[go: up one dir, main page]

 
 

Applications of Baculoviruses: Expression Factories, Vaccines and VLPs, Gene Delivery Vectors, Biological Control and Virus Genetics Models

A special issue of Viruses (ISSN 1999-4915). This special issue belongs to the section "Invertebrate Viruses".

Deadline for manuscript submissions: closed (15 June 2023) | Viewed by 22952

Special Issue Editors


E-Mail Website
Guest Editor
Hydrosciences Montpellier, Université de Montpellier, IMT Mines Ales, CNRS, IRD, Ales, France
Interests: baculovirus genetic diversity; virus-host interactions; biological control with viruses
Special Issues, Collections and Topics in MDPI journals

E-Mail Website
Guest Editor
Department of Biological and Medical Sciences, Oxford Brookes University, Oxford, UK
Interests: insect virus biology; baculovirus expression systems; baculovirus-host cell interactions at cellular level; molecular virology; insect virology; the biology and replication of insect baculoviruses in cultured insect cells and in larvae; the role of non-essential genes encoding proteins; the trafficking of virus proteins and particles through insect cells; baculoviruses as gene expression vectors
Special Issues, Collections and Topics in MDPI journals

Special Issue Information

Dear Colleagues,

Studies on baculoviruses represent a large share of the insect virus research portfolio, partly due to the widespread use of this virus family in biotechnological applications since the landmark studies of Miller and Summers in the 1980s. In recent years, growing knowledge of the replication cycle of these viruses has allowed expanding their use and improving the expression of heterologous genes, both in terms of quantity and quality. The use of baculoviruses for the control of insect pests has been specifically addressed in two recent special issues, but new approaches in this field continue to emerge. Baculoviruses are also used as models for other pathogens, as they possess a narrow host range. In this special issue, we would like to produce a collection of papers showing the variety of applications of this virus family across agriculture, biotechnology, and medicine. We also hope this collection will promote further developments due to the cross-fertilisation of the ideas presented across the various contributions.

Prof. Dr. Miguel López-Ferber
Prof. Dr. Linda King
Guest Editors

Manuscript Submission Information

Manuscripts should be submitted online at www.mdpi.com by registering and logging in to this website. Once you are registered, click here to go to the submission form. Manuscripts can be submitted until the deadline. All submissions that pass pre-check are peer-reviewed. Accepted papers will be published continuously in the journal (as soon as accepted) and will be listed together on the special issue website. Research articles, review articles as well as short communications are invited. For planned papers, a title and short abstract (about 100 words) can be sent to the Editorial Office for announcement on this website.

Submitted manuscripts should not have been published previously, nor be under consideration for publication elsewhere (except conference proceedings papers). All manuscripts are thoroughly refereed through a single-blind peer-review process. A guide for authors and other relevant information for submission of manuscripts is available on the Instructions for Authors page. Viruses is an international peer-reviewed open access monthly journal published by MDPI.

Please visit the Instructions for Authors page before submitting a manuscript. The Article Processing Charge (APC) for publication in this open access journal is 2600 CHF (Swiss Francs). Submitted papers should be well formatted and use good English. Authors may use MDPI's English editing service prior to publication or during author revisions.

Benefits of Publishing in a Special Issue

  • Ease of navigation: Grouping papers by topic helps scholars navigate broad scope journals more efficiently.
  • Greater discoverability: Special Issues support the reach and impact of scientific research. Articles in Special Issues are more discoverable and cited more frequently.
  • Expansion of research network: Special Issues facilitate connections among authors, fostering scientific collaborations.
  • External promotion: Articles in Special Issues are often promoted through the journal's social media, increasing their visibility.
  • e-Book format: Special Issues with more than 10 articles can be published as dedicated e-books, ensuring wide and rapid dissemination.

Further information on MDPI's Special Issue polices can be found here.

Related Special Issue

Published Papers (11 papers)

Order results
Result details
Select all
Export citation of selected articles as:

Research

Jump to: Review

16 pages, 6978 KiB  
Article
Comprehensive Comparison of Baculoviral and Plasmid Gene Delivery in Mammalian Cells
by Maria Toth, Manuel Reithofer, Gregory Dutra, Patricia Pereira Aguilar, Astrid Dürauer and Reingard Grabherr
Viruses 2024, 16(3), 426; https://doi.org/10.3390/v16030426 - 10 Mar 2024
Viewed by 1359
Abstract
(1) Recombinant protein production in mammalian cells is either based on transient transfection processes, often inefficient and underlying high batch-to-batch variability, or on laborious generation of stable cell lines. Alternatively, BacMam, a transduction process using the baculovirus, can be employed. (2) Six transfecting [...] Read more.
(1) Recombinant protein production in mammalian cells is either based on transient transfection processes, often inefficient and underlying high batch-to-batch variability, or on laborious generation of stable cell lines. Alternatively, BacMam, a transduction process using the baculovirus, can be employed. (2) Six transfecting agents were compared to baculovirus transduction in terms of transient and stable protein expression characteristics of the model protein ACE2-eGFP using HEK293-6E, CHO-K1, and Vero cell lines. Furthermore, process optimization such as expression enhancement using sodium butyrate and TSA or baculovirus purification was assessed. (3) Baculovirus transduction efficiency was superior to all transfection agents for all cell lines. Transduced protein expression was moderate, but an 18-fold expression increase was achieved using the enhancer sodium butyrate. Ultracentrifugation of baculovirus from a 3.5 L bioreactor significantly improved the transduction efficiency and protein expression. Stable cell lines were obtained with each baculovirus transduction, yet stable cell line generation after transfection was highly unreliable. (4) This study demonstrated the superiority of the BacMam platform to standard transfections. The baculovirus efficiently transduced an array of cell lines both transiently and stably and achieved the highest efficiency for all tested cell lines. The feasibility of the scale-up of baculovirus production was demonstrated and the possibility of baculovirus purification was successfully explored. Full article
Show Figures

Figure 1

Figure 1
<p>Workflow of the comparison of six transfecting agents to baculovirus transduction in three different cell lines.</p>
Full article ">Figure 2
<p>Determination of optimal transduction settings by baculovirus multiplicity of infection (MOI) titration. (<b>a</b>) Transduction efficiency expressed as a percentage of fluorescent HEK293-6E cells and (<b>b</b>) protein expression level measured as cellular fluorescence. Each bar represents four to five independent experiments. (<b>c</b>) Transduction efficiency of CHO-K1 cells and (<b>d</b>) cellular fluorescence resulting from a single experiment.</p>
Full article ">Figure 3
<p>Comparison of GFP expression between transfecting agents PEI, Lipofectamine, Fugene, calcium phosphate, Attractene, and Polyfect and baculovirus transduction at different MOIs with the three cell lines HEK293-6E (<b>a</b>,<b>b</b>), CHO-K1 (<b>c</b>,<b>d</b>), and Vero (<b>e</b>,<b>f</b>). (<b>a</b>,<b>c</b>,<b>e</b>) Transfection efficiency is assessed as the number of cells that exhibit fluorescence (%GFP pos.) and (<b>b</b>,<b>d</b>,<b>f</b>) GFP expression is measured as the geometric mean of fluorescence intensity (gMFI), calculated as a fold change in fluorescence over the negative control. Each bar represents five to six independent experiments and was evaluated using one-way ANOVA with Tukey’s post-test.</p>
Full article ">Figure 4
<p>Comparison of viability data between the indicated transfecting agents and baculovirus transduction at different MOIs with the three cell lines HEK293-6E (<b>a</b>,<b>b</b>), CHO-K1 (<b>c</b>,<b>d</b>), and Vero (<b>e</b>,<b>f</b>). (<b>a</b>,<b>c</b>,<b>e</b>) Viability expressed as a percentage of viable cells and (<b>b</b>,<b>d</b>,<b>f</b>) viable cell density (VCD) measured with the ViCell XR Cell Analyzer. Each bar represents three to five independent experiments and was evaluated using one-way ANOVA with Tukey’s post-test.</p>
Full article ">Figure 5
<p>Baculovirus transduction of HEK293-6E at MOI 10 with an increasing concentration of sodium butyrate and trichostatin A (TSA). (<b>a</b>) The impact of increasing sodium butyrate concentrations on transduction efficiency and (<b>b</b>) GFP expression is assessed. (<b>c</b>) Transduction efficiency and GFP expression (<b>d</b>) with varying TSA concentrations are shown. GFP expression is measured as geometric mean fluorescent intensity (gMFI) in relation to the autofluorescence of the cell. Each bar represents three independent experiments and was evaluated using one-way ANOVA with Dunnett’s post-test. The dotted line in (<b>b</b>,<b>d</b>) represents expression levels achieved with Attractene transfection.</p>
Full article ">Figure 6
<p>Comparison of transduction of HEK293-6E at varying MOIs with unpurified and purified baculovirus based on (<b>a</b>) transduction efficiency expressed as % GFP positive cells and (<b>b</b>) GFP expression assessed as geometric mean fluorescent intensity (gMFI) normalized to the autofluorescence of the cell. Each bar represents three independent experiments and was evaluated using Student’s <span class="html-italic">t</span>-test.</p>
Full article ">Figure 7
<p>Depiction of a fluorescent colony that demonstrated stable cell line generation. The upper side shows fluorescence microscopy while the lower images depict the confocal microscopy image of the same colony. (<b>a</b>,<b>b</b>) Image of a stably transfected CHO-K1 colony using the transfecting agent Lipofectamine while (<b>c</b>,<b>d</b>) show a CHO-K1 colony transduced with Attractene.</p>
Full article ">Figure A1
<p>Stringent sorting strategy of HEK (<b>a</b>) and CHO (<b>b</b>) cells after 2 weeks of selection.</p>
Full article ">Figure A2
<p>Stringent immunofluorescent staining of transduced CHO-K1 cells. Cells were transduced with baculovirus and expression was allowed for 48 h. Thereafter, cells were fixated with 4% PFA, and ACE2 was detected with His-tagged RBD (Klausberger et al. 2021 and 2022). The Histag was detected with mouse Anti Histidine Tag antibody (BioRad). As a secondary antibody to detect the mouse Anti Histidine Tag antibody, a goat anti-mouse* Phycoerythrin antibody (Sigma-Aldrich) was used.</p>
Full article ">
13 pages, 2881 KiB  
Article
Insecticidal Traits of Variants in a Genotypically Diverse Natural Isolate of Anticarsia Gemmatalis Multiple Nucleopolyhedrovirus (AgMNPV)
by Ana Parras-Jurado, Delia Muñoz, Inés Beperet, Trevor Williams and Primitivo Caballero
Viruses 2023, 15(7), 1526; https://doi.org/10.3390/v15071526 - 10 Jul 2023
Cited by 1 | Viewed by 1329
Abstract
Outbreaks of Anticarsia gemmatalis (Hübner, 1818) (Lepidoptera: Erebidae), a major pest of soybean, can be controlled below economic thresholds with methods that do not involve the application of synthetic insecticides. Formulations based on natural isolates of the Anticarsia gemmatalis multiple nucleopolyhedrovirus (AgMNPV) ( [...] Read more.
Outbreaks of Anticarsia gemmatalis (Hübner, 1818) (Lepidoptera: Erebidae), a major pest of soybean, can be controlled below economic thresholds with methods that do not involve the application of synthetic insecticides. Formulations based on natural isolates of the Anticarsia gemmatalis multiple nucleopolyhedrovirus (AgMNPV) (Baculoviridae: Alphabaculovirus) played a significant role in integrated pest management programs in the early 2000s, but a new generation of chemical insecticides and transgenic soybean have displaced AgMNPV-based products over the past decade. However, the marked genotypic variability present among and within alphabaculovirus isolates suggests that highly insecticidal genotypic variants can be isolated and used to reduce virus production costs or overcome isolate-dependent host resistance. This study aimed to select novel variants of AgMNPV with suitable insecticidal traits that could complement the existing AgMNPV active ingredients. Three distinct AgMNPV isolates were compared using their restriction endonuclease profile and in terms of their occlusion body (OB) pathogenicity. One isolate was selected (AgABB51) from which eighteen genotypic variants were plaque purified and characterized in terms of their insecticidal properties. The five most pathogenic variants varied in OB pathogenicity, although none of them was faster-killing or had higher OB production characteristics than the wild-type isolate. We conclude that the AgABB51 wild-type isolates appear to be genotypically structured for fast speed of kill and high OB production, both of which would favor horizontal transmission. Interactions among the component variants are likely to influence this insecticidal phenotype. Full article
Show Figures

Figure 1

Figure 1
<p>Restriction endonuclease profiles of the genomic DNA of three different AgMNPV isolates following treatment with HindIII. m denotes the molecular marker. Fragment size in kilobases (Kb) is shown on the left. Red arrowheads indicate restriction fragment length polymorphisms (RFLPs) and asterisks on the left of each lane indicate the presence of submolar bands.</p>
Full article ">Figure 2
<p>HindIII restriction endonuclease profiles of the genomic DNA of each of the 18 different AgABB51 genotypic variants (A to R). m denotes the molecular marker. Fragment size in kilobases (Kb) is shown on the left. The profile of genotype A is used as a reference to identify the presence (red arrowheads) or absence (green arrowheads) of characteristic restriction fragments in the other genotypic variants. Genotype A was selected for this purpose because of its similarity to the wild-type isolate (AgABB51) (shown in <a href="#viruses-15-01526-f001" class="html-fig">Figure 1</a>).</p>
Full article ">Figure 3
<p>Frequency of AgABB51 genotypic variants A through R. The <span class="html-italic">n</span>-value above each column indicates the number of clones exhibiting each variant’s restriction profile out of a total of 128 clones.</p>
Full article ">Figure 4
<p>Mean percentage of mortality caused by AgABB51 wild-type isolate and each of the genotypic variants A–R on <span class="html-italic">A. gemmatalis</span> second instars inoculated with 1.1 × 10<sup>4</sup> OBs/mL. Error bars indicate the standard error. Different lowercase letters indicate significant differences between variants (ANOVA, Tukey HSD; <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>(<b>a</b>) Median time to death (MTD) values for the AgABB51 isolate and selected genotypic variants in <span class="html-italic">A. gemmatalis</span> second instars. Error bars indicate standard error and different lowercase letters indicate significant differences between variants (Bonferroni-adjusted <span class="html-italic">t</span>-test; <span class="html-italic">p</span> &lt; 0.05). (<b>b</b>) OB production values obtained for AgABB51 and the selected genotypic variants in <span class="html-italic">A. gemmatalis</span> fifth instars. Error bars indicate standard error and different lowercase letters indicate significant differences between variants (Tukey HSD test; <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
15 pages, 1798 KiB  
Article
Evaluating Novel Quantification Methods for Infectious Baculoviruses
by Keven Lothert, Elena Bagrin and Michael W. Wolff
Viruses 2023, 15(4), 998; https://doi.org/10.3390/v15040998 - 19 Apr 2023
Viewed by 2675
Abstract
Accurate and rapid quantification of (infectious) virus titers is of paramount importance in the manufacture of viral vectors and vaccines. Reliable quantification data allow efficient process development at a laboratory scale and thorough process monitoring in later production. However, current gold standard applications, [...] Read more.
Accurate and rapid quantification of (infectious) virus titers is of paramount importance in the manufacture of viral vectors and vaccines. Reliable quantification data allow efficient process development at a laboratory scale and thorough process monitoring in later production. However, current gold standard applications, such as endpoint dilution assays, are cumbersome and do not provide true process analytical monitoring. Accordingly, flow cytometry and quantitative polymerase chain reaction have attracted increasing interest in recent years, offering various advantages for rapid quantification. Here, we compared different approaches for the assessment of infectious viruses, using a model baculovirus. Firstly, infectivity was estimated by the quantification of viral nucleic acids in infected cells, and secondly, different flow cytometric approaches were investigated regarding analysis times and calibration ranges. The flow cytometry technique included a quantification based on post-infection fluorophore expression and labeling of a viral surface protein using fluorescent antibodies. Additionally, the possibility of viral (m)RNA labeling in infected cells was investigated as a proof of concept. The results confirmed that infectivity assessment based on qPCR is not trivial and requires sophisticated method optimization, whereas staining of viral surface proteins is a fast and feasible approach for enveloped viruses. Finally, labeling of viral (m)RNA in infected cells appears to be a promising opportunity but will require further research. Full article
Show Figures

Figure 1

Figure 1
<p>Amplification plots for the qPCR evaluation of different baculovirus concentrations in the range of 0 (i.e., blank)−1.1 × 10<sup>7</sup> IU/mL (<b>A</b>) as well as the linear regression over the concentration range (<b>B</b>). The number of replicate measurements was <span class="html-italic">n</span> = 39 for 1.1 × 10<sup>2</sup> IU/mL and 1.1 × 10<sup>7</sup> IU/mL and <span class="html-italic">n</span> = 18 for all other concentrations and the blank.</p>
Full article ">Figure 2
<p>Amplification plots of the baculovirus DNA by qPCR measurements. The DNA was extracted from samples taken either from the supernatant or from the cell pellet of infected cells after 5 min (i.e., 0 h) of incubation and compared to a negative control (<b>A</b>). The amount of viral DNA detectable in the cell pellet was monitored over the course of 7 h (<b>B</b>) and compared to values after 18 h of infection (<b>C</b>). All samples were prepared in triplicates, with error bars only being displayed in (<b>A</b>) to allow a clearer view of the data.</p>
Full article ">Figure 3
<p>Development of the flow cytometric quantification protocol for baculoviruses (either wild-type or virus expressing the green fluorescent protein, BV-wt and BV-GFP, respectively). In a first approach, (<b>A</b>) the detection capability for both viruses was determined without further method optimization after 18 h of cell infection and subsequent flow cytometric detection of fluorescent cells. Depending on the staining procedure, green (no staining, left column), yellow (PE staining, middle column), or red fluorescence (APC staining, right column) was evaluated. The infection kinetics of the three approaches (<b>B</b>) indicate the earliest possible time of quantification at which the linear calibration range was subsequently determined for each individually optimized strategy (<b>C</b>). Error bars depict standard deviation of technical triplicates (<b>A</b>,<b>B</b>) and 18 replicates with <span class="html-italic">n</span> = 6 on three different days for (<b>C</b>).</p>
Full article ">Figure 4
<p>Quantification of the percentage of fluorescing cells after BV infection and fluorescent labeling of the mRNA of the viral gp64 protein within the first four hours of infection. A multiplicity of infection (MOI) of 1 was used and compared to a negative control without virus infection (error bars indicate the standard deviation of technical triplicates).</p>
Full article ">
20 pages, 3321 KiB  
Article
Transcriptional Reprogramming of Autographa Californica Multiple Nucleopolyhedrovirus Chitinase and Cathepsin Genes Enhances Virulence
by Jeffrey J. Hodgson, A. Lorena Passarelli and Peter J. Krell
Viruses 2023, 15(2), 503; https://doi.org/10.3390/v15020503 - 11 Feb 2023
Cited by 2 | Viewed by 1564
Abstract
The baculoviral chitinase (CHIA) and cathepsin (V-CATH) enzymes promote terminal insect host liquefaction, which aids viral progeny dissemination. Recombinant Autographa californica nucleopolyhedrovirus (AcMNPV)-derived viruses were previously generated with reprogrammed chiA transcription by replacing the native promoter with the AcMNPV polyhedrin (polh) [...] Read more.
The baculoviral chitinase (CHIA) and cathepsin (V-CATH) enzymes promote terminal insect host liquefaction, which aids viral progeny dissemination. Recombinant Autographa californica nucleopolyhedrovirus (AcMNPV)-derived viruses were previously generated with reprogrammed chiA transcription by replacing the native promoter with the AcMNPV polyhedrin (polh) or core protein (p6.9) promoter sequences, but of both these chiA-reprogrammed viruses lacked v-cath transcription and V-CATH enzymatic activity. Here, we report that dual p6.9/polh promoter reprogramming of the adjacent chiA/v-cath genes resulted in modulated temporal transcription of both genes without impacting infectious budded virus production. These promoter changes increased CHIA and V-CATH enzyme activities in infected Spodoptera frugiperda-derived cultured cells and Trichoplusia ni larvae. In addition, larvae infected with the dual reprogrammed virus had earlier mortalities and liquefaction. This recombinant baculovirus, lacking exogenous genomic elements and increased chiA/v-cath expression levels, may be desirable for and amenable to producing enhanced baculovirus-based biopesticides. Full article
Show Figures

Figure 1

Figure 1
<p>Reprogramming <span class="html-italic">chiA</span> and <span class="html-italic">v-cath</span> transcription. (<b>a</b>) Schematic depicting the method used for switching the AcMNPV <span class="html-italic">chiA/v-cath</span> intergenic promoters via an EGFP-selectable virus (AcEGFP) as described before [<a href="#B19-viruses-15-00503" class="html-bibr">19</a>]. The small arrows indicate which promoter (wt, <span class="html-italic">p6.9</span>, or <span class="html-italic">polh</span>) is driving <span class="html-italic">chiA</span> or <span class="html-italic">v-cath</span> transcription. To the right is a summary of the <span class="html-italic">chiA</span> and <span class="html-italic">v-cath</span> expression from each virus (+ = expressed, − = not expressed). (<b>b</b>) Intergenic <span class="html-italic">chiA/v-cath</span> promoter sequence of AcMNPV, AcMNPV-Rep and the dual reprogrammed virus (Acp6.9-chiA/polh-cath). The <span class="html-italic">chiA</span> and <span class="html-italic">v-cath</span> translation start codons are bolded and restriction enzyme recognition sequences (cloning sites) are italicized. For the dual reprogrammed virus, the sequence of the <span class="html-italic">p6.9</span> promoter (driving <span class="html-italic">chiA</span>) is shown from 3′-5′ and that of the adjacent <span class="html-italic">polh</span> promoter (driving <span class="html-italic">v-cath,</span> shaded in grey) are shown from 5′-3′ to reflect the coding strand for each antiparallel ORF. Potential mRNA transcription sites (TAAG) for <span class="html-italic">chiA</span> (dark grey) or <span class="html-italic">v-cath</span> (white) are boxed. The indicated <span class="html-italic">chiA</span> transcription site was mapped for AcMNPV and Acp6.9-chiA [<a href="#B19-viruses-15-00503" class="html-bibr">19</a>], but not Acp6.9-chiA/polh-cath. The <span class="html-italic">v-cath</span> transcription site was mapped previously [<a href="#B20-viruses-15-00503" class="html-bibr">20</a>].</p>
Full article ">Figure 2
<p>Temporal CHIA and proV-CATH expression by AcMNPV-Rep, Acp6.9-chiA, and Acp6.9-chiA/polh-cath in infected SF-21 cells. (<b>a</b>) Northern blots of <span class="html-italic">chiA</span> and <span class="html-italic">v-cath</span> mRNAs over a time-course of infection of SF-21 cells with AcMNPV-Rep, Acp6.9-chiA, and Acp6.9-chiA/polh-cath. Sizes are based on a high-range RNA ladder (Fermentas). Transcripts were detected using DIG-labeled ssDNA probes as described in [<a href="#B19-viruses-15-00503" class="html-bibr">19</a>]. Ethidium bromide-stained rRNA bands indicate RNA equivalency (10 µg/lane). Lane M is an uninfected control sample. (<b>b</b>) Temporal CHIA and proV-CATH production (0–48 h p.i.). Total proteins isolated from SF-21 cells infected with the indicated virus were loaded (in equivalent volumes) on gels for protein blots and probed with either anti-BmCHI-h antibody (to detect CHIA), anti-V-CATH antibody (to detect proV-CATH), or anti-GAPDH antibody (to detect host GAPDH) as a loading control. Lane M is protein from uninfected cells as control. Lane G is protein from AcEGFP-infected cells. (<b>c</b>) CHIA and proV-CATH solubility. Lysates from infected cells were separated by centrifugation into detergent (0.5% NP-40/1% Triton-X)-soluble (SOL) and -insoluble (INSOL) fractions and proteins were detected with anti-BmCHI-h or anti-V-CATH antibody.</p>
Full article ">Figure 3
<p>Chitinase and cathepsin (protease) assays. Homogenates from infected SF-21 cells were used to compare viral chitinase (CHIA) and protease (V-CATH) activities at 48 h p.i. with AcMNPV-Rep, Acp6.9-chiA, and Acp6.9-chiA/polh-cath. Mock-infected (Mock) and AcEGFP-infected cells were used as negative controls. (<b>a</b>) Chitinase assay. Total protein (50 µg) was used to measure viral CHIA activity. (<b>b</b>) Protease assays. Total protein (400 µg) was used to measure V-CATH activity in the presence or absence of 20 µM of the cysteine protease inhibitor E-64.</p>
Full article ">Figure 4
<p>Budded virus production of AcEGFP-, AcMNPV-Rep-, Acp6.9-chiA-, and Acp6.9-chiA/polh-cath-infected cells. SF-21 cells were infected (m.o.i. = 0.1 PFU/cell), and at the indicated timepoints, a small amount (0.2 mL) of supernatant was removed and titrated by end-point dilution. The data points represent the mean budded virus titers from three replicates. Error bars denote standard deviations.</p>
Full article ">Figure 5
<p>Temporal CHIA and V-CATH protein production from infected <span class="html-italic">T. ni</span> larvae. (<b>a</b>) Carcass. CHIA (50 µg/lane) and V-CATH/proV-CATH (500 µg/lane) were immunodetected from homogenates of infected larvae collected at 3 and 4 days p.i., as indicated, using anti-BmCHI-h antibody (to detect CHIA) and anti-V-CATH antibody (to detect proV-CATH). (<b>b</b>) Hemolymph. CHIA and V-CATH/proV-CATH were immunodetected in pooled (3 larvae each) hemolymph samples extracted at 3 days p.i. The same volume of hemolymph (15 µL) from virus-infected larvae were loaded in each lane. Migration of protein markers is shown on the right in kDa.</p>
Full article ">Figure 6
<p>Chitinase and cathepsin (protease) assays from infected <span class="html-italic">T. ni</span> larval tissues. Total protein in larval homogenates was used to compare viral chitinase (CHIA) and protease (V-CATH) activities at 3, 4, and 5 days p.i. for AcEGFP, AcMNPV-Rep, Acp6.9-chiA, and Acp6.9-chiA/polh-cath. (<b>a</b>) Total protein (200 µg) was used to measure viral chitinase (CHIA) activity. ns, not significant (<b>b</b>) Protease assays. Total protein (400 µg) was used to measure V-CATH activity in the presence (white bars) or absence (black bars) of E-64 (20 µM) inhibitor. Statistics were computed only for V-CATH samples without E-64. The indicated <span class="html-italic">p</span>-values (<span class="html-italic">t</span>-test, 95% CI) were ≤ 0.001 (*), ≤ 0.002 (**), and ≤ 0.003 (***).</p>
Full article ">Figure 7
<p>Survival of infected fifth instar <span class="html-italic">T. ni</span> larvae. Larvae were injected with 10 µL containing 5 × 10<sup>4</sup> TCID<sub>50</sub> units of each virus and were monitored every 8 h for viability at 27 °C. About thirty insects per experiment were injected with each virus and monitored for viability by prodding with a blunt object. The survival curves combine data from three independent experiments. The data were analyzed using the Kaplan–Meier method, and the survival curves were found to be significantly different from each other by Log-rank test (<span class="html-italic">p</span> &lt; 0.0001). The dashed horizontal line indicates 50%. Error bars indicate standard error.</p>
Full article ">Figure 8
<p>Pathology of infected fifth Instar <span class="html-italic">T. ni</span> larvae. Larvae were injected with 10 µL containing 5 × 10<sup>4</sup> TCID<sub>50</sub> units of each virus and incubated at 27 °C in individual containers with synthetic diet. They were monitored every 24 h for pathological changes (i.e., live, dead, and liquefied). (<b>a</b>) Insects with characteristic phenotypes due to infection with each of the viruses were photographed at 4 and 5 days as indicated. (<b>b</b>) The graphed results combine data for 3 to 6 days p.i. for three independent experiments. Error bars indicate standard deviation.</p>
Full article ">
15 pages, 4558 KiB  
Article
Integrated Analysis of MicroRNA and mRNA Expression Profiles in the Fat Bodies of MbMNPV-Infected Helicoverpa armigera
by Zhenpu Liang, Yanqing Yang, Xiaoyan Sun, Junyang Du, Qiuyun Wang, Guozhi Zhang, Jiran Zhang, Xinming Yin, Deepali Singh, Ping Su and Xiaoxia Zhang
Viruses 2023, 15(1), 19; https://doi.org/10.3390/v15010019 - 21 Dec 2022
Cited by 5 | Viewed by 1557
Abstract
MicroRNAs (miRNAs), are a novel class of gene expression regulators, that have been found to participate in regulating host–virus interactions. However, the function of insect-derived miRNAs in response to virus infection is poorly understood. We analyzed miRNA expression profiles in the fat bodies [...] Read more.
MicroRNAs (miRNAs), are a novel class of gene expression regulators, that have been found to participate in regulating host–virus interactions. However, the function of insect-derived miRNAs in response to virus infection is poorly understood. We analyzed miRNA expression profiles in the fat bodies of Helicoverpa armigera (H. armigera) infected with Mamestra brassicae multiple nucleopolyhedroviruses (MbMNPV). A total of 52 differentially expressed miRNAs (DEmiRNAs) were filtered out through RNA-seq analysis. The targets of 52 DEmiRNAs were predicted and 100 miRNA–mRNA interaction pairs were obtained. The predicted targets of DEmiRNAs were mainly enriched in the Wnt signaling pathway, phagosome, and mTOR signaling pathway, which are related to the virus infection. Real-time PCR was used to verify the RNA sequencing results. ame-miR-317-3p, mse-miR-34, novel1-star, and sfr-miR-6094-5p were shown to be involved in the host response to MbMNPV infection. Results suggest that sfr-miR-6094-5p can negatively regulate the expression of four host genes eIF3-S7, CG7583, CG16901, and btf314, and inhibited MbMNPV infection significantly. Further studies showed that RNAi-mediated knockdown of eIF3-S7 inhibited the MbMNPV infection. These findings suggest that sfr-miR-6094-5p inhibits MbMNPV infection by negatively regulating the expression of eIF3-S7. This study provides new insights into MbMNPV and H. armigera interaction mechanisms. Full article
Show Figures

Figure 1

Figure 1
<p>Effect of MbMNPV infection on body weight of <span class="html-italic">H. armigera</span>. ns, <span class="html-italic">p</span> &gt; 0.05; ***, <span class="html-italic">p</span> &lt; 0.001; <span class="html-italic">p</span> &lt; 0.05 indicated the significant difference (Student’s <span class="html-italic">t</span>-test). Each experiment was performed in three replicates, and data are shown as mean ± s.e.m.</p>
Full article ">Figure 2
<p>Verification for infection of <span class="html-italic">H. armigera</span> with MbMNPV. (<b>A</b>) PCR verification of polyhedrin gene of mock-infected and MbMNPV-infected larvae. Note: 1~6 are the amplification of polyhedrin gene in CK-1, CK-2, and CK-3 and MbMNPV-1, MbMNPV-2, MbMNPV-3. (<b>B</b>) The copy number of MbMNPV in the mock-infected and MbMNPV-infected larvae fat body. *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 3
<p>MiRNAs distribution in samples. (<b>A</b>) The distribution of total miRNA in the two samples. (<b>B</b>) The distribution of known miRNA in the two samples. (<b>C</b>) The distribution of novel predicted miRNA in the samples.</p>
Full article ">Figure 4
<p>Length distribution of known miRNA.</p>
Full article ">Figure 5
<p>Screening and enrichment analysis of DEmiRNAs in MbMNPV-infected group compared with control group. (<b>A</b>) The volcano map of different miRNA. (<b>B</b>) Top 30 significantly enriched GO terms of predicted targets for DEmiRNAs in molecular function, cellular components, and biological process. (<b>C</b>) Hierarchical clustering analysis (heatmap) for DEmiRNAs. (<b>D</b>) Top 20 significantly enriched KEGG analyses of predicted targets for DEmiRNAs.</p>
Full article ">Figure 6
<p>Interaction networks of differentially expressed miRNAs and putative target mRNAs in fat body of MbMNPV-infected <span class="html-italic">H. armigera</span>. The square nodes and circular nodes represent miRNAs and mRNA, respectively. Downregulated and upregulated RNAs are indicated by green and red colors, respectively.</p>
Full article ">Figure 7
<p>Relative expression of miRNAs in control and MbMNPV. ***, <span class="html-italic">p</span> &lt; 0.001; <span class="html-italic">p</span> &lt; 0.05 indicate significant difference (Student’s <span class="html-italic">t</span>-test). Each experiment was performed in three replicates, and data are shown as mean ± s.e.m.</p>
Full article ">Figure 8
<p>Relative expression of miRNAs at different time points in the control and MbMNPV-infected group. ns, <span class="html-italic">p</span> &gt; 0.05; *, <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, <span class="html-italic">p</span> &lt; 0.05 indicate significant difference (Student’s <span class="html-italic">t</span>-test). Each experiment was performed in three replicates, and data are shown as mean ± s.e.m.</p>
Full article ">Figure 9
<p>The relative expression of eIF3-S7 (<b>A</b>), CG7583 (<b>B</b>), CG16901 (<b>C</b>) and btf314 (<b>D</b>) after injection of sfr-miR-6094-5p agomir and antagomir. ns, <span class="html-italic">p</span> &gt; 0.05; *, <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; <span class="html-italic">p</span> &lt; 0.05 indicate significant difference (Student’s <span class="html-italic">t</span>-test). Each experiment was performed in three replicates, and data are shown as mean ± s.e.m.</p>
Full article ">Figure 10
<p>The relative expression of MbMNPV polyhedrin gene after injection of sfr-miR-6094-5p agomir and antagomir in the MbMNPV-infected larvae. *, <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; 0.05 indicate significant difference (Student’s <span class="html-italic">t</span>-test). Each experiment was performed in three replicates, and data are shown as mean ± s.e.m.</p>
Full article ">Figure 11
<p>The effects of HaeIF3-S7 RNAi on the larvae. At 1d fifth-instar, larvae were injected with dsRNA of HaeIF3-S7 (ds eIF3-S7), dsRNA of green fluorescent protein (dsGFP), and DEPC (control). (<b>A</b>) The expression levels of the eIF3-S7 gene. (<b>B</b>) The expression levels of the polyhedrin gene. (<b>C</b>,<b>D</b>) The MbMNPV-infected larvae weight were analyzed. The standard error is represented by the error bar. ns, <span class="html-italic">p</span> &gt; 0.05; *, <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; <span class="html-italic">p</span> &lt; 0.05 indicate significant difference (Student’s <span class="html-italic">t</span>-test). Each experiment was performed in three replicates, and data are shown as mean ± s.e.m.</p>
Full article ">
14 pages, 3113 KiB  
Article
Improved Expression of SARS-CoV-2 Spike RBD Using the Insect Cell-Baculovirus System
by Joaquín Poodts, Ignacio Smith, Joaquín Manuel Birenbaum, María Sol Rodriguez, Luciano Montero, Federico Javier Wolman, Juan Ignacio Marfía, Silvina Noemí Valdez, Leonardo Gabriel Alonso, Alexandra Marisa Targovnik and María Victoria Miranda
Viruses 2022, 14(12), 2794; https://doi.org/10.3390/v14122794 - 15 Dec 2022
Cited by 4 | Viewed by 2300
Abstract
Insect cell-baculovirus expression vector system is one of the most established platforms to produce biological products, and it plays a fundamental role in the context of COVID-19 emergency, providing recombinant proteins for treatment, diagnosis, and prevention. SARS-CoV-2 infection is mediated by the interaction [...] Read more.
Insect cell-baculovirus expression vector system is one of the most established platforms to produce biological products, and it plays a fundamental role in the context of COVID-19 emergency, providing recombinant proteins for treatment, diagnosis, and prevention. SARS-CoV-2 infection is mediated by the interaction of the spike glycoprotein trimer via its receptor-binding domain (RBD) with the host’s cellular receptor. As RBD is required for many applications, in the context of pandemic it is important to meet the challenge of producing a high amount of recombinant RBD (rRBD). For this reason, in the present study, we developed a process based on Sf9 insect cells to improve rRBD yield. rRBD was recovered from the supernatant of infected cells and easily purified by metal ion affinity chromatography, with a yield of 82% and purity higher than 95%. Expressed under a novel chimeric promoter (polh-pSeL), the yield of rRBD after purification was 21.1 ± 3.7 mg/L, which is the highest performance described in Sf9 cell lines. Finally, rRBD was successfully used in an assay to detect specific antibodies in COVID-19 serum samples. The efficient strategy herein described has the potential to produce high-quality rRBD in Sf9 cell line for diagnostic purpose. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Recombinant baculovirus (<span class="html-italic">Acpolh-pSeL</span>-gprRBD) for the expression of rRBD under the <span class="html-italic">polh-pSeL</span> promoter. GP64: viral secretion signal GP64; His-tag: six histidine tag. rRBD: SARS-CoV-2 receptor binding domain sequence (Wuhan-Hu-1 isolate) optimized for insects.</p>
Full article ">Figure 2
<p>Analysis of rRBD in supernatants by SDS-PAGE under reducing (<b>A</b>–<b>C</b>) and non-reducing (<b>D</b>) conditions followed by western blot developed with anti-His (α-His) and anti-S (α-S) antibodies. Sf9 cells were infected with <span class="html-italic">Acpolh-pSeL</span>-gprRBD. At different days post-infection, the culture medium was harvested and analyzed. Lanes: 1–4, culture supernatant from 1–4 days post-infection; M, protein marker; D (%): ratio of dimer to total rRBD expressed as determined by densitometry analysis. M (%): ratio of monomer to total rRBD expressed as determined by densitometry analysis.</p>
Full article ">Figure 3
<p>IMAC purification of rRBD. (<b>A</b>) SDS-PAGE analysis in the reducing condition of fraction collected during the purification process. (<b>B</b>) Western blot analysis of fraction collected during the purification process using anti-S antibody. Lanes: M, protein marker; 1, Sf9 cell expression supernatant; 2, diafiltrated sample (Input); 3, flow-through; 4, washing step (equilibration buffer with 80 mM imidazole); 5, IMAC fraction eluted by 500 mM imidazole.</p>
Full article ">Figure 4
<p>RP-HPLC analysis of rRBD purified by IMAC.</p>
Full article ">Figure 5
<p>MALDI-TOF MS analysis of purified rRBD.</p>
Full article ">Figure 6
<p>Characterization of purified rRBD. (<b>A</b>) rRBD glycosylation analysis. N-glycosidase F-mediated in-vitro deglycosylation of purified rRBD. Western blot in reducing condition developed with anti-S antibody. (<b>B</b>) Analysis of purified rRBD by SDS-PAGE in non-reducing condition. Lanes: M, protein marker; rRBD, purified rRBD; rRBD + PNGase. F, purified rRBD treated with N-glycosidase F.</p>
Full article ">Figure 7
<p>Size exclusion chromatography analysis of rRBD purified by IMAC. Chromatogram showing the elution profile of rRBD (<b>A</b>). The number indicates the peak containing rRBD dimer (1) and monomer (2). The peaks were analyzed by reducing SDS-PAGE (<b>B</b>) and western blot developed with specific anti-S antibody (<b>C</b>). Lanes: M, protein marker; 1, peak 1; 2, peak 2.</p>
Full article ">Figure 8
<p>Immunoreactivity of rRBD produced in Sf9 cell line by b-ELISA. SARS-CoV-2 antibody test results obtained by b-ELISA from pre-pandemic samples (n = 28) and samples obtained from seropositive COVID-19 patients (n = 30). Results are expressed as SDs. The cut-off value (SDs &gt; 5.0) is indicated by a dotted line and medians for each population are indicated as a full line (*** <span class="html-italic">p</span> &lt; 0.0001, statistically significant). NHS: normal human sera. COVID-19+: seropositive COVID-19 patients.</p>
Full article ">
13 pages, 4360 KiB  
Article
Successful Rescue of Synthetic AcMNPV with a ~17 kb Deletion in the C1 Region of the Genome
by Yijia Guo, Hengrui Hu, Han Xiao, Fei Deng, Jiang Li, Manli Wang and Zhihong Hu
Viruses 2022, 14(12), 2780; https://doi.org/10.3390/v14122780 - 13 Dec 2022
Cited by 1 | Viewed by 1892
Abstract
Baculoviruses have been widely used as expression vectors. However, numerous genes in the baculoviral genome are non-essential for cellular infection and protein expression, making the optimisation of baculovirus expression vectors possible. We used a synthetic biological method to reduce the number of genes [...] Read more.
Baculoviruses have been widely used as expression vectors. However, numerous genes in the baculoviral genome are non-essential for cellular infection and protein expression, making the optimisation of baculovirus expression vectors possible. We used a synthetic biological method to reduce the number of genes in a partial region of the autograph californica multiple nucleopolyhedrovirus (AcMNPV), the most widely used baculovirus expression vector. The C1 region of the AcMNPV is 46.4 kb and is subdivided into B1, B2, and B3 fragments. We first designed modified B1, B2, and B3 fragments by deleting the non-essential genes, and then synthesised complete viral genomes containing either individual modified B fragments or joint modified B fragments through transformation-related recombination in yeast. The synthetic genomes were then transfected into Sf9 cells to rescue the progeny viruses and test their infectivity. The design-build-test cycle was repeated until the ultimately rescued virus could produce progeny viruses efficiently. Finally, AcMNPV-Syn-mC1-1.1 by deleting approximately 17.2 kb, including 20 ORFs, in the C1 region, was obtained. This is essential to the synthesis of a minimal AcMNPV genome that can generate infectious progeny viruses and can be further used to optimise the foundation of baculovirus expression vectors. Full article
Show Figures

Figure 1

Figure 1
<p>Flowchart of constructing AcMNPV genome containing modified C1 fragment. (<b>A</b>) Circular maps of the parental genome of AcMNPV-WIV-Syn1 and the modified AcMNPV-Syn-mC1-1.1. (<b>B</b>) The design-build-test cycling method to modify C1 fragment. The design of the modified C1 was based on the modification of B1, B2, and B3 fragments. Non-essential genes of each fragment were removed to generate modified fragments of mB1-1.0, mB2-1.0, and mB3-1.0. mB-1.0, mB2-1.0, and mB3-1.0 by overlapping PCR and used to generate modified genomes of AcMNPV-Syn-mB1-1.0, AcMNPV-Syn-mB2-1.0 and AcMNPV-Syn-mB3-1.0, respectively, using two steps of TAR in yeast. Meanwhile, AcMNPV-Syn-mC1-1.0 was generated using mB1-1.0, mB2-1.0, and mB3-1.0. To test, transfection and infection were performed; and the modified genome, which could not generate efficient progeny viruses, was redesigned and subjected to the design-build-test cycle again.</p>
Full article ">Figure 2
<p>Design and characterization of version 1.0 of the modified genomes. (<b>A</b>) Design of mB1-1.0, mB2-1.0, and mB3-1.0. The modified fragments are shown on the backbones of the original B1, B2, and B3. The deleted genes and <span class="html-italic">hr</span>s are coloured white, the reserved genes green, and the reserved <span class="html-italic">hr1a</span> blue. The primers used for overlap extension PCR are indicated. (<b>B</b>) Transfection and infection results of AcMNPV-Syn-mB1-1.0, AcMNPV-Syn-mB2-1.0, AcMNPV-Syn-mB3-1.0, and AcMNPV-Syn-mC1-1.0. The images were obtained at 96 h post-transfection (p.t.) and 96 h post-infection (p.i.). The green dots represent infected Sf9 cells.</p>
Full article ">Figure 3
<p>Redesign and characterization of AcMNPV-Syn-mB1-1.1, AcMNPV-Syn-mB2-1.1, and AcMNPV-Syn-mC1-1.1. (<b>A</b>) Design of mB1-1.1 and mB2-1.1. The modified fragments are shown on the backbones of the original B1 and B2. The colouring is similar to that in <a href="#viruses-14-02780-f002" class="html-fig">Figure 2</a>A, with the eight repaired ORFs showed in brown. (<b>B</b>) Transfection and infection results of AcMNPV-Syn-mB1-1.1, AcMNPV-Syn-mB2-1.1, AcMNPV-Syn-mC1-1.1, and AcMNPV-WIV-Syn-1. The images were obtained at 96 h p.t. and 96 h p.i. The green dots represent infected Sf9 cells.</p>
Full article ">Figure 4
<p>Restriction enzyme analyses of the genomes of AcMNPV-Syn-mC1-1.1 and AcMNPV –WIV-Syn1. (<b>A</b>) Physical maps of the AcMNPV-Syn-mC1-1.1 and AcMNPV-WIV-Syn1 genomes digested with BglII, XhoI or SacII. The fragments were named alphabetically according to size. The fragments corresponding to the C1 region in AcMNPV-WIV-Syn1 and the modified C1 region in AcMNPV-Syn-mC1-1.1 are highlighted. (<b>B</b>) Computer-simulated restriction enzyme profiles of AcMNPV-WIV-Syn1 and AcMNPV-Syn-mC1-1.1. (<b>C</b>) Electrophoresis results of the restriction enzyme-digested DNA samples of AcMNPV-WIV-Syn1 and AcMNPV-Syn-mC1-1.1 in 1% agarose gel. Fragments with size change are indicated in red letters in AcMNPV-Syn-mC1-1.1 and the original fragments in AcMNNPV-WIV-Syn1 are marked with a red star.</p>
Full article ">Figure 5
<p>Characterization of AcMNPV-Syn-mC1-1.1. (<b>A</b>) One-step growth curves of AcMNPV-WIV-Syn1 and AcMNPV-Syn-mC1-1.1. Sf9 cells were infected with AcMNPV-WIV-Syn1 or AcMNPV-Syn-mC1-1.1 at an MOI of 1, the supernatants were collected at the indicated time points and viral titres were titrated by EPDA. The experiment was carried out in triplicate, and the titre was transformed logarithmically after averaging. Statistical significance was determined by 2way ANOVA. ***: <span class="html-italic">p</span> &lt; 0.001. (<b>B</b>) Results of electron microscopy. Sf9 cells were infected with AcMNPV-WIV-Syn1 or AcMNPV-Syn-mC1-1.1 at an MOI of 5, and cells were collected at 24, 48, and 72 h p.i. for electron microscopy. The representative nucleocapsids are marked with arrows. OB: occlusion body. Scale bars are 2 μm and 1 μm.</p>
Full article ">
18 pages, 3086 KiB  
Article
Utility of Alternative Promoters for Foreign Gene Expression Using the Baculovirus Expression Vector System
by Mark R. Bruder and Marc G. Aucoin
Viruses 2022, 14(12), 2670; https://doi.org/10.3390/v14122670 - 29 Nov 2022
Cited by 4 | Viewed by 2143
Abstract
The baculovirus expression vector system (BEVS) is a widely used platform for recombinant protein production for use in a wide variety of applications. Of particular interest is production of virus-like particles (VLPs), which consist of multiple viral proteins that self-assemble in strict stoichiometric [...] Read more.
The baculovirus expression vector system (BEVS) is a widely used platform for recombinant protein production for use in a wide variety of applications. Of particular interest is production of virus-like particles (VLPs), which consist of multiple viral proteins that self-assemble in strict stoichiometric ratios to mimic the structure of a virus but lacks its genetic material, while a significant amount of effort has been spent on optimizing expression ratios by co-infecting cells with multiple recombinant BEVs and modulating different process parameters, co-expressing multiple foreign genes from a single rBEV may offer more promise. However, there is currently a lack of promoters available with which to optimize co-expression of each foreign gene. To address this, previously published transcriptome data was used to identify promoters that have incrementally lower expression profiles and compared by expressing model cytoplasmic and secreted proteins. Bioinformatics was also used to identify sequence determinants that may be important for late gene transcription regulation, and translation initiation. The identified promoters and bioinformatics analyses may be useful for optimizing expression of foreign genes in the BEVS. Full article
Show Figures

Figure 1

Figure 1
<p>The promoters included on commercially available transfer plasmids have drastically different transcription profiles. (<b>Left panel</b>). The transcript abundance of <span class="html-italic">Ac</span>MNPV polh, p10, gp64, and ie1 ORFs, which are among the only promoters available on commercial transfer plasmids for foreign gene expression. (<b>Right panel</b>). The transcript abundance profiles of <span class="html-italic">Ac</span>MNPV ORFs selected for evaluation of the upstream promoter regions in this study. Promoters were selected for expression profiles between polh/p10 and gp64/ie1.</p>
Full article ">Figure 2
<p>Production of intracellular GFP from selected <span class="html-italic">Ac</span>MNPV promoters. (<b>A</b>) Median fluorescence intensity measured using flow cytometry and (<b>B</b>) relative transcript abundance measured using RT-qPCR at various times post infection.</p>
Full article ">Figure 3
<p>Production of extracellular SEAP from selected <span class="html-italic">Ac</span>MNPV promoters. (<b>A</b>) Yield of SEAP (mg/L) of culture supernatants measured using a colorimetric SEAP activity assay and (<b>B</b>) relative transcript abundance measured using RT-qPCR at 24, 48, and 72 h post infection.</p>
Full article ">Figure 4
<p>Evaluation of the 5<math display="inline"><semantics> <msup> <mrow/> <mo>′</mo> </msup> </semantics></math>UTRs of <span class="html-italic">Ac</span>MNPV ORFs categorized according to transcript abundance. (<b>A</b>) Length (in nucleotides) and (<b>B</b>) A/T content of the 5<math display="inline"><semantics> <msup> <mrow/> <mo>′</mo> </msup> </semantics></math>UTR between the late gene promoter motif (5<math display="inline"><semantics> <msup> <mrow/> <mo>′</mo> </msup> </semantics></math>-TAAG-3<math display="inline"><semantics> <msup> <mrow/> <mo>′</mo> </msup> </semantics></math>) and translation initiation codon (ATG). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 5
<p>Circular chromosome map of <span class="html-italic">Ac</span>MNPV ORFs colour-coded according to transcript abundance.</p>
Full article ">Figure 6
<p>Distance (in nucleotides) between the start codon and 5<math display="inline"><semantics> <msup> <mrow/> <mo>′</mo> </msup> </semantics></math> end of each homologous region for every <span class="html-italic">Ac</span>MNPV ORF categorized according to transcript abundance. The distance was calculated by subtracting the genomic location of the 5<math display="inline"><semantics> <msup> <mrow/> <mo>′</mo> </msup> </semantics></math> end of each <span class="html-italic">Ac</span>MNPV ORF from the genomic location of the 5<math display="inline"><semantics> <msup> <mrow/> <mo>′</mo> </msup> </semantics></math> end of each <span class="html-italic">hr</span>. Positive values represent ORFs located behind (clockwise) to the <span class="html-italic">hr</span> and negative values represent distances between ORFs that are located in front of (counterclockwise) the <span class="html-italic">hr</span>.</p>
Full article ">Figure 7
<p>The approximate positions of the upstream octamer (blue) and downstream octamer (red) in regions upstream of the late gene promoter motif for the most abundant <span class="html-italic">Ac</span>MNPV ORFs.</p>
Full article ">Figure 8
<p>Consensus sequences calculated from multiple sequence alignments. (<b>A</b>) Consensus sequence for the nucleotide sequences flanking the translation initiation site and (<b>B</b>) the late gene promoter motif. Consensus sequences were calculated from multiple sequence alignments of sequences extracted from <span class="html-italic">Ac</span>MNPV ORFs that were categorized according to transcript abundance.</p>
Full article ">
14 pages, 2227 KiB  
Article
Overexpression Bombyx mori HEXIM1 Facilitates Immune Escape of Bombyx mori Nucleopolyhedrovirus by Suppressing BmRelish-Driven Immune Responses
by Guanping Chen, Yuedong Li, Xiangshuo Kong, Shudi Zhao, Jiale Li and Xiaofeng Wu
Viruses 2022, 14(12), 2636; https://doi.org/10.3390/v14122636 - 25 Nov 2022
Viewed by 1581
Abstract
Bombyx mori nucleopolyhedrovirus (BmNPV), a typical arthropod-specific enveloped DNA virus, is one of the most serious pathogens in silkworm farming, but the potential mechanisms of the evasion of innate immune responses from BmNPV infection are still poorly understood. HEXIM1 is an RNA-binding protein, [...] Read more.
Bombyx mori nucleopolyhedrovirus (BmNPV), a typical arthropod-specific enveloped DNA virus, is one of the most serious pathogens in silkworm farming, but the potential mechanisms of the evasion of innate immune responses from BmNPV infection are still poorly understood. HEXIM1 is an RNA-binding protein, best known as an inhibitor of positive transcription elongation factor b (P-TEFb), which controls transcription elongation by RNA polymerase II. In this study, Bombyx mori HEXIM1 (BmHEXIM1) was cloned and characterized, and its expression was found to be remarkably upregulated after BmNPV infection. Furthermore, BmHEXIM1 was detected to increase the proliferation of BmNPV, and its full length is essential for assisting BmNPV immune escape by suppressing BmRelish-driven immune responses. This study brought new insights into the mechanisms of immune escape of BmNPV and provided theoretical guidance for the breeding of BmNPV-resistant silkworm varieties. Full article
Show Figures

Figure 1

Figure 1
<p>BmHEXIM1 showed a significant response to BmNPV infection. (<b>A</b>) The expression levels of BmHEXIM1 in BmN cells at different hours post infection with BmNPV. BmN cells without BmNPV infection as the control. (<b>B</b>) The expression levels of BmHEXIM1 in different tissues of silkworm post infection with BmNPV. Fat body of silkworm without BmNPV infection as the control. (<b>C</b>) Time course analysis of subcellular localization of BmHEXIM1 by immunofluorescence in BmN cells infected by BmNPV; DAPI, blue; and BmHEXIM1, green. The mRNA level of target genes was normalized to the internal control (BmRPL32). Data represent mean ± SEM of the three independent experiments. The number of asterisks represents the degree of significance with respect to <span class="html-italic">p</span>-value. <span class="html-italic">P</span>-values were provided as * <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; **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 2
<p>RNAi down-regulated BmHEXIM1 and inhibited the invasion of BmNPV. (<b>A</b>) The RNAi of the siBmHEXIM1 RNA decreases the expression level of protein HEXIM1 at 48 h post-transfection. (<b>B</b>) TCID<sub>50</sub> end-point dilution was used to evaluate the production of infectious BV at 24 h and 48 h. (<b>C</b>) The infected cells (the cells with the green fluorescence) were observed by an inverted fluorescence microscope at 24 h and 48 h (bar = 100 μm). (<b>D</b>) BmN cells were transfected with si-NC and siBmHEXIM1 RNA for 48 h. Then, the cells were infected with BmNPV for 48 h. The viral gene’s expression was measured by qRT-PCR analysis. (<b>E</b>) The BmNPV abundance was assessed by analyzing the expression of Bm14 by Western blot after knockdown of BmHEXIM1. The mRNA level of target genes was normalized to the internal control (BmRPL32). Data represent mean ± SEM of the three independent experiments. The number of asterisks represents the degree of significance with respect to <span class="html-italic">p</span>-value. <span class="html-italic">p</span>-values were provided as * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 3
<p>BmHEXIM1 promoted BmNPV proliferation, and its full length was essential to promote viral transcription. (<b>A</b>) BmN cells were transfected with pIZ/V5-BmHEXIM1 expressing plasmid, and the expression of BmHEXIM1 was analyzed by Western blot at 24 h, 48 h, 72 h, and 96 h. (<b>B</b>) TCID<sub>50</sub> end-point dilution was used to evaluate the production of infectious BV at 24 h and 48 h. (<b>C</b>) The infected cells (the cells with green fluorescence) were observed by an inverted fluorescence microscope at 24 h and 48 h (bar = 100 μm). (<b>D</b>) Determination of the region of BmHEXIM1 promoted BmNPV proliferation. (<b>E</b>) BmN cells were transfected with pIZ/V5-BmHEXIM1, pIZ/V5-BmHEXIM1-1, and pIZ/V5-BmHEXIM1-2 expressing plasmid for 48 h. Then, the cells were infected with BmNPV for 48 h. The viral gene’s expression was measured by qRT-PCR analysis. (<b>F</b>) BmNPV abundance was assessed by analyzing the expression levels of Bm14 by Western blot. The mRNA level of target genes was normalized to the internal control (BmRPL32). Data represent mean ± SEM of the three independent experiments. The number of asterisks represents the degree of significance with respect to <span class="html-italic">p</span>-value. <span class="html-italic">P</span>-values were provided as * <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; **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 4
<p>BmHEXIM1 suppresses host innate immunity during BmNPV infection. (<b>A</b>–<b>D</b>) Quantification of the expression of BmSTING, BmRelish, BmcecA, and BmcecB by qRT-PCR after overexpression of BmHEXIM1, BmHEXIM1-1, BmHEXIM1-2, or knockdown of BmHEXIM1 in BmN cells after BmNPV infection. (<b>E</b>) The phosphorylation level of RNP Ⅱ by Western blot after overexpression of BmHEXIM1, BmHEXIM1-1, BmHEXIM1-2, or knockdown of BmHEXIM1 in BmN cells. (<b>F</b>–<b>I</b>) In the absence or presence of PMA, the expression of BmSTING, BmRelish, BmcecA, and BmcecB was also analyzed by qRT-PCR or by the co-expression of HEXIM1. The mRNA level of target genes was normalized to the internal control (BmRPL32). Data represent mean ± SEM of the three independent experiments. The number of asterisks represents the degree of significance with respect to <span class="html-italic">p</span>-value. <span class="html-italic">p</span>-values were provided as * <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; **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 5
<p>HMBA inhibited BmRelish-mediated pathway and promoted BmNPV proliferation. (<b>A</b>) BmN cells were stimulated with increasing concentrations of HMBA (1, 5, 10, 20, and 30 mM) for 24 h, and the expression levels of BmHEXIM1 were measured by qRT-PCR analysis. (<b>B</b>) BmN cells stimulated or not with HMBA (10 mM) for 12 h, 24 h, and 48 h and the expression of BmHEXIM1 were measured by qRT-PCR analysis. (<b>C</b>) BmN cells stimulated with HMBA (10 mM) for 24 h and the expression of viral genes were measured by qRT-PCR analysis. (<b>D</b>) BmN cells stimulated with HMBA (10 mM) for 24 h, and then added BmNPV for 48 h. BmNPV abundance was assessed by analyzing the expression of Bm14 by Western blot. (<b>E</b>) BmN cells stimulated with HMBA (10 mM) for 24 h and the expression levels of BmSTING, BmRelish, BmcecA, and BmcecB were measured by qRT-PCR analysis. The mRNA level of target genes was normalized to the internal control (BmRPL32). Data represent mean ± SEM of the three independent experiments. The number of asterisks represents the degree of significance with respect to <span class="html-italic">p</span>-value. <span class="html-italic">p</span>-values were provided as ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001; **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 6
<p>Schematic representation of the putative mechanism(s) of BmHEXIM1 in assisting BmNPV evade antiviral immune response and promoting BmNPV proliferation. (A) BmNPV significantly induced BmHEXIM1 expression, while activating host antiviral pathways, such as STING and Relish. The innate immune pathway mediated by BmRelish can induce the expression of BmSTING antimicrobial peptide genes (BmcecA and BmcecB) to resist BmNPV infection. (B) Altered subcellular localization of BmHEXIM1 may be regulated by ubiquitination following viral infection. “? “ were provided as “not determined”. (C) BmHEXIM1 inhibits BmRelish-dependent transcription, leading to suppression of the host’s antiviral response. (D) Stimulation of PMA inhibits BmHEXIM1 to repress BmRelish-dependent transcription. (E) HMBA treatment induces BmHEXIM1 expression to suppress host antiviral response.</p>
Full article ">

Review

Jump to: Research

19 pages, 953 KiB  
Review
The History of Baculovirology in Africa
by Sean Moore and Michael Jukes
Viruses 2023, 15(7), 1519; https://doi.org/10.3390/v15071519 - 7 Jul 2023
Cited by 2 | Viewed by 1427
Abstract
Baculovirology has been studied on the African continent for the development of insect virus-based biopesticides and, to a much lesser extent, vaccine production and delivery, since the 1960s. In this review, we focus only on baculoviruses as biopesticides for agricultural pests in Africa. [...] Read more.
Baculovirology has been studied on the African continent for the development of insect virus-based biopesticides and, to a much lesser extent, vaccine production and delivery, since the 1960s. In this review, we focus only on baculoviruses as biopesticides for agricultural pests in Africa. At least 11 species of baculovirus have been discovered or studied on the African continent, some with several distinct isolates, with the objective in most cases being the development of a biopesticide. These include the nucleopolyhedroviruses of Helicoverpa armigera, Cryptophlebia peltastica, Spodoptera exempta, Spodoptera frugiperda, Spodoptera littoralis, and Maruca vitrata, as well as the granuloviruses of Cydia pomonella, Plutella xylostella, Thaumatotibia (Cryptophlebia) leucotreta, Choristoneura occidentalis, and Phthorimaea operculella. Eleven different baculovirus-based biopesticides are recorded as being registered and commercially available on the African continent. Baculoviruses are recorded to have been isolated, researched, utilised in field trials, and/or commercially deployed as biopesticides in at least 13 different African countries. Baculovirus research is ongoing in Africa, and researchers are confident that further novel species and isolates will be discovered, to the benefit of environmentally responsible agricultural pest management, not only in Africa but also elsewhere. Full article
Show Figures

Figure 1

Figure 1
<p>Overview of African countries where baculoviruses have been isolated, researched, utilised in field trials, and/or commercially deployed as biopesticides.</p>
Full article ">
17 pages, 848 KiB  
Review
Advances in CRISPR-Cas9 for the Baculovirus Vector System: A Systematic Review
by Duygu Sari-Ak, Omar Alomari, Raghad Al Shomali, Jackwee Lim and Deepak B. Thimiri Govinda Raj
Viruses 2023, 15(1), 54; https://doi.org/10.3390/v15010054 - 24 Dec 2022
Cited by 1 | Viewed by 4168
Abstract
The baculovirus expression vector systems (BEVS) have been widely used for the recombinant production of proteins in insect cells and with high insert capacity. However, baculovirus does not replicate in mammalian cells; thus, the BacMam system, a heterogenous expression system that can infect [...] Read more.
The baculovirus expression vector systems (BEVS) have been widely used for the recombinant production of proteins in insect cells and with high insert capacity. However, baculovirus does not replicate in mammalian cells; thus, the BacMam system, a heterogenous expression system that can infect certain mammalian cells, was developed. Since then, the BacMam system has enabled transgene expression via mammalian-specific promoters in human cells, and later, the MultiBacMam system enabled multi-protein expression in mammalian cells. In this review, we will cover the continual development of the BEVS in combination with CRPISPR-Cas technologies to drive genome-editing in mammalian cells. Additionally, we highlight the use of CRISPR-Cas in glycoengineering to potentially produce a new class of glycoprotein medicines in insect cells. Moreover, we anticipate CRISPR-Cas9 to play a crucial role in the development of protein expression systems, gene therapy, and advancing genome engineering applications in the future. Full article
Show Figures

Figure 1

Figure 1
<p>Evolution of Baculovirus expression vector system (BEVS), MultiBac technology and CRISPR/Cas9 system. MultiBac has made it simple to access baculoviral genomes ever since it was first used, in addition to the incorporation of CRISPR components on both natural and synthetic insect cells, and minimized genomes (SynBac). fdl: fused lobes gene; SfSWT-1: commercial transgenic insect cell line called mimic Sf9; YFP: Yellow Fluorescent Protein.</p>
Full article ">
Back to TopTop