytic cells with or without PSP treatment. (<b>A</b>) Representative Western blot data for the protein expression levels of cytoskeletal: (<b>B</b>) pCofilin1; (<b>C</b>) Cofilin1; (<b>D</b>) Gelsolin; and IFN-IP: (<b>E</b>) pPKR; (<b>F</b>) PKR. β-Actin was used as a loading control and for normalization of data. Images were quantified using the ImageJ software (NIH, version 1.52a) by comparing the integrated density value with the control group. Mean ± SEM and significant difference (*), <span class="html-italic">p</span> ≤ 0.05, (**), <span class="html-italic">p</span> ≤ 0.01, (***), <span class="html-italic">p</span> ≤ 0.001, (****), <span class="html-italic">p</span> ≤ 0.0001 are shown and were determined using one-way ANOVA with Tukey multiple comparisons test, <span class="html-italic">n</span> = 3. All Western blot images can be found in <a href="#app1-viruses-15-00804" class="html-app">Figure S2</a>.</p>
Full article ">Figure 6
<p>IRE1α modulates downstream phosphorylation activity of PKR and cofilin-1. Endoribonuclease activity of IRE1α was suppressed after 221.8 nM of 4µ8C inhibitor in THP1 monocytic cells with or without PSP treatment. (<b>A</b>) Representative Western blot data for the protein expression levels of cytoskeletal: (<b>B</b>) pCofilin1; (<b>C</b>) Cofilin1; (<b>D</b>) Gelsolin; IFN-IP: (<b>E</b>) p-PKR; (<b>F</b>) PKR and UPR: (<b>G</b>) XBP1s. β-Actin was used as a loading control and for normalization of data. Images were quantified using the ImageJ software (NIH, version 1.52a) by comparing the integrated density value with the control group. Mean ± SEM and significant difference (*), <span class="html-italic">p</span> ≤ 0.05, (**), <span class="html-italic">p</span> ≤ 0.01, (***), <span class="html-italic">p</span> ≤ 0.001, (****), <span class="html-italic">p</span> ≤ 0.0001 are shown and were determined using one-way ANOVA with Tukey multiple comparisons test, <span class="html-italic">n</span> = 3. All Western blot images can be found in <a href="#app1-viruses-15-00804" class="html-app">Figure S2</a>.</p>
Full article ">Figure 7
<p>Validation of western blot results using RT-qPCR analysis for cytoskeletal and IFN-IP markers. Relative gene expression analyzed through RT-qPCR approach. The mRNA expression levels for the ADF: (<b>A</b>) Cofilin-1; ABP: (<b>B</b>) Gelsolin; and IFN-IP: (<b>C</b>) PKR are shown. All data were normalized using 18S as a housekeeping gene in response to PSP treatment. Mean ± SEM and significant difference (*), <span class="html-italic">p</span> ≤ 0.05, (**), <span class="html-italic">p</span> ≤ 0.01, (***), <span class="html-italic">p</span> ≤ 0.001 are shown and were determined using one-way ANOVA, with Tukey multiple comparisons test, <span class="html-italic">n</span> = 3.</p>
Full article ">Figure 8
<p>Model depicting PSP signaling through a UPR/IFN-induced pathway. HIV-1 infection results in a chronic ER stress response, while simultaneously dephosphorylating cofilin-1 through SSH3 phosphatase. This results in actin depolymerization for viral entry. Prior to infection, PSP treatment induces actin polymerization via an acute UPR. PKR mediates downstream phosphorylation of IRE1α signals and reverses HIV-induced actin remodeling while infection persists. Legend colors: Green—PSP-mediated signaling and events. Cream- HIV-downstream pathways.</p>
Full article ">
19 pages, 8910 KiB  
Article
Design of Phage-Cocktail–Containing Hydrogel for the Treatment of Pseudomonas aeruginosa–Infected Wounds
by Fatemeh Shafigh Kheljan, Farzam Sheikhzadeh Hesari, Mohammad Sadegh Aminifazl, Mikael Skurnik, Sophia Goladze and Gholamreza Zarrini
Viruses 2023, 15(3), 803; https://doi.org/10.3390/v15030803 - 21 Mar 2023
Cited by 10 | Viewed by 3550
Abstract
Recently, the treatment of infected wounds has become a global problem due to increased antibiotic resistance in bacteria. The Gram-negative opportunistic pathogen Pseudomonas aeruginosa is often present in chronic skin infections, and it has become a threat to public health as it is [...] Read more.
Recently, the treatment of infected wounds has become a global problem due to increased antibiotic resistance in bacteria. The Gram-negative opportunistic pathogen Pseudomonas aeruginosa is often present in chronic skin infections, and it has become a threat to public health as it is increasingly multidrug resistant. Due to this, new measures to enable treatment of infections are necessary. Treatment of bacterial infections with bacteriophages, known as phage therapy, has been in use for a century, and has potential with its antimicrobial effect. The main purpose of this study was to create a phage-containing wound dressing with the ability to prevent bacterial infection and rapid wound healing without side effects. Several phages against P. aeruginosa were isolated from wastewater, and two polyvalent phages were used to prepare a phage cocktail. The phage cocktail was loaded in a hydrogel composed of polymers of sodium alginate (SA) and carboxymethyl cellulose (CMC). To compare the antimicrobial effects, hydrogels containing phages, ciprofloxacin, or phages plus ciprofloxacin were produced, and hydrogels without either. The antimicrobial effect of these hydrogels was investigated in vitro and in vivo using an experimental mouse wound infection model. The wound-healing process in different mouse groups showed that phage-containing hydrogels and antibiotic-containing hydrogels have almost the same antimicrobial effect. However, in terms of wound healing and pathological process, the phage-containing hydrogels performed better than the antibiotic alone. The best performance was achieved with the phage–antibiotic hydrogel, indicating a synergistic effect between the phage cocktail and the antibiotic. In conclusion, phage-containing hydrogels eliminate efficiently P. aeruginosa in wounds and may be a proper option for treating infectious wounds. Full article
(This article belongs to the Special Issue Phage and Antibiotic Combination Therapy against MDR Bacteria)
Show Figures

Figure 1

Figure 1
<p>Plaques of phages (<b>A</b>) PB10 (10<sup>7</sup> PFU/mL) and (<b>B</b>) PA19 (10<sup>7</sup> PFU/mL) on <span class="html-italic">P. aeruginosa</span> ATCC27853.</p>
Full article ">Figure 2
<p>Morphology of phage particles under transmission electron microscope. Phage PB10 (<b>A</b>) and PA19 (<b>B</b>). The scale bars represent 100 nm.</p>
Full article ">Figure 3
<p>Genome comparison of phages PB10 and PA19. Panel (<b>A</b>). Alignment of the genomes. The DNA-identity-% is indicated by the “Identity” track with 100% identity, indicated by green color, while the height of the graph indicates the identity%. The alignment was generated by the Geneious version 2022.2.2 using the Clustal Omega vs. 1.2.2. Panel (<b>B</b>). Alignment of amino acid sequences of the Gp62 tail fiber proteins demonstrating the lower similarity between the C-terminal thirds of the proteins; the differences are highlighted in grey.</p>
Full article ">Figure 4
<p>(<b>A</b>) Hydrogel in the form of gel, (<b>B</b>) Hydrogel film form.</p>
Full article ">Figure 5
<p>(<b>A</b>) The swelling index and (<b>B</b>) degradation of SA-CMC hydrogels. The results are based on three repeats showing the average with standard deviations. <span class="html-graphic" id="viruses-15-00803-i001"><img alt="Viruses 15 00803 i001" src="/viruses/viruses-15-00803/article_deploy/html/images/viruses-15-00803-i001.png"/></span> Based on the analysis of <a href="#viruses-15-00803-f005" class="html-fig">Figure 5</a>A, hydrogel swelling on day 5 is significant with the rest of the days (<span class="html-italic">p</span> &lt; 0.001) and the day 1 is significant with the rest of the days, except for day 2, at the level of <span class="html-italic">p</span> &lt; 0.05. Moreover, day 2 is significant with day 4 (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 6
<p>Water vapor transfer rate of hydrogel.</p>
Full article ">Figure 7
<p>FTIR spectra of (<b>A</b>). SA, (<b>B</b>). CMC, (<b>C</b>). Hydrogel. The peaks in the range 3400–3500 cm<sup>−1</sup> are related to (O-H), those in the 2900–3000 cm<sup>−1</sup> range to (C-H) bonds, those in the 1427–1639 cm<sup>−1</sup> range to (C=O) bonds, and at 1125 cm<sup>−1</sup> to (C-O-C) bonds.</p>
Full article ">Figure 8
<p>Surface morphology of (<b>A</b>). Pure hydrogel with uniform and smooth surface, (<b>B</b>). Phage-containing hydrogel with protrusion surface.</p>
Full article ">Figure 9
<p>Inhibition zones of the hydrogels against <span class="html-italic">P. aeruginosa</span>.</p>
Full article ">Figure 10
<p>Quantitative antibacterial assay. The colony count results of different groups. The results are based on four repeats showing the average with standard deviations. <span class="html-graphic" id="viruses-15-00803-i002"><img alt="Viruses 15 00803 i002" src="/viruses/viruses-15-00803/article_deploy/html/images/viruses-15-00803-i002.png"/></span> The statistical analysis of the five groups in all the tested hours indicated a significant difference between the two control and hydrogel groups with the other three groups at the level of <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 11
<p>Evaluation of the stability of the phages in the hydrogels over 8 weeks, as estimated by the reduced killing efficiency over time. The results are based on three repeats, showing the average with standard deviations. <span class="html-graphic" id="viruses-15-00803-i003"><img alt="Viruses 15 00803 i003" src="/viruses/viruses-15-00803/article_deploy/html/images/viruses-15-00803-i003.png"/></span> Based on the analysis, it was found that phage 4 is significant with all other weeks of the phage group (<span class="html-italic">p</span> &lt; 0.001). <span class="html-graphic" id="viruses-15-00803-i004"><img alt="Viruses 15 00803 i004" src="/viruses/viruses-15-00803/article_deploy/html/images/viruses-15-00803-i004.png"/></span> indicates that the phage–antibiotic 8 is significant with the other weeks of the phage–antibiotic group at the <span class="html-italic">p</span> &lt; 0.001 level.</p>
Full article ">Figure 12
<p>Bacterial load in different mouse groups on days 1, 3, 7, 10 and 12. C: Control, B: Untreated infectious, H: Hydrogel, PD: Phage (daily), PO: Phage (once), AD: Antibiotic (daily), AO: Antibiotic (once), PAD: Phage-Antibiotic (daily), PAO: Phage-Antibiotic (once). * <span class="html-italic">p</span> &lt; 0.001 means significant relative to untreated infectious group.</p>
Full article ">Figure 13
<p>Pictures of wound-healing process over 14 days in different groups.</p>
Full article ">Figure 14
<p>The wound healing rates based on wound sizes in different mouse groups. C: Control, B: Untreated infected wounds, H: Hydrogel, PD: Phage (daily), PO: Phage (once), AD: Antibiotic (daily), AO: Antibiotic (once), PAD: Phage-Antibiotic (daily), PAO: Phage-Antibiotic (once). According to the figure, the wound sizes of all the groups were not significantly different from each other on day one. <span class="html-graphic" id="viruses-15-00803-i005"><img alt="Viruses 15 00803 i005" src="/viruses/viruses-15-00803/article_deploy/html/images/viruses-15-00803-i005.png"/></span> indicates that the wound size of group AO was significantly different from those of groups PD and PAD on day 3 of the treatment (<span class="html-italic">p</span> &lt; 0.05). <span class="html-graphic" id="viruses-15-00803-i006"><img alt="Viruses 15 00803 i006" src="/viruses/viruses-15-00803/article_deploy/html/images/viruses-15-00803-i006.png"/></span> indicates that on day 7 of the treatment, the wound size of the AO group was significantly different to those of the PD, PAD and PAO groups (<span class="html-italic">p</span> &lt; 0.01). <span class="html-graphic" id="viruses-15-00803-i007"><img alt="Viruses 15 00803 i007" src="/viruses/viruses-15-00803/article_deploy/html/images/viruses-15-00803-i007.png"/></span> indicates that the wound sizes of groups C and PD were significantly different on day 10 of the treatment (<span class="html-italic">p</span> &lt; 0.01). <span class="html-graphic" id="viruses-15-00803-i008"><img alt="Viruses 15 00803 i008" src="/viruses/viruses-15-00803/article_deploy/html/images/viruses-15-00803-i008.png"/></span> Based on the analysis, the wound size of the group C was significantly different from those of groups AD, PAD and PD, and the wound size of group AO was significantly different from those of groups PD and PAO on day 14 of the treatment (<span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 15
<p>Histopathological sections of the skin in different groups. The black arrows show the repaired epidermis, the yellow arrows the macrophages and the green arrows the hair follicles.</p>
Full article ">
12 pages, 1260 KiB  
Article
Molecular Characterization and Cluster Analysis of SARS-CoV-2 Viral Isolates in Kahramanmaraş City, Turkey: The Delta VOC Wave within One Month
by Nadia Marascio, Merve Cilburunoglu, Elif Gulsum Torun, Federica Centofanti, Elida Mataj, Michele Equestre, Roberto Bruni, Angela Quirino, Giovanni Matera, Anna Rita Ciccaglione and Kezban Tulay Yalcinkaya
Viruses 2023, 15(3), 802; https://doi.org/10.3390/v15030802 - 21 Mar 2023
Cited by 2 | Viewed by 1992
Abstract
The SARS-CoV-2 pandemic has seriously affected the population in Turkey. Since the beginning, phylogenetic analysis has been necessary to monitor public health measures against COVID-19 disease. In any case, the analysis of spike (S) and nucleocapsid (N) gene mutations was crucial in determining [...] Read more.
The SARS-CoV-2 pandemic has seriously affected the population in Turkey. Since the beginning, phylogenetic analysis has been necessary to monitor public health measures against COVID-19 disease. In any case, the analysis of spike (S) and nucleocapsid (N) gene mutations was crucial in determining their potential impact on viral spread. We screened S and N regions to detect usual and unusual substitutions, whilst also investigating the clusters among a patient cohort resident in Kahramanmaraş city, in a restricted time span. Sequences were obtained by Sanger methods and genotyped by the PANGO Lineage tool. Amino acid substitutions were annotated comparing newly generated sequences to the NC_045512.2 reference sequence. Clusters were defined using phylogenetic analysis with a 70% cut-off. All sequences were classified as Delta. Eight isolates carried unusual mutations on the S protein, some of them located in the S2 key domain. One isolate displayed the unusual L139S on the N protein, while few isolates carried the T24I and A359S N substitutions able to destabilize the protein. Phylogeny identified nine monophyletic clusters. This study provided additional information about SARS-CoV-2 epidemiology in Turkey, suggesting local transmission of infection in the city by several transmission routes, and highlighting the necessity to improve the power of sequencing worldwide. Full article
(This article belongs to the Special Issue SARS-CoV-2 and Other Coronaviruses)
Show Figures

Figure 1

Figure 1
<p>Heatmap of S (light green) and N (orange) amino acid substitutions detected in each isolate. The appearance of substitutions was organized based on their inclusion in a cluster (colored box) or not (white box) by phylogenetic analysis. Nucleotide (nt) position of sequenced genes is reported in the lower part of the figure.</p>
Full article ">Figure 2
<p>ML phylogenetic tree was estimated using 99 sequences from Turkey and 58 SARS-CoV-2 newly generated sequences. The tree was rooted by the midpoint rooting according to the NC_045512.2 (highlighted in purple) outgroup sequence. The reliability of the phylogenetic clustering was evaluated using bootstrap analysis with 1000 replicates. Bootstrap support values (&gt;70%) are only shown for the clusters (colored collapsed cartoons) including Kahramanmaraş city isolates. The nine clusters were identified with the alphabetic letters from A to I. The scale bar at the bottom of the figure represents genetic distance (0.005).</p>
Full article ">
18 pages, 7178 KiB  
Article
Evaluation of SARS-CoV-2 ORF7a Deletions from COVID-19-Positive Individuals and Its Impact on Virus Spread in Cell Culture
by Maria Clara da Costa Simas, Sara Mesquita Costa, Priscila da Silva Figueiredo Celestino Gomes, Nádia Vaez Gonçalves da Cruz, Isadora Alonso Corrêa, Marcos Romário Matos de Souza, Marcos Dornelas-Ribeiro, Tatiana Lucia Santos Nogueira, Caleb Guedes Miranda dos Santos, Luísa Hoffmann, Amilcar Tanuri, Rodrigo Soares de Moura-Neto, Clarissa R. Damaso, Luciana Jesus da Costa and Rosane Silva
Viruses 2023, 15(3), 801; https://doi.org/10.3390/v15030801 - 21 Mar 2023
Cited by 4 | Viewed by 2890
Abstract
The spread of severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2), causing the COVID-19 outbreak, posed a primary concern of public health worldwide. The most common changes in SARS-CoV-2 are single nucleotide substitutions, also reported insertions and deletions. This work investigates the presence of [...] Read more.
The spread of severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2), causing the COVID-19 outbreak, posed a primary concern of public health worldwide. The most common changes in SARS-CoV-2 are single nucleotide substitutions, also reported insertions and deletions. This work investigates the presence of SARS-CoV-2 ORF7a deletions identified in COVID-19-positive individuals. Sequencing of SARS-CoV-2 complete genomes showed three different ORF7a size deletions (190-nt, 339-nt and 365-nt). Deletions were confirmed through Sanger sequencing. The ORF7a∆190 was detected in a group of five relatives with mild symptoms of COVID-19, and the ORF7a∆339 and ORF7a∆365 in a couple of co-workers. These deletions did not affect subgenomic RNAs (sgRNA) production downstream of ORF7a. Still, fragments associated with sgRNA of genes upstream of ORF7a showed a decrease in size when corresponding to samples with deletions. In silico analysis suggests that the deletions impair protein proper function; however, isolated viruses with partial deletion of ORF7a can replicate in culture cells similarly to wild-type viruses at 24 hpi, but with less infectious particles after 48 hpi. These findings on deleted ORF7a accessory protein gene, contribute to understanding SARS-CoV-2 phenotypes such as replication, immune evasion and evolutionary fitness as well insights into the role of SARS-CoV-2_ORF7a in the mechanism of virus-host interactions. Full article
(This article belongs to the Collection Coronaviruses)
Show Figures

Figure 1

Figure 1
<p>A 190-nucleotide deletion in the ORF7a gene (ORF7a∆190) and impact on the protein structure. (<b>A</b>) Top: Amplification and confirmation of 190-nt deletions amongst family members using primers forward (M_27145 F) and reverse (ORF8_28003R) (see scheme) indicating the positions on the SARS-CoV-2 genome. Bottom: ORF7a 190-nt deletions visualization in five family members on agarose gel (2%) stained with ethidium bromide revealed that the PCR products target the ORF7a gene. 100bp Molecular Weight Marker (lane M), 11785_ORF7a_wt, 58301_ORF7a_∆190, 59977_ORF7a_∆190, 59980_ORF7a_∆190, 58306_ORF7a∆190, 57817_ORF7a∆190. (<b>B</b>) Electropherograms from Sanger sequencing of the ORF7a PCR products from samples with 190-nt deletions: 59980, 59977, 58306, 58301 and 57817, the wt (without 190-nt deletion) 11785 aligned to the SARS-CoV-2 reference sequence (NCBI RefSeq SARS-CoV-2 genome sequence, NC_045512.2). (<b>C</b>) ORF7a deletions and impact on the protein structure. Left: Multiple sequence alignment of the reported ORF7a deletion (58301_ORF7a_∆190), and the corresponding sequences for SARS-CoV-2 and SARS-CoV ectodomain structures (PDB IDs: 7ci3 and 1xak, respectively) in respect of the reference sequence (NCBI RefSeq for SARS-CoV-2 ORF7a protein, YP_009724395.1). Identical residues are shaded in red; mutations from SARS-CoV to SARS-CoV-2 are colored in black and ORF7a altered residues are colored in grey. The labeled arrows indicate the secondary structure observed on the crystal structures. The different protein domains are indicated by the color bars below the alignment: N-terminal signaling region (residues 1–15, blue), Ig-like ectodomain (residues 16–96, yellow), hydrophobic transmembrane domain (residues 97–116, green) and ER retention motif (residues 117–121, brown). Right: Ribbon representation of the tridimensional SARS-CoV-2 ORF7a structure. The full-length structure of the ectodomain is colored in cyan. The β-strands are assigned numerically from the N-terminus to the C-terminus, corresponding to the alignment above. The deletion presented by the 58301_ORF7a_∆190 sample is colored in magenta, in both structures, covering a significant portion of the protein ectodomain, comprising the β sheets 3 to 7, the full transmembrane and ER retention motif. (<b>D</b>) Sars-CoV-2 whole genome sequencing of samples belonging to the five family members mapped to the reference sequence, aligned and visualized on IGV viewer. ORF7a region displays gaps evidencing the 190-nt deletion in all five samples.</p>
Full article ">Figure 1 Cont.
<p>A 190-nucleotide deletion in the ORF7a gene (ORF7a∆190) and impact on the protein structure. (<b>A</b>) Top: Amplification and confirmation of 190-nt deletions amongst family members using primers forward (M_27145 F) and reverse (ORF8_28003R) (see scheme) indicating the positions on the SARS-CoV-2 genome. Bottom: ORF7a 190-nt deletions visualization in five family members on agarose gel (2%) stained with ethidium bromide revealed that the PCR products target the ORF7a gene. 100bp Molecular Weight Marker (lane M), 11785_ORF7a_wt, 58301_ORF7a_∆190, 59977_ORF7a_∆190, 59980_ORF7a_∆190, 58306_ORF7a∆190, 57817_ORF7a∆190. (<b>B</b>) Electropherograms from Sanger sequencing of the ORF7a PCR products from samples with 190-nt deletions: 59980, 59977, 58306, 58301 and 57817, the wt (without 190-nt deletion) 11785 aligned to the SARS-CoV-2 reference sequence (NCBI RefSeq SARS-CoV-2 genome sequence, NC_045512.2). (<b>C</b>) ORF7a deletions and impact on the protein structure. Left: Multiple sequence alignment of the reported ORF7a deletion (58301_ORF7a_∆190), and the corresponding sequences for SARS-CoV-2 and SARS-CoV ectodomain structures (PDB IDs: 7ci3 and 1xak, respectively) in respect of the reference sequence (NCBI RefSeq for SARS-CoV-2 ORF7a protein, YP_009724395.1). Identical residues are shaded in red; mutations from SARS-CoV to SARS-CoV-2 are colored in black and ORF7a altered residues are colored in grey. The labeled arrows indicate the secondary structure observed on the crystal structures. The different protein domains are indicated by the color bars below the alignment: N-terminal signaling region (residues 1–15, blue), Ig-like ectodomain (residues 16–96, yellow), hydrophobic transmembrane domain (residues 97–116, green) and ER retention motif (residues 117–121, brown). Right: Ribbon representation of the tridimensional SARS-CoV-2 ORF7a structure. The full-length structure of the ectodomain is colored in cyan. The β-strands are assigned numerically from the N-terminus to the C-terminus, corresponding to the alignment above. The deletion presented by the 58301_ORF7a_∆190 sample is colored in magenta, in both structures, covering a significant portion of the protein ectodomain, comprising the β sheets 3 to 7, the full transmembrane and ER retention motif. (<b>D</b>) Sars-CoV-2 whole genome sequencing of samples belonging to the five family members mapped to the reference sequence, aligned and visualized on IGV viewer. ORF7a region displays gaps evidencing the 190-nt deletion in all five samples.</p>
Full article ">Figure 2
<p>Amplification and confirmation of 339 and 365-nt deletion amongst co-workers. (<b>A</b>) ORF7a 339-nt and 365-nt deletions visualization in a group of co-workers on agarose gel (2%) stained with ethidium bromide revealed the PCR products targeting ORF7a gene: 16305_ORF7a_wt, 16553_ORF7a_wt, 16538_ORF7a_∆365, 16991_ORF7a_∆339; 100bp Molecular Weight Marker (lane M). (<b>B</b>) Prevalence of viral genomic deletion analyzed through ORF7a PCR products from serial dilutions (10<sup>–1</sup>–10<sup>−4</sup>, lane 2–5, respectively) of sample 16538_ORF7a_∆365 viral stock and visualized on agarose gel (2%) stained with ethidium bromide. The two PCR fragments are present in 10<sup>−1</sup> and 10<sup>−2</sup>, and only the non-deleted fragment is present in 10<sup>−3</sup> and 10<sup>−4</sup>. (<b>C</b>) Eleven virus plaques were isolated from sample 16538. RNA extraction, reverse transcription, and PCR using primers that target the ORF7a region were performed. PCR products were visualized on agarose gel (2%) stained with ethidium bromide. Each lane corresponds to a virus plaque (P1–P11) derived from sample 16538. Virus plaques P2, P4 and P11 correspond to the virus without deletion in the ORF7a region, and virus plaques P5, P6 and P7 correspond to the virus with 365-nt deletions in the ORF7a region (ORF7a_∆365). M = 100 bp Molecular Weight Marker. (<b>D</b>) Electropherograms from Sanger sequencing of the ORF7a PCR products from sample 16991_ORF7a_∆339 aligned to SARS-CoV-2 reference and its deletion-resulting sequence (<b>E</b>) Electropherograms of both ORF7a PCR products isolated on a gel, from clinical samples and viruses isolated from cell culture (lanes P2 and P5, (<b>B</b>)) of 16538 ORF7a WT and the ORF7a_∆ 365 of 16538_ORF7a_∆365 and the non-deleted counterpart of 16538 * Samples submitted to NGS sequencing † Samples submitted to Sanger sequencing.</p>
Full article ">Figure 2 Cont.
<p>Amplification and confirmation of 339 and 365-nt deletion amongst co-workers. (<b>A</b>) ORF7a 339-nt and 365-nt deletions visualization in a group of co-workers on agarose gel (2%) stained with ethidium bromide revealed the PCR products targeting ORF7a gene: 16305_ORF7a_wt, 16553_ORF7a_wt, 16538_ORF7a_∆365, 16991_ORF7a_∆339; 100bp Molecular Weight Marker (lane M). (<b>B</b>) Prevalence of viral genomic deletion analyzed through ORF7a PCR products from serial dilutions (10<sup>–1</sup>–10<sup>−4</sup>, lane 2–5, respectively) of sample 16538_ORF7a_∆365 viral stock and visualized on agarose gel (2%) stained with ethidium bromide. The two PCR fragments are present in 10<sup>−1</sup> and 10<sup>−2</sup>, and only the non-deleted fragment is present in 10<sup>−3</sup> and 10<sup>−4</sup>. (<b>C</b>) Eleven virus plaques were isolated from sample 16538. RNA extraction, reverse transcription, and PCR using primers that target the ORF7a region were performed. PCR products were visualized on agarose gel (2%) stained with ethidium bromide. Each lane corresponds to a virus plaque (P1–P11) derived from sample 16538. Virus plaques P2, P4 and P11 correspond to the virus without deletion in the ORF7a region, and virus plaques P5, P6 and P7 correspond to the virus with 365-nt deletions in the ORF7a region (ORF7a_∆365). M = 100 bp Molecular Weight Marker. (<b>D</b>) Electropherograms from Sanger sequencing of the ORF7a PCR products from sample 16991_ORF7a_∆339 aligned to SARS-CoV-2 reference and its deletion-resulting sequence (<b>E</b>) Electropherograms of both ORF7a PCR products isolated on a gel, from clinical samples and viruses isolated from cell culture (lanes P2 and P5, (<b>B</b>)) of 16538 ORF7a WT and the ORF7a_∆ 365 of 16538_ORF7a_∆365 and the non-deleted counterpart of 16538 * Samples submitted to NGS sequencing † Samples submitted to Sanger sequencing.</p>
Full article ">Figure 3
<p>Amplification and examination of sgRNAs from samples with and without deletions on ORF7a. (<b>A</b>) Schematic illustration of SARS-CoV-2 genome encompassing TRS-L leader sequence (light orange box) and the primers’ position for sgRNAs amplification. Green dots represent the TRS-B sequences at the 5′ of each ORF of the structural and accessory proteins. (<b>B</b>) Sanger sequencing visualization of sgRNAs fragments from samples ORF7a_wt, ORF7a_∆190 and ORF7a_∆339 as illustrated in <a href="#app1-viruses-15-00801" class="html-app">Figure S2</a>. The light orange box indicates the 75-nt leader sequence, the red bar represents the forward primer SARS-Leader 18-41F, yellow boxes represent coding sequences (ORF7a, ORF8 and N), the reverse primers are not shown.</p>
Full article ">Figure 4
<p>SARS-CoV-2 ORF7a 190-nt deletion reduces virus infectivity in Vero E6 cells. (<b>A</b>) Vero E6 cells were infected at MOI of 0.1 of SARS-CoV-2_wt (11785_wt) or SARS-CoV-2_ORF7a_∆190 (58301_ORF7a_∆190). (<b>B</b>) Cell supernatant was collected at 24, 48 and 72 hpi and used for virus titration in Vero E6 cells. Virus titration (mean ± SD) was plotted in a graphic. The virus titles at each time point were compared using multiple unpaired t-tests. PFU = plaque forming units (<b>C</b>) Cell supernatant was collected at 24, 48 and 72 hpi and used to perform RT-qPCR for SARS-CoV-2 N. SARS-CoV-2 RNA equivalent to PFU (mean ± SD) was plotted in the graphic. The virus titrations at each time point were compared using multiple unpaired <span class="html-italic">t</span>-tests. Results are from three independent experiments. * <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; ns indicates no significant difference (<span class="html-italic">p</span> &gt; 0.05).</p>
Full article ">
12 pages, 3151 KiB  
Article
Two Point Mutations in the Glycoprotein of SFTSV Enhance the Propagation Recombinant Vesicular Stomatitis Virus Vectors at Assembly Step
by Qiang Hu, Yuhang Zhang, Jiafu Jiang and Aihua Zheng
Viruses 2023, 15(3), 800; https://doi.org/10.3390/v15030800 - 21 Mar 2023
Cited by 2 | Viewed by 2019
Abstract
Severe fever with thrombocytopenia syndrome virus (SFTSV) is an emerging tick-borne pathogen for which approved therapeutic drugs or vaccines are not available. We previously developed a recombinant vesicular stomatitis virus-based vaccine candidate (rVSV-SFTSV) by replacing the original glycoprotein with Gn/Gc from SFTSV, which [...] Read more.
Severe fever with thrombocytopenia syndrome virus (SFTSV) is an emerging tick-borne pathogen for which approved therapeutic drugs or vaccines are not available. We previously developed a recombinant vesicular stomatitis virus-based vaccine candidate (rVSV-SFTSV) by replacing the original glycoprotein with Gn/Gc from SFTSV, which conferred complete protection in a mouse model. Here, we found that two spontaneous mutations, M749T/C617R, emerged in the Gc glycoprotein during passaging that could significantly increase the titer of rVSV-SFTSV. M749T/C617R enhanced the genetic stability of rVSV-SFTSV, and no further mutations appeared after 10 passages. Using immunofluorescence analysis, we found that M749T/C617R could increase glycoprotein traffic to the plasma membrane, thus facilitating virus assembly. Remarkably, the broad-spectrum immunogenicity of rVSV-SFTSV was not affected by the M749T/C617R mutations. Overall, M749T/C617R could enhance the further development of rVSV-SFTSV into an effective vaccine in the future. Full article
(This article belongs to the Special Issue Advances in Antiviral Immunity and Virus Vaccines)
Show Figures

Figure 1

Figure 1
<p>Characterization of rVSV-SFTSV variants. (<b>a</b>) Schematic of the mutations that emerged in the SFTSV Gc protein. Mutation (Mut). Mutated nucleotides (Red) (<b>b</b>) Gc expression in the supernatants and cell lysates of rVSV-SFTSV WT- and variant-infected Vero cells (MOI = 0.01). Forty-eight hours post-infection, purified particles and cell lysates were blotted with anti-SFTSV Gc polyclonal antibodies (GAPDH as control). Con. indicates mock-infected cells. (<b>c</b>) Quantification of (<b>b</b>). (<b>d</b>) Growth kinetics of rVSVs (MOI = 0.01). *** <span class="html-italic">t</span> test, <span class="html-italic">p</span> &lt; 0.001; **** <span class="html-italic">t</span> test, <span class="html-italic">p</span> &lt; 0.0001. (<b>e</b>) Representative image of plaques formed using rVSV-SFTSV variants (6 days), rVSV-HRTV (6 days), and rVSV-G (3 days) at indicated days post-infection in Vero cells. Scale bar: 500 μm. (<b>f</b>) Genetic stability of rVSV-WT and rVSV-M749T + C617R during passages (P) in Vero cells. The above data are representative of three independent experiments.</p>
Full article ">Figure 2
<p>The effect of M749T and C617R mutations on the plasma membrane localization of Gc. (<b>a</b>) Vero cells were infected with rVSV-WT and variants 24 h after transfection with a plasmid encoding the ER reporter DsRed-KDEL (red) and stained with anti-Gc polyclonal antibodies (green) 24 h post-infection. (<b>b</b>) Live Vero cells were infected with rVSVs for 16 h and stained with anti-Gc polyclonal antibodies (green). (<b>c</b>) Quantification of the density of Gc in the whole cell from (<b>a</b>) (<span class="html-italic">n</span> = 60). (<b>d</b>) Quantification of Gc in plasma membrane from (<b>b</b>) (<span class="html-italic">n</span> = 60). Statistical significance was determined using an unpaired Student’s <span class="html-italic">t</span>-test. ns, <span class="html-italic">p</span> &gt; 0.05; **** <span class="html-italic">p</span> &lt; 0.0001. The above data are representative of three independent experiments.</p>
Full article ">Figure 3
<p>The humoral responses in mice elicited using rVSV-WT and rVSV-M749T + C617R. Three groups of C57/BL6 IFNAR<sup>−/−</sup> mice (<span class="html-italic">n</span> = 5 per group) were i.p. immunized with rVSV-WT, rVSV-M749T + C617R, and DMEM (2 × 10<sup>4</sup> PFU). (<b>a</b>) Mouse body weight changes after immunization. (<b>b</b>) NAb titers were determined against rVSV-GFP-SFTSV AH12 (<b>b</b>), rVSV-GFP-SFTSV YG1 (<b>c</b>), rVSV-GFP-HRTV (<b>d</b>), and rVSV-GFP-G (<b>e</b>) 28 days after immunization. Titers were calculated using the Reed–Muench method. The above data are representative of two independent experiments. The <span class="html-italic">p</span>-value was determined using a two-sided multiple <span class="html-italic">t</span>-test; **** <span class="html-italic">t</span>-test, <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 4
<p>rVSV-WT and rVSV-M749T + C617R confer complete protection against lethal SFTSV Wuhan strain challenge in mice. (<b>a</b>) Survival rate of the mice after challenge. Immunized mice were i.p. challenged with a lethal dose of SFTSV (2 × 10<sup>4</sup> PFU) 30 days after immunization. (<b>b</b>) Body weight changes in mice after SFTSV challenge. Data are presented as the means  ±  SD. (<b>c</b>) SFTSV viral RNAs in the mice after challenge as measured using real-time PCR. The black bar is the mock-immunized control (con.); The dotted line represents the detection limit. (<b>d</b>) Viremia in the sera was measured using a plaque assay in Vero cells. Statistical significance was determined using multiple <span class="html-italic">t</span>-test. The above data are representative of two independent experiments. The <span class="html-italic">p</span>-value was determined using a two-sided multiple <span class="html-italic">t</span>-test; **** <span class="html-italic">t</span>-test, <span class="html-italic">p</span> &lt; 0.0001; *** <span class="html-italic">t</span>-test, <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">
13 pages, 2421 KiB  
Article
Dynamic of Mayaro Virus Transmission in Aedes aegypti, Culex quinquefasciatus Mosquitoes, and a Mice Model
by Larissa Krokovsky, Carlos Ralph Batista Lins, Duschinka Ribeiro Duarte Guedes, Gabriel da Luz Wallau, Constância Flávia Junqueira Ayres and Marcelo Henrique Santos Paiva
Viruses 2023, 15(3), 799; https://doi.org/10.3390/v15030799 - 21 Mar 2023
Cited by 5 | Viewed by 2097
Abstract
Mayaro virus (MAYV) is transmitted by Haemagogus spp. mosquitoes and has been circulating in Amazon areas in the North and Central West regions of Brazil since the 1980s, with an increase in human case notifications in the last 10 years. MAYV introduction in [...] Read more.
Mayaro virus (MAYV) is transmitted by Haemagogus spp. mosquitoes and has been circulating in Amazon areas in the North and Central West regions of Brazil since the 1980s, with an increase in human case notifications in the last 10 years. MAYV introduction in urban areas is a public health concern as infections can cause severe symptoms similar to other alphaviruses. Studies with Aedes aegypti have demonstrated the potential vector competence of the species and the detection of MAYV in urban populations of mosquitoes. Considering the two most abundant urban mosquito species in Brazil, we investigated the dynamics of MAYV transmission by Ae. aegypti and Culex quinquefasciatus in a mice model. Mosquito colonies were artificially fed with blood containing MAYV and infection (IR) and dissemination rates (DR) were evaluated. On the 7th day post-infection (dpi), IFNAR BL/6 mice were made available as a blood source to both mosquito species. After the appearance of clinical signs of infection, a second blood feeding was performed with a new group of non-infected mosquitoes. RT-qPCR and plaque assays were carried out with animal and mosquito tissues to determine IR and DR. For Ae. aegypti, we found an IR of 97.5–100% and a DR reached 100% in both 7 and 14 dpi. While IR and DR for Cx. quinquefasciatus was 13.1–14.81% and 60% to 80%, respectively. A total of 18 mice were used (test = 12 and control = 6) for Ae. aegypti and 12 (test = 8 and control = 4) for Cx. quinquefasciatus to evaluate the mosquito–mice transmission rate. All mice that were bitten by infected Ae. aegypti showed clinical signs of infection while all mice exposed to infected Cx. quinquefasciatus mosquitoes remained healthy. Viremia in the mice from Ae. aegypti group ranged from 2.5 × 108 to 5 × 109 PFU/mL. Ae. aegypti from the second blood feeding showed a 50% IR. Our study showed the applicability of an efficient model to complete arbovirus transmission cycle studies and suggests that the Ae. aegypti population evaluated is a competent vector for MAYV, while highlighting the vectorial capacity of Ae. aegypti and the possible introduction into urban areas. The mice model employed here is an important tool for arthropod–vector transmission studies with laboratory and field mosquito populations, as well as with other arboviruses. Full article
(This article belongs to the Special Issue Animal Flaviviruses and Alphaviruses)
Show Figures

Figure 1

Figure 1
<p>Illustration of the vector competence study design of Mayaro virus transmission in a mouse model. (<b>A</b>) Vector competence study design in a timeline. (<b>B</b>) Mayaro transmission cycle model in a timeline. MID—midguts; SG—salivary glands; and MAYV—Mayaro virus. The illustration was created by the authors using the software Inkscape v 1.2.</p>
Full article ">Figure 2
<p>Infection and dissemination rates of <span class="html-italic">Aedes aegypti</span> (RecLab) and <span class="html-italic">Culex quinquefasciatus</span> (CqSLab) mosquitoes experimentally fed with blood containing Mayaro. (<b>A</b>) Representative graph of infection rates found in <span class="html-italic">Aedes aegypti</span> (RecLab) and <span class="html-italic">Culex quinquefasciatus</span> (CqSLab). (<b>B</b>) Representative graph of dissemination rates found in <span class="html-italic">Aedes aegypti</span> (RecLab) and <span class="html-italic">Culex quinquefasciatus</span> (CqSLab). Dpi—day post-infection; ns—not significant. Statistical analysis was performed using R software (R DEVELOPMENT CORE TEAM) (*** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 3
<p>Positive midgut and salivary glands of <span class="html-italic">Aedes aegypti</span> and <span class="html-italic">Culex quinquefasciatus</span> by RT-qPCR for Mayaro virus. (<b>A</b>,<b>B</b>) Quantification of RNA viral copy numbers in the positive midguts and salivary glands of <span class="html-italic">Aedes aegypti</span> and <span class="html-italic">Culex quinquefasciatus</span> mosquitoes experimentally fed with blood containing MAYV. Dpi—day post-infection; ns—not significant. Statistical analysis was performed using the R software (R DEVELOPMENT CORE TEAM) ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 4
<p>Animal experimentation panel with results of IFNAR BL/6 mice infected with Mayaro and mosquitoes <span class="html-italic">Aedes aegypti</span>. (<b>A</b>) Foot edema presented in mice hind foot after Mayaro virus infection compared with mock hind foot showed in bottom view; (<b>B</b>) foot edema presented in mice hind foot after Mayaro infection and mock hind foot showed in top view; (<b>C</b>) visualization of plaque-forming units in different mice samples at 10<sup>−5</sup> dilution in VERO cell plaque assay; (<b>D</b>) Mayaro titer in mice tissues infected with Mayaro virus after <span class="html-italic">Aedes aegypti</span> blood feeding; (<b>E</b>) RNA copy number of mice tissues and <span class="html-italic">Aedes aegypti</span> whole-body sample infected with Mayaro virus after <span class="html-italic">Aedes aegypti</span> blood feeding. Gray circles—mice serum samples; Red circles—mice brain samples; Green diamond—mice liver samples; Blue triangle—mice gonad samples; Black circles—mosquito body samples; ns—not significant. Statistical analysis was performed using Graph Pad Prism 9 (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">
14 pages, 1392 KiB  
Review
Elucidating the Implications of Norovirus N- and O-Glycosylation, O-GlcNAcylation, and Phosphorylation
by Chia-Chi Cheng, Guan-Ming Ke, Pei-Yu Chu and Liang-Yin Ke
Viruses 2023, 15(3), 798; https://doi.org/10.3390/v15030798 - 21 Mar 2023
Cited by 1 | Viewed by 2680
Abstract
Norovirus is the most common cause of foodborne gastroenteritis, affecting millions of people worldwide annually. Among the ten genotypes (GI–GX) of norovirus, only GI, GII, GIV, GVIII, and GIX infect humans. Some genotypes reportedly exhibit post-translational modifications (PTMs), including N- and O [...] Read more.
Norovirus is the most common cause of foodborne gastroenteritis, affecting millions of people worldwide annually. Among the ten genotypes (GI–GX) of norovirus, only GI, GII, GIV, GVIII, and GIX infect humans. Some genotypes reportedly exhibit post-translational modifications (PTMs), including N- and O-glycosylation, O-GlcNAcylation, and phosphorylation, in their viral antigens. PTMs have been linked to increased viral genome replication, viral particle release, and virulence. Owing to breakthroughs in mass spectrometry (MS) technologies, more PTMs have been discovered in recent years and have contributed significantly to preventing and treating infectious diseases. However, the mechanisms by which PTMs act on noroviruses remain poorly understood. In this section, we outline the current knowledge of the three common types of PTM and investigate their impact on norovirus pathogenesis. Moreover, we summarize the strategies and techniques for the identification of PTMs. Full article
(This article belongs to the Special Issue Norovirus and Foodborne Diseases)
Show Figures

Figure 1

Figure 1
<p>Replication cycle of noroviruses and protein post-translational modifications. (1) Attachment: Human norovirus (HuNoV) attaches to the HBGAs on the host cell surface, allowing viral entrance. Note that fucosylation of HBGA by fucosyltransferase 2 (FUT2) is required for certain genotypes of norovirus infection. In contrast, deamination on Asn373 of norovirus capsid protein VP1 impairs the recognition of HBGAs. (2) Binding. (3) Uncoating through undefined pathways [<a href="#B80-viruses-15-00798" class="html-bibr">80</a>]. (4) Translation: The positive-sense RNA genome may serve as a template for viral protein translation. After translation, viral proteins could undergo <span class="html-italic">N</span>- or <span class="html-italic">O</span>-glycosylation at the endoplasmic reticulum or Golgi apparatus. Mechanisms of noroviral protein glycosylation remain unclear. (5) Assembly: viral proteins assemble to form new viral particles. (6) Release: viral particle release from host cell [<a href="#B81-viruses-15-00798" class="html-bibr">81</a>]. Abbreviations: HBGA, histo-blood group antigen.</p>
Full article ">Figure 2
<p>Replication cycle of noroviruses and post-translational modification crosstalk between phosphorylation and <span class="html-italic">O</span>-GlcNAcylation. Co-occurring PTMs on proteins are common, and PTM crosstalk between phosphorylation and <span class="html-italic">O</span>-GlcNAcylation is the most common since they will compete for the same residues (serine/threonine residues). Phosphorylation is required for viral function. In contrast, <span class="html-italic">O</span>-GlcNAcylation of norovirus capsid protein VP1 could interfere with receptor binding. Studies on the <span class="html-italic">O</span>-GlcNAcylation of noroviruses are few and lack direct evidence from receptor binding assays. Abbreviations: HBGA, histo-blood group antigen; VP1, major capsid protein VP1; ATP, adenosine triphosphate; ADP, adenosine diphosphate; OGT, <span class="html-italic">O</span>-linked <span class="html-italic">N</span>-acetylglucosamine (GlcNAc) transferase; OGA, <span class="html-italic">O</span>-GlcNAcase; <span class="html-italic">O</span>-GlcNAc, <span class="html-italic">O</span>-linked <span class="html-italic">N</span>-acetylglucosamine; UDP-GlcNAc, uridine diphosphate <span class="html-italic">N</span>-acetylglucosamine.</p>
Full article ">Figure 3
<p>PTM crosstalk between <span class="html-italic">O</span>-GlcNAcylation and phosphorylation. Phosphorylation occurs when protein kinase attaches a phosphate group from ATP to the substrate; in contrast, PPase removes a phosphate group from a phosphoprotein. Similarly, OGT adds <span class="html-italic">O</span>-GlcNAc from UDP-GlcNAc to the substrates for <span class="html-italic">O</span>-GlcNAcylation. Conversely, OGA removes <span class="html-italic">O</span>-GlcNAc. These two processes are both reversible. Because phosphorylation and <span class="html-italic">O</span>-GlcNAcylation can compete for serine and threonine residues, PTM crosstalk occurs [<a href="#B83-viruses-15-00798" class="html-bibr">83</a>]. Abbreviations: <span class="html-italic">O</span>-GlcNAc, <span class="html-italic">O</span>-linked β-<span class="html-italic">N</span>-acetylglucosamine; OGT, <span class="html-italic">O</span>-GlcNAc transferase; OGA, <span class="html-italic">O</span>-GlcNAcase; Ser, serine; Thr, threonine; ATP, adenosine triphosphate; ADT, adenosine diphosphate; UPD, uridine diphosphate; PPase, protein phosphatase.</p>
Full article ">
22 pages, 1977 KiB  
Review
B and T Cell Epitopes of the Incursionary Foot-and-Mouth Disease Virus Serotype SAT2 for Vaccine Development
by Qian Li, Ashenafi Kiros Wubshet, Yang Wang, Livio Heath and Jie Zhang
Viruses 2023, 15(3), 797; https://doi.org/10.3390/v15030797 - 21 Mar 2023
Cited by 7 | Viewed by 3247
Abstract
Failure of cross-protection among interserotypes and intratypes of foot-and-mouth disease virus (FMDV) is a big threat to endemic countries and their prevention and control strategies. However, insights into practices relating to the development of a multi-epitope vaccine appear as a best alternative approach [...] Read more.
Failure of cross-protection among interserotypes and intratypes of foot-and-mouth disease virus (FMDV) is a big threat to endemic countries and their prevention and control strategies. However, insights into practices relating to the development of a multi-epitope vaccine appear as a best alternative approach to alleviate the cross-protection-associated problems. In order to facilitate the development of such a vaccine design approach, identification and prediction of the antigenic B and T cell epitopes along with determining the level of immunogenicity are essential bioinformatics steps. These steps are well applied in Eurasian serotypes, but very rare in South African Territories (SAT) Types, particularly in serotype SAT2. For this reason, the available scattered immunogenic information on SAT2 epitopes needs to be organized and clearly understood. Therefore, in this review, we compiled relevant bioinformatic reports about B and T cell epitopes of the incursionary SAT2 FMDV and the promising experimental demonstrations of such designed and developed vaccines against this serotype. Full article
(This article belongs to the Special Issue Foot-and-Mouth Disease Virus and Other Vesicular Disease Viruses)
Show Figures

Figure 1

Figure 1
<p>This simple schematic diagram shows the genome organization of FMDV, the encoded polyprotein, and matured proteins. It describes the long process of protein cleavage. The single ORF is illustrated in the box, and the viral proteins are named according to Rueckert and Wimmer’s nomenclature of picornavirus proteins [<a href="#B18-viruses-15-00797" class="html-bibr">18</a>]. Inside the boxes, there is a leader protein (L pro), four structural viral proteins (VP1-4), and seven non-structural proteins (2A-C, 3A-D). In addition, cleaved protein products and the cascade of cleavage are also shown in the diagram. The mature functional protein elements after cleavage are categorized as structural and non-structural proteins. The main cleavage sites are also shown in the box at the left side. The right and left untranslated regions (UTR) of the open reading frame are 3’ UTR and 5’ UTR respectively. The right extreme flank is shorter than the left one.</p>
Full article ">Figure 2
<p>Incursionary features of FMDV serotype SAT2. In this map, SAT2 topotype VII was taken as an example. The directional arrows on the map show the epidemiological dynamics of SAT2 across different African regions, jumping the WOAH demarcated pools.</p>
Full article ">Figure 3
<p>The illustration describes successive steps in epitope-based vaccine designing (this figure was adopted from Aryandra Arya and Sunil K Arora [<a href="#B103-viruses-15-00797" class="html-bibr">103</a>]. (1) In vitro antigenic epitope selection, identification, and analysis by using multiple immunoinformatic software from proteomic databases; (2) immunogenic B and T cell epitope selection by epitope insertion and sequence analysis; (3) synthesis of the designed neutralizing epitopes in the form of particulate, and evaluation of humoral and cellular immune responses by in vitro testing; (4) In vivo animal immunization to analyze the antibody responses to functionally characterize the anticipated neutralizing Abs- and cell-mediated immune responses.</p>
Full article ">Figure 4
<p>Adjuvants targeting both innate and adaptive immunity and the mechanism action. (<b>A</b>) Innate immunity; molecular adjuvants (cytokines, chemokines, and dendritic stimulating molecules) incorporated as a gene in the same plasmid and, when co-delivered with the vaccine, can elicit efficient DC recruitment, activation, and Ag presentation to T cells. The CD8+ T lymphocytes are authorized by dendritic cells to become effector CTLs. For this, antigens need to be taken up, processed, and presented by dendritic cells (DCs) in association with MHC molecules. Cytokine adjuvants help sustain CD25+ regulatory T cells (T-regs) of the CD4+ cells. This can significantly improve the immune response. (<b>B</b>) Vaccine with adjuvants, such as oil-in-water emulsions and almunium hydroxide, can stimulate the adaptive arm of the immune system. The maturation of DCs are firstly recognized by CD4+ and CD8+ T cells. CD4+ T lymphocytes undergo a clonal expansion into two distinct T-helper subpopulations following contact with the MHC I-peptide complex. TH1 and TH2 cells stimulate the humoral and cell-mediated immune response through a different cytokine expression pattern. TLR4 agonists are considered to be the main ligand to activate DCs. This kind of adjuvant elicits a TH2-type functional immune response with preferential IgG1 antibody production [<a href="#B157-viruses-15-00797" class="html-bibr">157</a>,<a href="#B158-viruses-15-00797" class="html-bibr">158</a>,<a href="#B159-viruses-15-00797" class="html-bibr">159</a>]. CTL: cytotoxic T lymphocyte; DC: dendritic cell; PAMP: pathogen-associated molecular pattern; PRR: pattern recognition receptor; TLR: toll-like receptor (TLR is a receptor family belonging to PRRs); Treg: regulatory T cell; Th: T helper cell.</p>
Full article ">
13 pages, 1954 KiB  
Article
Antibody Immunity to Zika Virus among Young Children in a Flavivirus-Endemic Area in Nicaragua
by Omar Zepeda, Daniel O. Espinoza, Evelin Martinez, Kaitlyn A. Cross, Sylvia Becker-Dreps, Aravinda M. de Silva, Natalie M. Bowman, Lakshmanane Premkumar, Elizabeth M. Stringer, Filemón Bucardo and Matthew H. Collins
Viruses 2023, 15(3), 796; https://doi.org/10.3390/v15030796 - 21 Mar 2023
Cited by 2 | Viewed by 2600
Abstract
Objective: To understand the dynamics of Zika virus (ZIKV)-specific antibody immunity in children born to mothers in a flavivirus-endemic region during and after the emergence of ZIKV in the Americas. Methods: We performed serologic testing for ZIKV cross-reactive and type-specific IgG in two [...] Read more.
Objective: To understand the dynamics of Zika virus (ZIKV)-specific antibody immunity in children born to mothers in a flavivirus-endemic region during and after the emergence of ZIKV in the Americas. Methods: We performed serologic testing for ZIKV cross-reactive and type-specific IgG in two longitudinal cohorts, which enrolled pregnant women and their children (PW1 and PW2) after the beginning of the ZIKV epidemic in Nicaragua. Quarterly samples from children over their first two years of life and maternal blood samples at birth and at the end of the two-year follow-up period were studied. Results: Most mothers in this dengue-endemic area were flavivirus-immune at enrollment. ZIKV-specific IgG (anti-ZIKV EDIII IgG) was detected in 82 of 102 (80.4%) mothers in cohort PW1 and 89 of 134 (66.4%) mothers in cohort PW2, consistent with extensive transmission observed in Nicaragua during 2016. ZIKV-reactive IgG decayed to undetectable levels by 6–9 months in infants, whereas these antibodies were maintained in mothers at the year two time point. Interestingly, a greater contribution to ZIKV immunity by IgG3 was observed in babies born soon after ZIKV transmission. Finally, 43 of 343 (13%) children exhibited persistent or increasing ZIKV-reactive IgG at ≥9 months, with 10 of 30 (33%) tested demonstrating serologic evidence of incident dengue infection. Conclusions: These data inform our understanding of protective and pathogenic immunity to potential flavivirus infections in early life in areas where multiple flaviviruses co-circulate, particularly considering the immune interactions between ZIKV and dengue and the future possibility of ZIKV vaccination in women of childbearing potential. This study also shows the benefits of cord blood sampling for serologic surveillance of infectious diseases in resource-limited settings. Full article
(This article belongs to the Section Human Virology and Viral Diseases)
Show Figures

Figure 1

Figure 1
<p>Seroprevalence of ZIKV-reactive IgG in cord blood. ZIKV IgG (<b>A</b>) and ZIKV EDIII (<b>B</b>) ELISA were performed on cord blood plasma diluted 1:200 from Cohort 1 (<span class="html-italic">n</span> = 104) and Cohort 2 (<span class="html-italic">n</span> = 135). The cutoff for positivity was determined on each plate by running negative control plasma. The percent seropositivity for each assay is shown beneath the cohort names, and the geometric mean (GM) of OD for positive samples is shown beneath that. (<b>C</b>) The OD value from the ZIKV EDIII ELISA for available paired samples (<span class="html-italic">n</span> = 102) from mother peripheral blood and cord blood are graphed in a scatter plot to assess the efficiency of ZIKV-reactive IgG transfer across the placenta. Pearson correlation analysis was performed, and the R<sup>2</sup> and p values are included in the upper left corner of each graph. ns, not significant; ****, <span class="html-italic">p</span> &lt; 0.0001 according to an unpaired Student’s t-test comparing differences between GM of Cohort 1 vs. Cohort 2.</p>
Full article ">Figure 2
<p>ZIKV-reactive antibody kinetics and IgG subtype. Decay of IgG reactive to ZIKV (<b>A</b>) and ZIKV EDIII (<b>B</b>) is shown for infants in both PW1 and PW2 cohorts. Spaghetti plots of trajectories were generated from the OD from each ELISA as a ratio to the LLOD, using a continuous time scale of weeks of life for each infant. Data from individual infants are shown as semi-transparent lines. Fit lines summarizing all data were calculated using an exponential decay (red dashed line) and a quadratic polynomial (blue dashed line) model. (<b>C</b>) IgG3 reactivity to ZIKV (left panel) and ZIKV EDIII (right panel) is shown in a subset of cord blood samples (<span class="html-italic">n</span> = 10 per subgroup) from subgroups indicated on the <span class="html-italic">x</span>-axis. OD, optical density; LLOD, lower limit of detection. X-axis legend of left graph of panel C: needs space between LOW and EDIII in the 2 group from left.</p>
Full article ">Figure 3
<p>Incident flavivirus infection detected after initial decay of maternal-derived anti-ZIKV IgG in early life. Paired samples available from time points before and after the observed increase in anti-ZIKV IgG were selected for NAb testing. (<b>A</b>) Representative raw FRNT data measuring NAb to ZIKV (left) and DENV2 (right) are shown for three subjects. (<b>B</b>,<b>C</b>) eFRNT50 values for ZIKV (left) and DENV2 (right) are shown for suspected cases of incident flavivirus infection in Cohort 1 (<b>B</b>) and Cohort 2 (<b>C</b>). For nine samples with suspected incident flavivirus infection., and eFRNT50 values for paired samples are linked by a solid line.</p>
Full article ">
14 pages, 2788 KiB  
Article
Tissue and Time Optimization for Real-Time Detection of Apple Mosaic Virus and Apple Necrotic Mosaic Virus Associated with Mosaic Disease of Apple (Malus domestica)
by Sajad Un Nabi, Javid Iqbal Mir, Salwee Yasmin, Ambreena Din, Wasim H. Raja, G. S. Madhu, Shugufta Parveen, Sheikh Mansoor, Yong Suk Chung, Om Chand Sharma, Muneer Ahmad Sheikh, Fahad A. Al-Misned and Hamed A. El-Serehy
Viruses 2023, 15(3), 795; https://doi.org/10.3390/v15030795 - 21 Mar 2023
Cited by 5 | Viewed by 2379
Abstract
Besides apple mosaic virus (ApMV), apple necrotic mosaic virus (ApNMV) has also been found to be associated with apple mosaic disease. Both viruses are unevenly distributed throughout the plant and their titer decreases variably with high temperatures, hence requiring proper tissue and time [...] Read more.
Besides apple mosaic virus (ApMV), apple necrotic mosaic virus (ApNMV) has also been found to be associated with apple mosaic disease. Both viruses are unevenly distributed throughout the plant and their titer decreases variably with high temperatures, hence requiring proper tissue and time for early and real-time detection within plants. The present study was carried out to understand the distribution and titer of ApMV and ApNMV in apple trees from different plant parts (spatial) during different seasons (temporal) for the optimization of tissue and time for their timely detection. The Reverse Transcription-Polymerase Chain Reaction (RT-PCR) and Reverse Transcription-quantitative Polymerase Chain Reaction (RT-qPCR) was carried out to detect and quantify both viruses in the various plant parts of apple trees during different seasons. Depending on the availability of tissue, both ApMV and ApNMV were detected in all the plant parts during the spring season using RT-PCR. During the summer, both viruses were detected only in seeds and fruits, whereas they were detected in leaves and pedicel during the autumn season. The RT-qPCR results showed that during the spring, the ApMV and ApNMV expression was higher in leaves, whereas in the summer and autumn, the titer was mostly detected in seeds and leaves, respectively. The leaves in the spring and autumn seasons and the seeds in the summer season can be used as detection tissues through RT-PCR for early and rapid detection of ApMV and ApNMV. This study was validated on 7 cultivars of apples infected with both viruses. This will help to accurately sample and index the planting material well ahead of time, which will aid in the production of virus-free, quality planting material. Full article
(This article belongs to the Special Issue Plant Virus Epidemiology and Control 2022)
Show Figures

Figure 1

Figure 1
<p>Different plant parts/tissues with presence or absence of symptoms noted for virus detection and quantification. (<b>a</b>) Symptomatic leaf and pedicel, (<b>b</b>) petals, (<b>c</b>) anther, (<b>d</b>) flowers without petal, (<b>e</b>) bark, (<b>f</b>) fruit and (<b>g</b>) seed.</p>
Full article ">Figure 2
<p>The symptoms of mosaic and necrotic mosaic on two cultivars: Golden Delicious (<b>a</b>) and Oregon Spur (<b>b</b>).</p>
Full article ">Figure 3
<p>Amplicon of 260 bp amplified from ApMV-infected tissues of cultivar Golden Delicious (GD) plant tissues by conventional RT-PCR during the spring season (<b>a</b>), summer season (<b>b</b>) and autumn season (<b>c</b>), where L: ladder 100 bp, LF: Leaf, FL: Flowers, GD4:Bark, PT: Petals, PD: Pedicel, BK: Bark, FR: Fruit, SD: Seed, NC: Negative control, and PC: Positive control.</p>
Full article ">Figure 4
<p>Amplicon of 670 bp amplified from ApNMV-infected tissues of cultivar Golden Delicious (GD) plant tissues by conventional RT-PCR during the spring season (<b>a</b>), summer season (<b>b</b>) and autumn season (<b>c</b>), where L: ladder 1 kb, LF: Leaf, FL: Flowers, GD4:Bark, PT: Petals, PD: Pedicel, BK: Bark, FR: Fruit, SD: Seed, NC: Negative control, and PC: Positive control.</p>
Full article ">Figure 5
<p>Comparative fold change ofthe CP gene of ApMV and ApNMV in different tissues in cultivar Golden Delicious (GD) and Oregon spur (OS) during the spring season. <span class="html-italic">Y</span>-axis represents the fold change in the viral infection, whereas <span class="html-italic">X</span>-axis represents different tissues tested for both viruses.</p>
Full article ">Figure 6
<p>Comparative fold change of the CP gene of ApMV and ApNMV in different tissues in cultivar Golden Delicious (GD) and Oregon spur (OS) during the summer season. <span class="html-italic">Y</span>-axis represents the fold change in the virus infection, whereas <span class="html-italic">X</span>-axis represents different tissues tested for both viruses.</p>
Full article ">Figure 7
<p>Comparative fold change of the CP gene of APMV and ApNMV in different tissues in cultivar Golden Delicious (GD) and Oregon spur (OS) during the autumn season. <span class="html-italic">Y</span>-axis represents the fold change in the virus infection, whereas <span class="html-italic">X</span>-axis represents different tissues tested for both viruses.</p>
Full article ">
23 pages, 4901 KiB  
Article
Circulating Plasma Exosomal Proteins of Either SHIV-Infected Rhesus Macaque or HIV-Infected Patient Indicates a Link to Neuropathogenesis
by Partha K. Chandra, Stephen E. Braun, Sudipa Maity, Jorge A. Castorena-Gonzalez, Hogyoung Kim, Jeffrey G. Shaffer, Sinisa Cikic, Ibolya Rutkai, Jia Fan, Jessie J. Guidry, David K. Worthylake, Chenzhong Li, Asim B. Abdel-Mageed and David W. Busija
Viruses 2023, 15(3), 794; https://doi.org/10.3390/v15030794 - 21 Mar 2023
Cited by 2 | Viewed by 2787
Abstract
Despite the suppression of human immunodeficiency virus (HIV) replication by combined antiretroviral therapy (cART), 50–60% of HIV-infected patients suffer from HIV-associated neurocognitive disorders (HAND). Studies are uncovering the role of extracellular vesicles (EVs), especially exosomes, in the central nervous system (CNS) due to [...] Read more.
Despite the suppression of human immunodeficiency virus (HIV) replication by combined antiretroviral therapy (cART), 50–60% of HIV-infected patients suffer from HIV-associated neurocognitive disorders (HAND). Studies are uncovering the role of extracellular vesicles (EVs), especially exosomes, in the central nervous system (CNS) due to HIV infection. We investigated links among circulating plasma exosomal (crExo) proteins and neuropathogenesis in simian/human immunodeficiency virus (SHIV)-infected rhesus macaques (RM) and HIV-infected and cART treated patients (Patient-Exo). Isolated EVs from SHIV-infected (SHIV-Exo) and uninfected (CTL-Exo) RM were predominantly exosomes (particle size < 150 nm). Proteomic analysis quantified 5654 proteins, of which 236 proteins (~4%) were significantly, differentially expressed (DE) between SHIV-/CTL-Exo. Interestingly, different CNS cell specific markers were abundantly expressed in crExo. Proteins involved in latent viral reactivation, neuroinflammation, neuropathology-associated interactive as well as signaling molecules were expressed at significantly higher levels in SHIV-Exo than CTL-Exo. However, proteins involved in mitochondrial biogenesis, ATP production, autophagy, endocytosis, exocytosis, and cytoskeleton organization were significantly less expressed in SHIV-Exo than CTL-Exo. Interestingly, proteins involved in oxidative stress, mitochondrial biogenesis, ATP production, and autophagy were significantly downregulated in primary human brain microvascular endothelial cells exposed with HIV+/cART+ Patient-Exo. We showed that Patient-Exo significantly increased blood–brain barrier permeability, possibly due to loss of platelet endothelial cell adhesion molecule-1 protein and actin cytoskeleton structure. Our novel findings suggest that circulating exosomal proteins expressed CNS cell markers—possibly associated with viral reactivation and neuropathogenesis—that may elucidate the etiology of HAND. Full article
(This article belongs to the Special Issue Viruses and Extracellular Vesicles 2023)
Show Figures

Figure 1

Figure 1
<p>Characterization of circulating plasma exosomal proteome of SHIV-infected and uninfected Rhesus macaque (RM). (<b>a</b>,<b>b</b>) Exosomes were isolated by exoEasy Maxi Kit from the plasma of SHIV-infected and uninfected RM. The size (nm) and concentration (particles/mL) of the isolated exosomes were characterized by ZetaView Particle Metrix system. (<b>c</b>) Total and significantly differentially expressed (DE) proteins in plasma exosomes of SHIV-infected (SHIV-Exo) and control (CTL-Exo) RM quantified by proteomic analysis. (<b>d</b>) Comparison of the number of unique peptide(s) detected per quantified proteins. (<b>e</b>) Comparison of the number of unique peptide(s) detected in significant DE proteins. (<b>f</b>) A hierarchical cluster analysis of significant DE proteins. A heatmap was generated for all the 236 significant DE proteins using “Manhattan” clustering and “complete” linkage method.</p>
Full article ">Figure 2
<p>(<b>a</b>,<b>b</b>) A hierarchical cluster analysis of significant top 50 up-/down-regulated proteins in SHIV-Exo. Proteins were filtered based on <span class="html-italic">p</span> &lt; 0.05. Median values were calculated for CTL-Exo and SHIV-Exo to calculate the ratio (SHIV-/CTL-Exo). The ratio was used to generate the list of top 50 up and down regulated proteins. Z scores were calculated for each significant protein, and then heatmaps were generated for all the significant proteins: top 50 upregulated and top 50 downregulated proteins using “Manhattan” clustering and “complete” linkage method. Protein coding gene names are presented here. (<b>c</b>) To test if the dataset samples were separated, we performed a principal component analysis (PCA) analysis on all protein expression data and as observed the CTL-Exo (in red) are clustered separately from the SHIV-Exo (in green). Unsupervised PCA plot is generated in R Studio by using median normalized and log<sub>10</sub> transformed data from each sample type. CTL-Exo: Plasma exosomes isolated from uninfected RM; SHIV-Exo: Plasma exosomes isolated from SHIV-infected RM.</p>
Full article ">Figure 3
<p>Hallmark exosomal proteins quantified by proteomic analysis in circulating exosomes of SHIV-infected and uninfected RM. (<b>a</b>–<b>e</b>) Relative abundance of quantified proteins in CTL-/SHIV-Exo. Proteins that exhibited group differences shown as bar graphs with green and red for CTL-Exo and SHIV-Exo, respectively. Plasma exosomes were isolated from three SHIV-infected and three uninfected RM (N = 3/group). Graphs show mean ± SD of relative abundance that was calculated from four experimental replicates of each sample. Protein group ‘b’ and ‘e’ passed the D’Agostino and Pearson normality test and protein group ‘c’ and ‘d’ passed the Shapiro–Wilk normality test. Protein group ‘a’ did not pass the normality test. The data sets were followed by unpaired t test with Welch correction, performed for normally distributed data. When the data did not pass the normality test, a non-parametric Mann–Whitney test was performed. Significant differences (<span class="html-italic">p</span> &lt; 0.05) between groups are indicated. (<b>f</b>) The expression of CD63, CD81, GAPDH, and FLOT1 in CTL-/SHIV-Exo was further validated by Western blotting (N = 5/group). The protein coding gene names are presented here and were also described in the text. CTL-Exo: Plasma exosomes isolated from uninfected RM; SHIV-Exo: Plasma exosomes isolated from SHIV-infected RM.</p>
Full article ">Figure 4
<p>Relative abundance of CNS cell markers in circulating plasma exosomes. (<b>a</b>) Central nervous system (CNS) cells-specific exosomal proteins in CTL-/SHIV-Exo were quantified by proteomic analysis. Proteins that exhibited group differences shown as bar graphs with white and gray for CTL-Exo and SHIV-Exo, respectively. Plasma exosomes were isolated from three SHIV-infected and three uninfected RM (N = 3/group). Graphs show mean ± SD of relative abundance calculated from four experimental replicates per sample. For a protein group not passing the normality test, a non-parametric Mann–Whitney test was performed. There were no significant differences (<span class="html-italic">p</span> &lt; 0.05) between groups. (<b>b</b>) The indicated proteins were further validated by Western blotting. For this assay, plasma exosomes were isolated from five SHIV-infected and five uninfected RM (N = 5/group). Protein coding gene names are presented here and were described in the text. CTL-Exo: Plasma exosomes isolated from uninfected RM; SHIV-Exo: Plasma exosomes isolated from SHIV-infected RM. BMVEC: Brain microvascular endothelial cells.</p>
Full article ">Figure 5
<p>Increased expression of proteins involved in viral reactivation, inflammation, and neuropathology-associated interactive/signaling proteins in SHIV-Exo. (<b>a</b>–<b>c</b>) Relative protein abundance in CTL-/SHIV-Exo was quantified by proteomic analysis. Proteins that exhibited group differences are shown as bar graphs with green and red for CTL-Exo and SHIV-Exo, respectively. Plasma exosomes were isolated from three SHIV-infected and three uninfected RM (N = 3/group). Graphs show mean ± SD relative abundance that was calculated from four experimental replicates of each sample. All protein groups (<b>a</b>–<b>c</b>) passed the Shapiro–Wilk normality test, and the data sets followed by unpaired t test with Welch correction for normally distributed data. The protein coding gene names are presented here, and they were described in the text. CTL-Exo: Plasma exosomes isolated from uninfected RM; SHIV-Exo: Plasma exosomes isolated from SHIV-infected RM. q-values for: (<b>a</b>) all of them are 0.035; (<b>b</b>) all of them are 0.040; and (<b>c</b>) 0.022, 0.036, 0.017, 0.022, 0.017. Age-adjusted <span class="html-italic">p</span>-values were presented in the <a href="#app1-viruses-15-00794" class="html-app">Supplementary Materials File S1</a>.</p>
Full article ">Figure 6
<p>Decreased expression of proteins involved in mitochondrial biogenesis and ATP production in SHIV-Exo. (<b>a</b>–<b>c</b>) Differential expression of mitochondrial proteins in CTL-/SHIV-Exo quantified by proteomic analysis. Proteins that exhibited group differences are shown as bar graphs with green and red for CTL-Exo and SHIV-Exo, respectively. Plasma exosomes were isolated from three SHIV-infected and three uninfected RM (N = 3/group). Graphs show mean ± SD of relative abundance calculated from four experimental replicates/sample. Protein group ‘a’ and ‘c’ passed the Shapiro–Wilk normality test and protein group ‘b’ passed the D’Agostino and Pearson normality test. The data sets were followed by unpaired t test with Welch correction for normally distributed data. The protein coding gene names are presented here, and they were described in the text. CTL-Exo: Plasma exosomes isolated from uninfected RM; SHIV-Exo: Plasma exosomes isolated from SHIV-infected RM. q-values for: (<b>a</b>) 0.022, 0.022, 0.039, 0.022; (<b>b</b>) 0.024, 0.037, 0.013, 0.024; and (<b>c</b>) all of them are 0.043. Age-adjusted <span class="html-italic">p</span>-values were presented in the <a href="#app1-viruses-15-00794" class="html-app">Supplementary Materials File S1</a>.</p>
Full article ">Figure 7
<p>Decreased expression of proteins involved in autophagy, endosomal recycling, exocytosis, and sprouting angiogenesis in SHIV-Exo. (<b>a</b>–<b>c</b>) Differential expression of proteins involved in autophagy, endosomal recycling, exocytosis, and sprouting angiogenesis in CTL-/SHIV-Exo quantified by proteomic analysis. Proteins exhibiting group differences are shown as bar graphs with green and red for CTL-Exo and SHIV-Exo, respectively. Graphs show mean ± SD of relative abundance calculated from four experimental replicates/sample. All protein groups passed the Shapiro–Wilk normality test. The data sets were followed by unpaired t test with Welch correction for normally distributed data. The protein coding gene names are presented here and were described in the text. CTL-Exo: Plasma exosomes isolated from uninfected RM; SHIV-Exo: Plasma exosomes isolated from SHIV-infected RM (N = 3/group). q-values for: (<b>a</b>) 0.036, 0.048, 0.041, 0.038; (<b>b</b>) all of them are 0.042; and (<b>c</b>) 0.037, 0.048, 0.037, 0.037, 0.048. Age-adjusted <span class="html-italic">p</span>-values were presented in the <a href="#app1-viruses-15-00794" class="html-app">Supplementary Materials File S1</a>.</p>
Full article ">Figure 8
<p>The effect of HIV-infected and cART-treated Patient-Exo on primary HBMVECs. (<b>a</b>) Cells were treated with 10 µg/mL of both Patient-Exo as well as exosomes isolated from healthy human plasma (hCTL-Exo) for 24 h. The indicated proteins were qualitatively detected by Western blotting from equal amount (15 µg) of clarified cell lysates and β-actin was used as an internal control. Three independent experiments were performed and are presented by Arabic numbers. (<b>b</b>) The relative band intensities of CAT, LC3B-II, pDRP1, and MC-III were compared by Image-J software (version 1.50). Values were mean ± standard deviation (SD). The data set passed the Shapiro–Wilk normality test and unpaired t test with Welch correction was performed for normally distributed data. q-values: all of them are 0.041 for HBMVEC only vs Patient-Exo; and 0.054, 0.043, 0.041, 0.041 for hCTL-Exo vs. Patient-Exo.</p>
Full article ">Figure 9
<p>In vitro BBB permeability assays and the detection of PECAM-1 by immunofluorescence confocal microscopy in primary HBMVECs exposed to HIV+ Patient-Exo. (<b>a</b>) Schematic diagram of transwell migration-based BBB permeability in vitro assay. (<b>b</b>) Graph showing the BBB permeability efficiency of fluorescence labeled high molecular weight dextran (FITC-Dextran, 150 kDa) in the absence (black line) and presence (red, blue, and green lines) of cell monolayer. Cells without treatment (Cells only) or treated with normal human plasma exosomes (+CTL-Exo) were used as controls. The increased BBB permeability in cells treated with HIV-infected and cART-treated patient plasma exosomes (+Patient-Exo) is indicated by the red line. N = 4 experimental replicates were performed. (<b>c</b>) Representative immunofluorescent confocal microscope images of HBMVEC expressing PECAM-1 after 24 h treatment (10 µg/mL) with either hCTL-Exo or Patient-Exo. (<b>d</b>) The relative fluorescence intensity of PECAM-1 per view field was compared.</p>
Full article ">Figure 10
<p>Differential expression of actin cytoskeleton proteins in RM plasma exosomes and the expression pattern of F-actin in primary HBMVECs after exposure to HIV+ Patient-Exo. (<b>a</b>) The heat map of the abundance actin cytoskeleton proteins quantified by proteomic analysis was generated by using GraphPad Prism version 9.0.0 for Windows. (<b>b</b>) Representative immunofluorescent confocal microscope images of the HBMVEC expressing F-actin after 24 h treatment (10 µg/mL) with either hCTL-Exo or Patient-Exo.</p>
Full article ">
11 pages, 1607 KiB  
Communication
Correlation of SARS-CoV-2 Neutralization with Antibody Levels in Vaccinated Individuals
by Shazeda Haque Chowdhury, Sean Riley, Riley Mikolajczyk, Lauren Smith, Lakshmanan Suresh and Amy Jacobs
Viruses 2023, 15(3), 793; https://doi.org/10.3390/v15030793 - 21 Mar 2023
Cited by 1 | Viewed by 1858
Abstract
Neutralizing antibody titers are an important measurement of the effectiveness of vaccination against SARS-CoV-2. Our laboratory has set out to further verify the functionality of these antibodies by measuring the neutralization capacity of patient samples against infectious SARS-CoV-2. Samples from patients from Western [...] Read more.
Neutralizing antibody titers are an important measurement of the effectiveness of vaccination against SARS-CoV-2. Our laboratory has set out to further verify the functionality of these antibodies by measuring the neutralization capacity of patient samples against infectious SARS-CoV-2. Samples from patients from Western New York who had been vaccinated with the original Moderna and Pfizer vaccines (two doses) were tested for neutralization of both Delta (B.1.617.2) and Omicron (BA.5). Strong correlations between antibody levels and neutralization of the delta variant were attained; however, antibodies from the first two doses of the vaccines did not have good neutralization coverage of the subvariant omicron BA.5. Further studies are ongoing with local patient samples to determine correlation following updated booster administration. Full article
(This article belongs to the Special Issue RNA Viruses and Antibody Response)
Show Figures

Figure 1

Figure 1
<p>Example of plaque reduction neutralization performed in duplicate. Overlay fixed with 4% paraformaldehyde and stained with crystal violet before plaques were counted.</p>
Full article ">Figure 2
<p>Representative PRNT50 analysis performed according to Bewley et al. [<a href="#B8-viruses-15-00793" class="html-bibr">8</a>].</p>
Full article ">Figure 3
<p>Correlation of PRNT50 titers with IgG COI when challenged with Delta and Omicron variants of SARS-CoV-2. Viral load of 6 × 10<sup>3</sup> PFU/mL (30 plaques/well) were added to Vero-E6 cells. Data represented mean ± SD of two biological replicates. µ = 0.05.</p>
Full article ">Figure 4
<p>Correlation of PRNT50 titers with age when challenged with Delta and Omicron variants of SARS-CoV-2. Viral load of 6 × 10<sup>3</sup> PFU/mL (30 plaques/well) were added to Vero-E6 cells. Data represented mean ± SD of two biological replicates. µ = 0.05.</p>
Full article ">Figure 5
<p>Comparison of PRNT50 titers obtained following vaccination with the primary doses of the Moderna or Pfizer/BioNTech mRNA vaccines. Viral load of 6 × 10<sup>3</sup> PFU/mL (30 plaques/well) were added to Vero-E6 cells. Data represented median ± 95% CI of two biological replicates.</p>
Full article ">
8 pages, 2600 KiB  
Communication
Importance of Cellular Immunity and IFN-γ Concentration in Preventing SARS-CoV-2 Infection and Reinfection: A Cohort Study
by Dragan Primorac, Petar Brlek, Eduard Stjepan Pavelić, Jana Mešić, David Glavaš Weinberger, Vid Matišić, Vilim Molnar, Saša Srića and Renata Zadro
Viruses 2023, 15(3), 792; https://doi.org/10.3390/v15030792 - 20 Mar 2023
Cited by 5 | Viewed by 2247
Abstract
Recent studies have highlighted the underestimated importance of the cellular immune response after the emergence of variants of concern (VOCs) of SARS-CoV-2, and the significantly reduced neutralizing power of antibody titers in individuals with previous SARS-CoV-2 infection or vaccination. Our study included 303 [...] Read more.
Recent studies have highlighted the underestimated importance of the cellular immune response after the emergence of variants of concern (VOCs) of SARS-CoV-2, and the significantly reduced neutralizing power of antibody titers in individuals with previous SARS-CoV-2 infection or vaccination. Our study included 303 participants who were tested at St. Catherine Specialty Hospital using the Quan-T-Cell SARS-CoV-2 in combination with the Quan-T-Cell ELISA (Euroimmun Medizinische Labordiagnostika, Lübeck, Germany) for the analysis of IFN-γ concentration, and with Anti-SARS-CoV-2 QuantiVac ELISA IgG (Euroimmun Medizinische Labordiagnostika, Lübeck, Germany) for the detection of human antibodies of the immunoglobulin class IgG against the S1 domain of the SARS-CoV-2 spike protein. The statistical analysis showed a significant difference in the concentration of IFN-γ between reinfected participants and those without infection (p = 0.012). Participants who were not infected or reinfected with SARS-CoV-2 after vaccination and/or previous SARS-CoV-2 infection had a significantly higher level of cellular immunity. Furthermore, in individuals without additional vaccination, those who experienced infection/reinfection had significantly lower levels of IFN-γ compared to uninfected participants (p = 0.016). Our findings suggest a long-lasting effect of cellular immunity, measured by IFN-γ concentrations, which plays a key role in preventing infections and reinfections after the emergence of SARS-CoV-2 variants of concern. Full article
Show Figures

Figure 1

Figure 1
<p>Level of cellular immunity (IFN-γ concentration) in patients with a SARS-CoV-2 infection or reinfection and additional vaccination after initial testing of IFN-γ. Group A—participants with (re)infection and additional vaccination; Group B—participants with (re)infection without additional vaccination; Group C—participants without (re)infection with additional vaccination; Group D—participants without (re)infection and additional vaccination. Mean values are shown above each bar. *—Mann–Whitney, <span class="html-italic">p</span> = 0.016.</p>
Full article ">Figure 2
<p>Percentage of study participants with and without a SARS-CoV-2 infection or reinfection (after IFN-γ test) in participants with previous COVID-19 (before IFN-γ test) (group 1), participants vaccinated with one of the SARS-CoV-2 vaccines (group 2), participants who had both the SARS-CoV-2 infection and vaccination history (group 3), and patients without a history of SARS-CoV-2 infection or vaccination (group 4). Groups 2 and 4 have a higher percentage of participants who were infected with SARS-CoV-2, while groups 1 and 3 have a higher percentage of participants who did not have a SARS-CoV-2 reinfection.</p>
Full article ">
35 pages, 1395 KiB  
Article
First Expert Elicitation of Knowledge on Possible Drivers of Observed Increasing Human Cases of Tick-Borne Encephalitis in Europe
by Claude Saegerman, Marie-France Humblet, Marc Leandri, Gaëlle Gonzalez, Paul Heyman, Hein Sprong, Monique L’Hostis, Sara Moutailler, Sarah I. Bonnet, Nadia Haddad, Nathalie Boulanger, Stephen L. Leib, Thierry Hoch, Etienne Thiry, Laure Bournez, Jana Kerlik, Aurélie Velay, Solveig Jore, Elsa Jourdain, Emmanuelle Gilot-Fromont, Katharina Brugger, Julia Geller, Marie Studahl, Nataša Knap, Tatjana Avšič-Županc, Daniel Růžek, Tizza P. Zomer, René Bødker, Thomas F. H. Berger, Sandra Martin-Latil, Nick De Regge, Alice Raffetin, Sandrine A. Lacour, Matthias Klein, Tinne Lernout, Elsa Quillery, Zdeněk Hubálek, Francisco Ruiz-Fons, Agustín Estrada-Peña, Philippe Fravalo, Pauline Kooh, Florence Etore, Céline M. Gossner and Bethan Purseadd Show full author list remove Hide full author list
Viruses 2023, 15(3), 791; https://doi.org/10.3390/v15030791 - 20 Mar 2023
Cited by 11 | Viewed by 4763
Abstract
Tick-borne encephalitis (TBE) is a viral disease endemic in Eurasia. The virus is mainly transmitted to humans via ticks and occasionally via the consumption of unpasteurized milk products. The European Centre for Disease Prevention and Control reported an increase in TBE incidence over [...] Read more.
Tick-borne encephalitis (TBE) is a viral disease endemic in Eurasia. The virus is mainly transmitted to humans via ticks and occasionally via the consumption of unpasteurized milk products. The European Centre for Disease Prevention and Control reported an increase in TBE incidence over the past years in Europe as well as the emergence of the disease in new areas. To better understand this phenomenon, we investigated the drivers of TBE emergence and increase in incidence in humans through an expert knowledge elicitation. We listed 59 possible drivers grouped in eight domains and elicited forty European experts to: (i) allocate a score per driver, (ii) weight this score within each domain, and (iii) weight the different domains and attribute an uncertainty level per domain. An overall weighted score per driver was calculated, and drivers with comparable scores were grouped into three terminal nodes using a regression tree analysis. The drivers with the highest scores were: (i) changes in human behavior/activities; (ii) changes in eating habits or consumer demand; (iii) changes in the landscape; (iv) influence of humidity on the survival and transmission of the pathogen; (v) difficulty to control reservoir(s) and/or vector(s); (vi) influence of temperature on virus survival and transmission; (vii) number of wildlife compartments/groups acting as reservoirs or amplifying hosts; (viii) increase of autochthonous wild mammals; and (ix) number of tick species vectors and their distribution. Our results support researchers in prioritizing studies targeting the most relevant drivers of emergence and increasing TBE incidence. Full article
(This article belongs to the Special Issue Zoonotic Viral Diseases: Drivers, Causes, Prevention and Cure)
Show Figures

Figure 1

Figure 1
<p>Boxplot of the relative importance of the eight domains of possible drivers of the observed emergence or increasing incidence of TBE in humans (N = 40 European experts). Legend: The bold line represents the median of the score distribution between the different experts attributed to each domain; the solid lines at the top and bottom of each rectangle represent, respectively, the first and third quartiles; adjacent lines to the whiskers represent the limits of the 95% confidence interval; small circles represent outside values. The eight domains of drivers are: D1, disease/pathogen characteristics; D2, distance to Europe and the country of the expert (spatial-temporal scales); D3, ability to monitor, treat, and control the disease; D4, European farm characteristics; D5, global change; D6, wildlife interface; D7, human activity; and D8, economic and trade activities.</p>
Full article ">Figure 2
<p>Ranking of the overall weighted score for each potential driver of the observed emergence or increasing incidence of TBE in humans (boxplot based on input from 40 European experts). Legend: The x-axis represents the drivers with the following codification: D1 to D8 refer to the eight domains of drivers, and D1_1 to D8_11 refer to a specific driver (for the codification, see <a href="#viruses-15-00791-t0A1" class="html-table">Table A1</a>). A relation to <a href="#viruses-15-00791-f003" class="html-fig">Figure 3</a> was provided by the group, which named “very high importance”, “high importance” and “less importance”, as possible drivers of the observed emergence or increasing incidence of TBE in humans.</p>
Full article ">Figure 3
<p>Aggregation of drivers of the observed emergence or increasing incidence of TBE in humans into three homogenous groups using a regression tree analysis. Legend: N, number; SD, standard deviation. D1-01 to D8-11 refer to a specific driver (for the codification, see <a href="#viruses-15-00791-t0A1" class="html-table">Table A1</a>).</p>
Full article ">Figure 4
<p>Level of uncertainty per domain of drivers. Legend: The bold line represents the median of the level of uncertainty attributed by experts using a scale from 0 (minimal uncertainty in the scoring) to 100 (maximum uncertainty in the scoring); the solid lines at the top and bottom of each rectangle represent, respectively, the first and third quartiles; adjacent lines to the whiskers represent the limits of the 95% confidence interval; small circles represent outside values. The eight domains of drivers are: D1, disease/pathogen characteristics; D2, distance to Europe and the country of the expert (spatial-temporal scales); D3, ability to monitor, treat, and control the disease; D4, European farm characteristics; D5, global change; D6, wildlife interface; D7, human activity; and D8, economic and trade activities.</p>
Full article ">
24 pages, 3139 KiB  
Article
One Health Surveillance Highlights Circulation of Viruses with Zoonotic Potential in Bats, Pigs, and Humans in Viet Nam
by Alice Latinne, Nguyen Thi Thanh Nga, Nguyen Van Long, Pham Thi Bich Ngoc, Hoang Bich Thuy, PREDICT Consortium, Nguyen Van Long, Pham Thanh Long, Nguyen Thanh Phuong, Le Tin Vinh Quang, Nguyen Tung, Vu Sinh Nam, Vu Trong Duoc, Nguyen Duc Thinh, Randal Schoepp, Keersten Ricks, Ken Inui, Pawin Padungtod, Christine K. Johnson, Jonna A. K. Mazet, Chris Walzer, Sarah H. Olson and Amanda E. Fineadd Show full author list remove Hide full author list
Viruses 2023, 15(3), 790; https://doi.org/10.3390/v15030790 - 20 Mar 2023
Cited by 4 | Viewed by 7954
Abstract
A One Health cross-sectoral surveillance approach was implemented to screen biological samples from bats, pigs, and humans at high-risk interfaces for zoonotic viral spillover for five viral families with zoonotic potential in Viet Nam. Over 1600 animal and human samples from bat guano [...] Read more.
A One Health cross-sectoral surveillance approach was implemented to screen biological samples from bats, pigs, and humans at high-risk interfaces for zoonotic viral spillover for five viral families with zoonotic potential in Viet Nam. Over 1600 animal and human samples from bat guano harvesting sites, natural bat roosts, and pig farming operations were tested for coronaviruses (CoVs), paramyxoviruses, influenza viruses, filoviruses and flaviviruses using consensus PCR assays. Human samples were also tested using immunoassays to detect antibodies against eight virus groups. Significant viral diversity, including CoVs closely related to ancestors of pig pathogens, was detected in bats roosting at the human–animal interfaces, illustrating the high risk for CoV spillover from bats to pigs in Viet Nam, where pig density is very high. Season and reproductive period were significantly associated with the detection of bat CoVs, with site-specific effects. Phylogeographic analysis indicated localized viral transmission among pig farms. Our limited human sampling did not detect any known zoonotic bat viruses in human communities living close to the bat cave and harvesting bat guano, but our serological assays showed possible previous exposure to Marburg virus-like (Filoviridae), Crimean–Congo hemorrhagic fever virus-like (Bunyaviridae) viruses and flaviviruses. Targeted and coordinated One Health surveillance helped uncover this viral pathogen emergence hotspot. Full article
(This article belongs to the Special Issue Viruses and Bats 2023)
Show Figures

Figure 1

Figure 1
<p>Map of sampling locations in targeted provinces in Viet Nam.</p>
Full article ">Figure 2
<p>Maximum Likelihood phylogenetic tree summarizing phylogenetic relationships among bat and pig coronavirus RdRp sequences (concatenated dataset including two fragments of the RdRp gene) identified in Viet Nam. Well-supported nodes (bootstrap &gt; 75%) are indicated by a black dot. Virus sequences isolated in bat samples from Viet Nam are indicated by a bat symbol colored in green while bat viruses from other countries are black. Sequences isolated in pig samples in Viet Nam are indicated by a pig symbol colored in orange while pig viruses from other countries are grey.</p>
Full article ">Figure 3
<p>Maximum Likelihood phylogenetic tree summarizing phylogenetic relationships among bat and pig paramyxovirus Pol gene sequences identified in Viet Nam. Well-supported nodes (bootstrap &gt; 75%) are indicated by a black dot. Virus sequences isolated in bat samples from Viet Nam are indicated by a bat symbol colored in green, while bat viruses from other countries are black bat. Sequences isolated in pig samples in Viet Nam are indicated by a pig symbol colored in orange while pig viruses from other countries are grey.</p>
Full article ">Figure 4
<p>Median-joining networks of (<b>A</b>) Betacoronavirus 1 (RdRp gene, [<a href="#B36-viruses-15-00790" class="html-bibr">36</a>], 393 bp), (<b>B</b>) porcine parainfluenza virus 1 (Pol gene, 546 bp), and (<b>C</b>) influenza A virus (PB1 gene, 384 bp). Circles correspond to distinct viral sequences and circle sizes are proportional to the number of identical sequences in the dataset. Small black circles represent median vectors (ancestral or unsampled intermediate sequences). The numbers of mutational steps between sequences are represented as hatch marks along branches.</p>
Full article ">Figure 5
<p>Median-joining networks of (<b>A</b>) bat CoV 512/2005 (RdRp gene, [<a href="#B36-viruses-15-00790" class="html-bibr">36</a>], 393 bp), and (<b>B</b>) PREDICT_CoV-35 (RdRp gene, [<a href="#B36-viruses-15-00790" class="html-bibr">36</a>], 393 bp). Circles correspond to distinct viral sequences and circle sizes are proportional to the number of identical sequences in the dataset. Small black circles represent median vectors (ancestral or unsampled intermediate sequences). The numbers of mutational steps between sequences are represented as hatch marks along branches.</p>
Full article ">Figure 6
<p>Spatiotemporal dispersal of bat coronavirus 512/2005 (grey shades) and PREDICT_CoV-35 (brown shades) in Southeast Asia inferred from a fragment of the RdRp gene (386 bp, [<a href="#B36-viruses-15-00790" class="html-bibr">36</a>]). Arrows indicate the direction of dispersal routes. Darker arrow colors indicate older dispersal events for both viruses.</p>
Full article ">
13 pages, 2294 KiB  
Review
Opportunities for CAR-T Cell Immunotherapy in HIV Cure
by Gerard Campos-Gonzalez, Javier Martinez-Picado, Talia Velasco-Hernandez and Maria Salgado
Viruses 2023, 15(3), 789; https://doi.org/10.3390/v15030789 - 19 Mar 2023
Cited by 9 | Viewed by 5924
Abstract
Chimeric antigen receptor (CAR) technology is having a huge impact in the blood malignancy field and is becoming a well-established therapy for many types of leukaemia. In recent decades, efforts have been made to demonstrate that CAR-T cells have potential as a therapy [...] Read more.
Chimeric antigen receptor (CAR) technology is having a huge impact in the blood malignancy field and is becoming a well-established therapy for many types of leukaemia. In recent decades, efforts have been made to demonstrate that CAR-T cells have potential as a therapy to achieve a sterilizing cure for human immunodeficiency virus (HIV) infection. However, translation of this technology to the HIV scenario has not been easy, as many challenges have appeared along the way that hinder the consolidation of CAR-T cells as a putative therapy. Here, we review the origin and development of CAR-T cells, describe the advantages of CAR-T cell therapy in comparison with other therapies, and describe the major obstacles currently faced regarding application of this technology in the HIV field, specifically, viral escape, CAR-T cell infectivity, and accessibility to hidden reservoirs. Nonetheless, promising results in successfully tackling some of these issues that have been obtained in clinical trials suggest a bright future for CAR-T cells as a consolidated therapy. Full article
(This article belongs to the Special Issue CAR-T Cell Therapy for HIV Cure 2023)
Show Figures

Figure 1

Figure 1
<p>Schematic representation of the main limitations of CAR-T cells as a therapy for HIV and new approaches to overcome the limitations. (<b>A</b>) CD4-CAR-T cells can be infected by HIV-1 through interaction with CD4 and CCR5 receptors present in the CAR. Alternative approaches proposed: (i) use of bNAbs to build the CAR; (ii) deletion of the <span class="html-italic">CCR5</span> gene through gene editing techniques to prevent infection; or (iii) both approaches combined. (<b>B</b>) The high mutation rate of HIV-1 can generate HIV-1 strains resistant to the CAR-T cell cytotoxic effect. Some of the new CAR-T cell designs: (i) duoCAR-T cells, which are able to target more than one gp160 epitope, and (ii) convertibleCAR-T built with a common binding site for a variety of gp160-specific bNAbs that can mitigate viral escape. (<b>C</b>) Latently infected cells from reservoirs are not accessible to CAR-T cells, but shock-and-kill approaches such as using LRAs can transcriptionally reactivate these cells making them accessible targets for CAR-T cells.</p>
Full article ">
13 pages, 4381 KiB  
Article
Small RNA Profiling of Cucurbit Yellow Stunting Disorder Virus from Susceptible and Tolerant Squash (Cucurbita pepo) Lines
by Saritha Raman Kavalappara, Sudeep Bag, Alex Luckew and Cecilia E. McGregor
Viruses 2023, 15(3), 788; https://doi.org/10.3390/v15030788 - 19 Mar 2023
Cited by 3 | Viewed by 2706
Abstract
RNA silencing is a crucial mechanism of the antiviral immunity system in plants. Small RNAs guide Argonaut proteins to target viral RNA or DNA, preventing virus accumulation. Small RNA profiles in Cucurbita pepo line PI 420328 with tolerance to cucurbit yellow stunting disorder [...] Read more.
RNA silencing is a crucial mechanism of the antiviral immunity system in plants. Small RNAs guide Argonaut proteins to target viral RNA or DNA, preventing virus accumulation. Small RNA profiles in Cucurbita pepo line PI 420328 with tolerance to cucurbit yellow stunting disorder virus (CYSDV) were compared with those in Gold Star, a susceptible cultivar. The lower CYSDV symptom severity in PI 420328 correlated with lower virus titers and fewer sRNAs derived from CYSDV (vsRNA) compared to Gold Star. Elevated levels of 21- and 22-nucleotide (nt) size class vsRNAs were observed in PI 420328, indicating more robust and efficient RNA silencing in PI 420328. The distribution of vsRNA hotspots along the CYSDV genome was similar in both PI 420328 and Gold Star. However, the 3’ UTRs, CPm, and p26 were targeted at a higher frequency in PI 420328. Full article
(This article belongs to the Special Issue Crop Resistance to Viral Infections)
Show Figures

Figure 1

Figure 1
<p>Impact of CYSDV infection on squash line Gold Star and PI 420328, 30 days after inoculation: (<b>A</b>): mock-inoculated plants with non-viruliferous whiteflies displayed no phenotypic symptoms; (<b>B</b>): Gold Star developed prominent interveinal chlorosis and yellowing; (<b>C</b>) PI 420328 only displayed yellowing symptoms on the lower leaves; (<b>D</b>) estimated copy numbers of CYSDV in Gold Star (blue) and PI 420328 (red) at 30 DAI. Viral titer is represented as the mean Log concentration per ng total RNA. Each value is the average of three biological replicates with standard error bars; * indicates a significant difference (<span class="html-italic">p</span> &lt; 0.05); (<b>E</b>) number of CYSDV-derived vsRNA reads per million total reads averaged over three replicates. Levels not connected by the same letter are significantly different (Student’s <span class="html-italic">t</span>-test, <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>Relative abundance of different size classes in total vsRNAs 30 days after inoculation in susceptible (Gold Star) and tolerant (PI 420328) squash, systemically infected with cucurbit yellow stunting disorder virus (CYSDV). Each value is the average of three replicates with standard error bars. Asterix <sup>(</sup>*<sup>)</sup> indicates that corresponding values in Gold Star and PI 420328 are significantly different (<span class="html-italic">t</span>-test, <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>Distribution of vsRNAs along the genomes of CYSDV in PI 420328 and Gold Star 30 days after inoculation with cucurbit yellow stunting disorder virus (CYSDV). The histograms plot the numbers of 21 to 24 nt viral siRNA reads at each nucleotide position of CYSDV genomic RNA 1 and RNA 2, (mapped with zero mismatches). The bars above the axis (blue) represent sense (forward) reads starting at each position, and those below (red) represent antisense (reverse) reads ending at the respective position.</p>
Full article ">Figure 4
<p>Accumulation of 21–25 nt virus-derived sRNAs corresponding to UTRs and cistrons of CYSDV in <span class="html-italic">Cucurbita pepo</span> leaves at 30 DAI. Values are the average of three biological replicates and are represented as the frequency of vsRNA/100 bp of gene length/million total vsRNA reads. Asterix <sup>(</sup>*<sup>)</sup> indicates that corresponding values in Gold Star and PI 420328 are significantly different (<span class="html-italic">t</span>-test, <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>The 5′-terminal nucleotide frequency of 20–24 nt CYSDV-derived small RNAs (sRNA) in Gold Star and PI 420328. The composite bar graphs indicate the percentage of 5′ U (red), 5′ G (orange), 5′ C (blue), and 5′ A (green) for each of the size classes of virus CYSDV-derived sRNAs.</p>
Full article ">
16 pages, 687 KiB  
Review
COVID-19 Pharmacotherapy in Pregnancy: A Literature Review of Current Therapeutic Choices
by Karolina Akinosoglou, Georgios Schinas, Emmanouil-Angelos Rigopoulos, Eleni Polyzou, Argyrios Tzouvelekis, George Adonakis and Charalambos Gogos
Viruses 2023, 15(3), 787; https://doi.org/10.3390/v15030787 - 19 Mar 2023
Cited by 9 | Viewed by 3087
Abstract
The clinical management of COVID-19 in pregnant women, who are considered a vulnerable population, remains uncertain even as the pandemic subsides. SARS-CoV-2 affects pregnant individuals in multiple ways and has been associated with severe maternal morbidity and mortality, as well as neonatal complications. [...] Read more.
The clinical management of COVID-19 in pregnant women, who are considered a vulnerable population, remains uncertain even as the pandemic subsides. SARS-CoV-2 affects pregnant individuals in multiple ways and has been associated with severe maternal morbidity and mortality, as well as neonatal complications. The unique anatomy and physiology of gestation make managing COVID-19 in this population a complex and challenging task, emphasizing the importance of spreading knowledge and expertise in this area. Therapeutic interventions require distinct clinical consideration, taking into account differences in pharmacokinetics, vertical transmission, drug toxicities, and postnatal care. Currently, there is limited data on antiviral and immunomodulating COVID-19 pharmacotherapy in pregnancy. Some medication has been shown to be safe and well tolerated among pregnant women with COVID-19; however, the lack of randomized clinical trials and studies in this patient population is evident. Available vaccines are considered safe and effective, with no evidence of harm to the fetus, embryo development, or short-term postnatal development. Pregnant women should be counseled about the risks of SARS-CoV-2 infection and informed of available ways to protect themselves and their families. Effective treatments for COVID-19 should not be withheld from pregnant individuals, and more research is needed to ensure the best outcomes. Full article
(This article belongs to the Special Issue COVID-19 Pharmacotherapy)
Show Figures

Figure 1

Figure 1
<p>Study flowchart.</p>
Full article ">Figure 2
<p>Recommended therapeutic regimens in pregnancy according to NIH living guidance. “?” indicates uncertainty or lack of evidence.</p>
Full article ">
13 pages, 6382 KiB  
Article
Above-Standard Survival of Hepatocellular Carcinoma as the Final Outcome of Comprehensive Hepatology Care Programs in a Remote HCV-Endemic Area
by Wei-Ru Cho, Hui-Ling Huang, Nien-Tzu Hsu, Tung-Jung Huang and Te-Sheng Chang
Viruses 2023, 15(3), 786; https://doi.org/10.3390/v15030786 - 19 Mar 2023
Viewed by 1724
Abstract
Early detection and prompt linkage to care are critical for hepatocellular carcinoma (HCC) care. Chang Gung Memorial Hospital (CGMH) Yunlin branch, a local hospital in a rural area, undertakes health checkup programs in addition to its routine clinical service. Patients with HCC are [...] Read more.
Early detection and prompt linkage to care are critical for hepatocellular carcinoma (HCC) care. Chang Gung Memorial Hospital (CGMH) Yunlin branch, a local hospital in a rural area, undertakes health checkup programs in addition to its routine clinical service. Patients with HCC are referred to CGMH Chiayi branch, a tertiary referral hospital, for treatment. This study enrolled 77 consecutive patients with newly diagnosed HCCs between 2017 and 2022, with a mean age of 65.7 ± 11.1 years. The screening group included HCC patients detected through health checkups, and those detected by routine clinical service served as the control group. Compared to the 24 patients in the control group, the 53 patients in the screening group had more cases with early stage cancer (Barcelona Clinic Liver Cancer or BCLC stage 0 + A 86.8% vs. 62.5%, p = 0.028), better liver reserve (albumin–bilirubin or ALBI grade I 77.3% vs. 50%, p = 0.031) and more prolonged survival (p = 0.036). The median survival rates of the 77 patients were >5 years, 3.3 years, and 0.5 years in the BCLC stages 0 + A, B, and C, respectively, which were above the expectations of the BCLC guideline 2022 for stages 0, A, and B. This study provides a model of HCC screening and referral to high-quality care in remote viral-hepatitis-endemic areas. Full article
(This article belongs to the Special Issue Hepatitis-Associated Liver Cancer)
Show Figures

Figure 1

Figure 1
<p>Screening and linkage to accessible care.</p>
Full article ">Figure 2
<p>Patient enrollment flow diagram.</p>
Full article ">Figure 3
<p>Kaplan–Meier analysis of overall survival between the screening group and control group.</p>
Full article ">Figure 4
<p>Kaplan–Meier analysis of overall survival among BCLC 0 + A, BCLC B and BCLC C patients.</p>
Full article ">
23 pages, 1077 KiB  
Review
EV-A71 Mechanism of Entry: Receptors/Co-Receptors, Related Pathways and Inhibitors
by Kanghong Hu, Rominah Onintsoa Diarimalala, Chenguang Yao, Hanluo Li and Yanhong Wei
Viruses 2023, 15(3), 785; https://doi.org/10.3390/v15030785 - 18 Mar 2023
Cited by 6 | Viewed by 3728
Abstract
Enterovirus A71, a non-enveloped single-stranded (+) RNA virus, enters host cells through three stages: attachment, endocytosis and uncoating. In recent years, receptors/co-receptors anchored on the host cell membrane and involved in this process have been continuously identified. Among these, hSCARB-2 was the first [...] Read more.
Enterovirus A71, a non-enveloped single-stranded (+) RNA virus, enters host cells through three stages: attachment, endocytosis and uncoating. In recent years, receptors/co-receptors anchored on the host cell membrane and involved in this process have been continuously identified. Among these, hSCARB-2 was the first receptor revealed to specifically bind to a definite site of the EV-A71 viral capsid and plays an indispensable role during viral entry. It actually acts as the main receptor due to its ability to recognize all EV-A71 strains. In addition, PSGL-1 is the second EV-A71 receptor discovered. Unlike hSCARB-2, PSGL-1 binding is strain-specific; only 20% of EV-A71 strains isolated to date are able to recognize and bind it. Some other receptors, such as sialylated glycan, Anx 2, HS, HSP90, vimentin, nucleolin and fibronectin, were discovered successively and considered as “co-receptors” because, without hSCARB-2 or PSGL-1, they are not able to mediate entry. For cypA, prohibitin and hWARS, whether they belong to the category of receptors or of co-receptors still needs further investigation. In fact, they have shown to exhibit an hSCARB-2-independent entry. All this information has gradually enriched our knowledge of EV-A71’s early stages of infection. In addition to the availability of receptors/co-receptors for EV-A71 on host cells, the complex interaction between the virus and host proteins and various intracellular signaling pathways that are intricately connected to each other is critical for a successful EV-A71 invasion and for escaping the attack of the immune system. However, a lot remains unknown about the EV-A71 entry process. Nevertheless, researchers have been continuously interested in developing EV-A71 entry inhibitors, as this study area offers a large number of targets. To date, important progress has been made toward the development of several inhibitors targeting: receptors/co-receptors, including their soluble forms and chemically designed compounds; virus capsids, such as capsid inhibitors designed on the VP1 capsid; compounds potentially interfering with related signaling pathways, such as MAPK-, IFN- and ATR-inhibitors; and other strategies, such as siRNA and monoclonal antibodies targeting entry. The present review summarizes these latest studies, which are undoubtedly of great significance in developing a novel therapeutic approach against EV-A71. Full article
(This article belongs to the Special Issue Enteroviruses 2023)
Show Figures

Figure 1

Figure 1
<p>Enterovirus A71: mode of entry, virus–host interaction and viral life cycle. (<b>A</b>) Schematic illustration of the 3 different modes of entry identified to date for EV-A71 to invade cells. EV-A71 enters cells through receptor-mediated endocytosis, which includes clathrin-mediated (I), caveolae-mediated (II) and endophilin-A2-mediated (III). (I) In presence of the main receptor, hSCARB2, EV-A71 binds to it with or without the help of one or more co-receptors. The binding induces recruitment of adaptor proteins on the receptor cytoplasmic tail, which afterward bind to clathrin and form “a clathrin-coated pit” (CCP), leading to EV-A71 entry through clathrin-mediated endocytosis. hSCARB2 delivers β-GC from the ER to the lysosomes under physiological conditions. hSCARB2 is abundant in the lysosomal and endosomal compartments, and it also shuttles to the plasma membrane where it encounters EV-A71. After the binding of the virus on the cell surface, the virus–receptor complex is internalized. In the endosome or lysosome, where the pH is low, the virus initiates a conformational change that leads to uncoating. (II) In an alternative process, caveolins are the proteins binding to PSGL-1 cytoplasmic tail in the presence of actin cytoskeleton. Thus, this entrance way is called caveolae-mediated endocytosis. PSGL-1 can bind to EV-A71 and internalize via caveolin-mediated endocytosis, but PSGL-1 cannot initiate uncoating. (III) Recently, endophilin-A2-mediated endocytosis was identified. However, how receptor/co-receptor mediates this kind of entry is not yet elucidated. Once endocytosis initiates, the virus is internalized and delivered to early endosome for translocation, which is assured by the endosomal sorting complex required for transport to multivesicular bodies (ESCRT-MVBs). (<b>B</b>) EV-A71 promotes its production through interaction with intracellular signaling pathways. Once the virus is captured by main or co-receptors, several intracellular signaling pathways related to immune response are activated in order to eliminate the viral infection. However, EV-A71 has to escape immune response and overcome cellular apoptosis/autophagy for its survival by interfering with these pathways by interacting with host proteins. MAPK signaling cascade is activated after release of IL-2, IL-4, IL-10 and TNF-α. Subsequently, together with activated P13K/Akt pathway, MAPK downregulates GSK3, resulting in the delay of apoptosis. Multiple proteins involved in IFN-, apoptosis- and autophagy-related pathways are also regulated, such as JAK/STAT, TRAFs, p53, bax and mTOR. Consequently, many aspects of viral production are promoted, including viral polyprotein processing, RNA synthesis, assembly and maturation and release of new virions.</p>
Full article ">
13 pages, 2077 KiB  
Article
Open Reading Frame 4 Is Not Essential in the Replication and Infection of Genotype 1 Hepatitis E Virus
by Huimin Bai, Yasushi Ami, Yuriko Suzaki, Yen Hai Doan, Masamichi Muramatsu and Tian-Cheng Li
Viruses 2023, 15(3), 784; https://doi.org/10.3390/v15030784 - 18 Mar 2023
Cited by 3 | Viewed by 2165
Abstract
Genotype 1 hepatitis E virus (HEV-1), unlike other genotypes of HEV, has a unique small open reading frame known as ORF4 whose function is not yet known. ORF4 is located in an out-framed manner in the middle of ORF1, which encodes putative 90 [...] Read more.
Genotype 1 hepatitis E virus (HEV-1), unlike other genotypes of HEV, has a unique small open reading frame known as ORF4 whose function is not yet known. ORF4 is located in an out-framed manner in the middle of ORF1, which encodes putative 90 to 158 amino acids depending on the strains. To explore the role of ORF4 in HEV-1 replication and infection, we cloned the complete genome of wild-type HEV-1 downstream of a T7 RNA polymerase promoter, and the following ORF4 mutant constructs were prepared: the first construct had TTG instead of the initiation codon ATG (A2836T), introducing an M→L mutation in ORF4 and a D→V mutation in ORF1. The second construct had ACG instead of the ATG codon (T2837C), introducing an M→T mutation in ORF4. The third construct had ACG instead of the second in-frame ATG codon (T2885C), introducing an M→T mutation in ORF4. The fourth construct contained two mutations (T2837C and T2885C) accompanying two M→T mutations in ORF4. For the latter three constructs, the accompanied mutations introduced in ORF1 were all synonymous changes. The capped entire genomic RNAs were generated by in vitro transcription and used to transfect PLC/PRF/5 cells. Three mRNAs containing synonymous mutations in ORF1, i.e., T2837CRNA, T2885CRNA, and T2837C/T2885CRNA, replicated normally in PLC/PRF/5 cells and generated infectious viruses that successfully infected Mongolian gerbils as the wild-type HEV-1 did. In contrast, the mutant RNA, i.e., A2836TRNA, accompanying an amino acid change (D937V) in ORF1 generated infectious viruses upon transfection, but they replicated slower than the wild-type HEV-1 and failed to infect Mongolian gerbils. No putative viral protein(s) derived from ORF4 were detected in the wild-type HEV-1- as well as the mutant virus-infected PLC/PRF/5 cells by Western blot analysis using a high-titer anti-HEV-1 IgG antibody. These results demonstrated that the ORF4-defective HEV-1s had the ability to replicate in the cultured cells, and that these defective viruses had the ability to infect Mongolian gerbils unless the overlapping ORF1 was accompanied by non-synonymous mutation(s), confirming that ORF4 is not essential in the replication and infection of HEV-1. Full article
(This article belongs to the Section Human Virology and Viral Diseases)
Show Figures

Figure 1

Figure 1
<p>HEV-1 (LC061267) genome organization and the position of the nucleotide mutations. ORF1 to ORF4 and the putative functional domains observed in ORF1 are depicted. Hel: helicase, HVR: hypervariable region, MT: methyltransferase, PCP: papain-like cysteine protease, RdRp: RNA-dependent RNA polymerase, X: X-domain, and Y: Y-domain. The numbers indicate the nucleotide position from the 5′-end. Mutated nucleotides are shown by <span class="html-italic">bold</span> and <span class="html-italic">italics</span>.</p>
Full article ">Figure 2
<p>Generation and replication of the original and ORF4-defective HEV-1 in PLC/PRF/5 cells. PLC/PRF/5 cells were transfected with capped HEV-1<sup>RNA</sup>, A2836T<sup>RNA</sup>, T2837C<sup>RNA</sup>, T2885C<sup>RNA</sup>, and T2837<sup>RNA</sup>/T2885<sup>RNA</sup>, respectively. The culture supernatant was collected every 4 days, and the viral RNA was measured by RT-qPCR. The viral RNA copy numbers are shown in the HEV-1<sup>RNA</sup>-transfected cells (≤), A2836T<sup>RNA</sup>-transfected cells (○), T2837C<sup>RNA</sup>-transfected cells (◊), T2885C<sup>RNA</sup>-transfected cells (△), and T2837C/T2885C<sup>RNA</sup>-transfected cells (×).</p>
Full article ">Figure 3
<p>Infectivity of the original and ORF4-defective HEV-1s in PLC/PRF/5 cells. PLC/PRF/5 cells were inoculated with HEV-1p0, A2836Tp0, T2837Cp0, T2885Cp0, and T2837C/T2885Cp0 containing the same RNA copy number, respectively. The culture supernatant was collected every 4 days, and the viral RNA was measured by RT-qPCR. The RNA titers are shown in the HEV-1p0-infected cells (≤), A2836Tp0-infected cells (○), T2837Cp0-infected cells (◊), T2885Cp0-infected cells (△), and T2837C/T2885Cp0-infected cells (×).</p>
Full article ">Figure 4
<p>Infection of original and ORF4-defective viruses in Mongolian gerbils. Fifteen Mongolia gerbils were randomly separated into five groups (n = 3 per group). Individual gerbils are indicated by ○, △, and ☐. Each group received HEV-1p0, A2836Tp0, T2837Cp0, T2885Cp0, or T2837C/T2885Cp0 via intraperitoneal injection. The kinetics of the viral RNA in the fecal specimens were measured by RT-qPCR (<b>a</b>). The serum samples were collected at the end of the experiment (day 28 post-inoculation [p.i.]), and the anti-HEV-IgG antibody titers were determined by an ELISA with the virus-like particles (VLPs) of HEV-1 as the antigens (<b>b</b>). <span class="html-italic">Dotted lines</span>: the cut-off values. The minimum endpoints of the antibody titers are <span class="html-italic">blackened</span>.</p>
Full article ">Figure 5
<p>Detection of viral proteins in HEV-1-infected PLC/PRF/5 cells. A cynomolgus monkey’s serum bearing the anti-HEV-1 IgG antibody titer 1:3,276,800 by ELISA was used for Western blot analyses. The minimum endpoints of the antibody titers are <span class="html-italic">blackened</span> (<b>a</b>). HEV-1p0-, T2837Cp0-, and T2837C/T2885Cp0-infected PLC/PRF/5 cells were harvested on day 48 p.i., and the virus proteins were detected by Western blot analysis (<b>b</b>). ORF4 was expressed by an <span class="html-italic">E. coli</span> expression system, and the related protein was analyzed by SDS-PAGE (<b>c</b>) and a Western blot analysis with monkey anti-HEV-1 serum (<b>d</b>) and anti-HEV-7 serum (<b>e</b>). M: molecular weight, NC: no-infected PLC/PRF/5 cells (<b>b</b>) or no-transformed BL21 (DE3) cells (<b>c</b>–<b>e</b>). pET32a (+): vector pET32a (+)-transformed BL21 (DE3). pET32aORF4: pET32aORF4-transformed BL21 (DE3).</p>
Full article ">
12 pages, 282 KiB  
Perspective
Neurological Dysfunction in Long COVID Should Not Be Labelled as Functional Neurological Disorder
by Christina M. Van der Feltz-Cornelis, Andrew S. Moriarty and William David Strain
Viruses 2023, 15(3), 783; https://doi.org/10.3390/v15030783 - 18 Mar 2023
Cited by 2 | Viewed by 9892
Abstract
There have been suggestions that Long COVID might be purely functional (meaning psychological) in origin. Labelling patients with neurological dysfunction in Long COVID as having functional neurological disorder (FND) in the absence of proper testing may be symptomatic of that line of thought. [...] Read more.
There have been suggestions that Long COVID might be purely functional (meaning psychological) in origin. Labelling patients with neurological dysfunction in Long COVID as having functional neurological disorder (FND) in the absence of proper testing may be symptomatic of that line of thought. This practice is problematic for Long COVID patients, as motor and balance symptoms have been reported to occur in Long COVID frequently. FND is characterized by the presentation of symptoms that seem neurological but lack compatibility of the symptom with a neurological substrate. Although diagnostic classification according to the ICD-11 and DSM-5-TR is dependent predominantly on the exclusion of any other medical condition that could account for the symptoms, current neurological practice of FND classification allows for such comorbidity. As a consequence, Long COVID patients with motor and balance symptoms mislabeled as FND have no longer access to Long COVID care, whereas treatment for FND is seldom provided and is ineffective. Research into underlying mechanisms and diagnostic methods should explore how to determine whether motor and balance symptoms currently diagnosed as FND should be considered one part of Long COVID symptoms, in other words, one component of symptomatology, and in which cases they correctly represent FND. Research into rehabilitation models, treatment and integrated care are needed, which should take into account biological underpinnings as well as possible psychological mechanisms and the patient perspective. Full article
(This article belongs to the Section SARS-CoV-2 and COVID-19)
Previous Issue
Next Issue
Back to TopTop