[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (49)

Search Parameters:
Keywords = DNA replicon

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
24 pages, 2287 KiB  
Review
A Comprehensive Review of Our Understanding and Challenges of Viral Vaccines against Swine Pathogens
by Aman Kamboj, Shaurya Dumka, Mumtesh Kumar Saxena, Yashpal Singh, Bani Preet Kaur, Severino Jefferson Ribeiro da Silva and Sachin Kumar
Viruses 2024, 16(6), 833; https://doi.org/10.3390/v16060833 - 24 May 2024
Viewed by 1312
Abstract
Pig farming has become a strategically significant and economically important industry across the globe. It is also a potentially vulnerable sector due to challenges posed by transboundary diseases in which viral infections are at the forefront. Among the porcine viral diseases, African swine [...] Read more.
Pig farming has become a strategically significant and economically important industry across the globe. It is also a potentially vulnerable sector due to challenges posed by transboundary diseases in which viral infections are at the forefront. Among the porcine viral diseases, African swine fever, classical swine fever, foot and mouth disease, porcine reproductive and respiratory syndrome, pseudorabies, swine influenza, and transmissible gastroenteritis are some of the diseases that cause substantial economic losses in the pig industry. It is a well-established fact that vaccination is undoubtedly the most effective strategy to control viral infections in animals. From the period of Jenner and Pasteur to the recent new-generation technology era, the development of vaccines has contributed significantly to reducing the burden of viral infections on animals and humans. Inactivated and modified live viral vaccines provide partial protection against key pathogens. However, there is a need to improve these vaccines to address emerging infections more comprehensively and ensure their safety. The recent reports on new-generation vaccines against swine viruses like DNA, viral-vector-based replicon, chimeric, peptide, plant-made, virus-like particle, and nanoparticle-based vaccines are very encouraging. The current review gathers comprehensive information on the available vaccines and the future perspectives on porcine viral vaccines. Full article
(This article belongs to the Special Issue Novel Vaccines for Porcine Viruses)
Show Figures

Figure 1

Figure 1
<p>Milestones in animal vaccine development.</p>
Full article ">Figure 2
<p>Available veterinary viral vaccines.</p>
Full article ">Figure 3
<p>Schematic representation of the working principle of the viral vector vaccine.</p>
Full article ">Figure 4
<p>Overview of vaccine administration routes in pigs.</p>
Full article ">
27 pages, 4778 KiB  
Article
West Nile Virus Subgenomic RNAs Modulate Gene Expression in a Neuronal Cell Line
by Maria Bampali, Adamantia Kouvela, Nikolaos Kesesidis, Katerina Kassela, Nikolas Dovrolis and Ioannis Karakasiliotis
Viruses 2024, 16(5), 812; https://doi.org/10.3390/v16050812 - 20 May 2024
Cited by 1 | Viewed by 1197
Abstract
Subgenomic flaviviral RNAs (sfRNAs) are small non-coding products of the incomplete degradation of viral genomic RNA. They accumulate during flaviviral infection and have been associated with many functional roles inside the host cell. Studies so far have demonstrated that sfRNA plays a crucial [...] Read more.
Subgenomic flaviviral RNAs (sfRNAs) are small non-coding products of the incomplete degradation of viral genomic RNA. They accumulate during flaviviral infection and have been associated with many functional roles inside the host cell. Studies so far have demonstrated that sfRNA plays a crucial role in determining West Nile virus (WNV) pathogenicity. However, its modulatory role on neuronal homeostasis has not been studied in depth. In this study, we investigated the mechanism of sfRNA biosynthesis and its importance for WNV replication in neuronal cells. We found that sfRNA1 is functionally redundant for both replication and translation of WNV. However, the concurrent absence of sfRNA1 and sfRNA2 species is detrimental for the survival of the virus. Differential expression analysis on RNA-seq data from WT and ΔsfRNA replicon cell lines revealed transcriptional changes induced by sfRNA and identified a number of putative targets. Overall, it was shown that sfRNA contributes to the viral evasion by suppressing the interferon-mediated antiviral response. An additional differential expression analysis among replicon and control Neuro2A cells also clarified the transcriptional changes that support WNV replication in neuronal cells. Increased levels of translation and oxidative phosphorylation, post-translational modification processes, and activated DNA repair pathways were observed in replicon cell lines, while developmental processes such as axonal growth were deficient. Full article
(This article belongs to the Section Invertebrate Viruses)
Show Figures

Figure 1

Figure 1
<p>Coverage map of West Nile virus (WNV) replicon 3′ UTR showing the starting points of subgenomic flaviviral RNA (sfRNA) species. Normalized reads per nucleotide position are shown for the 3′ UTR of WT (orange), ΔsfRNA1 (magenta), and ΔsfRNA2 (cyan) replicons. The position of inserted mutations (mut1, mut2) is shown with pink dashed lines. Sharp fluctuation of counts reflects sfRNA start. Black dashed lines indicate the starting points of sfRNAs. The structure of the replicon constructs are also presented as an inset in the figure (WT, ΔsfRNA1, ΔsfRNA2).</p>
Full article ">Figure 2
<p>Translation and replication efficiency of WT and ΔsfRNA replicons in Neuro2A cells. (<b>A</b>) Comparison of the translation efficiency for WT and ΔsfRNA replicons. Renilla luciferase measurements are indicative of their translation capability. Luciferase was measured over the span of four days at indicated time points. The expression was normalized to that of a co-transfected Firefly luciferase plasmid. (<b>B</b>) Relative abundance of the genomic and antigenomic replicon RNA of WT and ΔsfRNA replicons at 48 h post-transfection. <span class="html-italic">YWHAZ</span> expression was used for normalization. (<b>C</b>) Relative abundance of the genomic and antigenomic replicon RNA in the WT and ΔsfRNA stable replicon cell lines. (<b>D</b>) Relative abundance of sfRNA in the WT and ΔsfRNA stable replicon cell lines. Viral NS1 gene expression was used for normalization. Error bars correspond to the standard deviation from 3 technical replicates. * <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 (Student’s <span class="html-italic">t</span> test).</p>
Full article ">Figure 3
<p>Differentially regulated genes and altered pathways among the WT and ΔsfRNA replicon cell lines. (<b>A</b>) Differentially expressed genes in the ΔsfRNA1 (<b>left</b> image) and ΔsfRNA2 (<b>right</b> image) replicon cell lines plotted as volcano plots. (<b>B</b>) Top downregulated and upregulated genes of the ΔsfRNA1 and ΔsfRNA2 replicon cell lines when compared against the WT replicon cell line. (<b>C</b>) Pathway enrichment analysis results for upregulated and downregulated genes in the ΔsfRNA1 and ΔsfRNA2 replicon cell lines. (<b>D</b>) Cnetplot depicting the top 10 upregulated and downregulated Reactome pathways in ΔsfRNA1 and ΔsfRNA2 cells as networks that connect the genes with the respective pathways. (<b>E</b>) qPCR validation results for a selected set of genes. Error bars correspond to the standard deviation from 3 technical replicates. * <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 (Student’s <span class="html-italic">t</span> test).</p>
Full article ">Figure 3 Cont.
<p>Differentially regulated genes and altered pathways among the WT and ΔsfRNA replicon cell lines. (<b>A</b>) Differentially expressed genes in the ΔsfRNA1 (<b>left</b> image) and ΔsfRNA2 (<b>right</b> image) replicon cell lines plotted as volcano plots. (<b>B</b>) Top downregulated and upregulated genes of the ΔsfRNA1 and ΔsfRNA2 replicon cell lines when compared against the WT replicon cell line. (<b>C</b>) Pathway enrichment analysis results for upregulated and downregulated genes in the ΔsfRNA1 and ΔsfRNA2 replicon cell lines. (<b>D</b>) Cnetplot depicting the top 10 upregulated and downregulated Reactome pathways in ΔsfRNA1 and ΔsfRNA2 cells as networks that connect the genes with the respective pathways. (<b>E</b>) qPCR validation results for a selected set of genes. Error bars correspond to the standard deviation from 3 technical replicates. * <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 (Student’s <span class="html-italic">t</span> test).</p>
Full article ">Figure 3 Cont.
<p>Differentially regulated genes and altered pathways among the WT and ΔsfRNA replicon cell lines. (<b>A</b>) Differentially expressed genes in the ΔsfRNA1 (<b>left</b> image) and ΔsfRNA2 (<b>right</b> image) replicon cell lines plotted as volcano plots. (<b>B</b>) Top downregulated and upregulated genes of the ΔsfRNA1 and ΔsfRNA2 replicon cell lines when compared against the WT replicon cell line. (<b>C</b>) Pathway enrichment analysis results for upregulated and downregulated genes in the ΔsfRNA1 and ΔsfRNA2 replicon cell lines. (<b>D</b>) Cnetplot depicting the top 10 upregulated and downregulated Reactome pathways in ΔsfRNA1 and ΔsfRNA2 cells as networks that connect the genes with the respective pathways. (<b>E</b>) qPCR validation results for a selected set of genes. Error bars correspond to the standard deviation from 3 technical replicates. * <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 (Student’s <span class="html-italic">t</span> test).</p>
Full article ">Figure 3 Cont.
<p>Differentially regulated genes and altered pathways among the WT and ΔsfRNA replicon cell lines. (<b>A</b>) Differentially expressed genes in the ΔsfRNA1 (<b>left</b> image) and ΔsfRNA2 (<b>right</b> image) replicon cell lines plotted as volcano plots. (<b>B</b>) Top downregulated and upregulated genes of the ΔsfRNA1 and ΔsfRNA2 replicon cell lines when compared against the WT replicon cell line. (<b>C</b>) Pathway enrichment analysis results for upregulated and downregulated genes in the ΔsfRNA1 and ΔsfRNA2 replicon cell lines. (<b>D</b>) Cnetplot depicting the top 10 upregulated and downregulated Reactome pathways in ΔsfRNA1 and ΔsfRNA2 cells as networks that connect the genes with the respective pathways. (<b>E</b>) qPCR validation results for a selected set of genes. Error bars correspond to the standard deviation from 3 technical replicates. * <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 (Student’s <span class="html-italic">t</span> test).</p>
Full article ">Figure 4
<p>Differentially regulated genes and altered pathways among the replicon and control Neuro2A cell lines. (<b>A</b>) Venn diagram showing the number of common upregulated and downregulated genes among the WT, ΔsfRNA1, and ΔsfRNA2 replicon cells when compared with the control Neuro2A cell lines. (<b>B</b>) Reactome plots depicting the results from pathway enrichment analysis for the upregulated and downregulated genes in each category.</p>
Full article ">Figure 4 Cont.
<p>Differentially regulated genes and altered pathways among the replicon and control Neuro2A cell lines. (<b>A</b>) Venn diagram showing the number of common upregulated and downregulated genes among the WT, ΔsfRNA1, and ΔsfRNA2 replicon cells when compared with the control Neuro2A cell lines. (<b>B</b>) Reactome plots depicting the results from pathway enrichment analysis for the upregulated and downregulated genes in each category.</p>
Full article ">Figure 5
<p>Induction of interferon signaling in the Neuro2A WT and ΔsfRNA replicon cell lines. (<b>A</b>) Luciferase measurements from the ISRE plasmid in the Neuro2A WT and ΔsfRNA cell lines. Cells were tested under normal conditions (-IFN-containing conditioned medium) or after their treatment with conditioned medium containing IFN secreted from poly:IC-transfected cells (+IFN-containing conditioned medium). Luciferase activity is indicative of the cell’s IFN competency. Error bars correspond to the standard deviation from 3 biological replicates. *** <span class="html-italic">p</span> &lt; 0.001 (Student’s <span class="html-italic">t</span> test). (<b>B</b>) Comparison of the WNV titer produced from the Neuro2A WT and ΔsfRNA replicon cell lines at 48 hpi. *** <span class="html-italic">p</span> &lt; 0.001 (Student’s <span class="html-italic">t</span> test). (<b>C</b>) WNV titer produced in the Neuro2A ΔsfRNA2 CURED cell line in comparison with that of the plain Neuro2A cells at indicated time points.</p>
Full article ">
17 pages, 2775 KiB  
Article
Development of a Cell Culture Model for Inducible SARS-CoV-2 Replication
by Xiaoyan Wang, Yuanfei Zhu, Qiong Wu, Nan Jiang, Youhua Xie and Qiang Deng
Viruses 2024, 16(5), 708; https://doi.org/10.3390/v16050708 - 29 Apr 2024
Viewed by 1075
Abstract
Severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) induces direct cytopathic effects, complicating the establishment of low-cytotoxicity cell culture models for studying its replication. We initially developed a DNA vector-based replicon system utilizing the CMV promoter to generate a recombinant viral genome bearing reporter [...] Read more.
Severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) induces direct cytopathic effects, complicating the establishment of low-cytotoxicity cell culture models for studying its replication. We initially developed a DNA vector-based replicon system utilizing the CMV promoter to generate a recombinant viral genome bearing reporter genes. However, this system frequently resulted in drug resistance and cytotoxicity, impeding model establishment. Herein, we present a novel cell culture model with SARS-CoV-2 replication induced by Cre/LoxP-mediated DNA recombination. An engineered SARS-CoV-2 transcription unit was subcloned into a bacterial artificial chromosome (BAC) vector. To enhance biosafety, the viral spike protein gene was deleted, and the nucleocapsid gene was replaced with a reporter gene. An exogenous sequence was inserted within NSP1 as a modulatory cassette that is removable after Cre/LoxP-mediated DNA recombination and subsequent RNA splicing. Using the PiggyBac transposon strategy, the transcription unit was integrated into host cell chromatin, yielding a stable cell line capable of inducing recombinant SARS-CoV-2 RNA replication. The model exhibited sensitivity to the potential antivirals forsythoside A and verteporfin. An innovative inducible SARS-CoV-2 replicon cell model was introduced to further explore the replication and pathogenesis of the virus and facilitate screening and assessment of anti-SARS-CoV-2 therapeutics. Full article
Show Figures

Figure 1

Figure 1
<p>Design of the Rep-S-EGFP/NLuc replicon. (<b>A</b>) Design of the SARS-CoV-2 viral genome structure and replicon model. We selectively excised a portion of the S gene coding sequence while preserving its transcriptional regulatory sequence (TRS) and inserted the “self-cleaving peptide” T2A-linked EGFP or IgK secretion signal peptide-tagged NLuc sequences, allowing NLuc or EGFP to be expressed freely. The CMV promoter and hepatitis D virus anti-ribozyme sequence (HDV RZ) were introduced to the 5′ and 3′ ends of the SARS-CoV-2 genome, respectively, to construct the transcription unit. (<b>B</b>) Flowchart of the SARS-CoV-2 recombinant replicon assembly strategy. Utilizing a bacterial artificial chromosome (pBAC) as the vector, the pBAC-CMV-5′UTR-EGFP/NLuc-N-3′UTR precursor clone was generated using fusion PCR. This precursor clone was subsequently digested with restriction enzymes and ligated with ORF1ab segments (F1/F2/F3/F4) harboring homologous arms, culminating in the assembly of the replicon clone plasmid Rep-S-EGFP/NLuc using Gibson assembly.</p>
Full article ">Figure 2
<p>Rep-S-EGFP replication in transfected cells. (<b>A</b>) Changes in nucleocapsid (N) protein levels in Rep-S-EGFP-transfected cells and the effect of remdesivir on its expression. The Rep-S-EGFP replicon plasmid was transfected into VeroE6 and Huh7.5 cells. After 6 h, 10 μM remdesivir was added. Cell lysates were collected at 24, 48, 60, and 72 h post-transfection for western blotting analysis. (<b>B</b>) At 48 h post-transfection, VeroE6 cells as depicted in (<b>A</b>) were imaged under a fluorescence microscope. Representative images from three independent experiments are shown. (<b>C</b>) Schematic representation of qRT-PCR primers N-sg-F and N-sg-R targeting N gene sgRNAs. The replication of SARS-CoV-2 generates numerous subgenomic RNAs, as indicated in the diagram. (<b>D</b>) Rep-S-EGFP replicon-transfected BHK-21 cells were treated with remdesivir at indicated concentrations for 36 h. N gene sgRNA was quantified using qRT-RCR with specific primers. Each group included 3 biological replicates. * indicates <span class="html-italic">p</span> &lt; 0.05, ** indicates <span class="html-italic">p</span> &lt; 0.01, *** indicates <span class="html-italic">p</span> &lt; 0.001. (<b>E</b>) The Rep-S-EGFP replicon plasmid-transfected BHK-21 cells were treated with different concentrations of remdesivir for 36 h. N sgRNA was determined as depicted in (<b>D</b>). Cytotoxicity was evaluated using the CCK-8 assay. Cells mock-treated with DMSO were used as control. Selectivity index (value of CC<sub>50</sub>/IC<sub>50</sub>, SI). (<b>F</b>) Differential expression of progeny N sgRNAs between the replicons Rep-S-EGFP and Rep-S-EGFP-RdRp<sub>mut</sub>. Data were normalized to the Rep-S-EGFP-RdRp<sub>mut</sub> control. *** indicates <span class="html-italic">p</span> &lt; 0.001. (<b>G</b>) Comparison of N protein levels in Rep-S-EGFP- or Rep-S-EGFP-RdRp<sub>mut</sub>-transfected BHK-21 cells. Treated with remdesivir for 36 h after transfection for 6 h. Cell samples were collected and analyzed using western blotting. (<b>H</b>) Rep-S-EGFP replicon-transfected BHK-21 cells were treated with nirmatrelvir at indicated concentrations for 36 h. The inhibitory effect and cytotoxicity of nirmatrelvir on Rep-S-EGFP were investigated as depicted in (<b>E</b>). (<b>I</b>) The N protein expression in nirmatrelvir-treated cells, as described in (<b>H</b>), was determined using western blotting.</p>
Full article ">Figure 3
<p>Impact of remdesivir on reporter genes in Rep-S-NLuc and optimized Rep-N-NLuc replicons. (<b>A</b>) Inhibitory effect of remdesivir on Rep-S-NLuc NLuc activity. The Rep-S-NLuc plasmid was co-transfected with a Renilla luciferase-expressing plasmid into VeroE6, Huh7.5, and BHK-21 cells. After a 6-hour transfection, 10 μM remdesivir was added. Cell lysates were collected at 36 h post-transfection for NLuc activity analysis. Each treatment group included three biological replicates. Data were normalized to that of Renilla and analyzed for differential expression, with the DMSO group serving as the control. Results are presented as the mean ± SEM. *** indicates <span class="html-italic">p</span> &lt; 0.001, **** indicates <span class="html-italic">p</span> &lt; 0.0001. (<b>B</b>) Optimization of the reporter gene position for the Rep-N-NLuc replicon plasmid. The N gene fragment was deleted while retaining its upstream transcription regulatory sequence (TRS). The non-secreted NLuc gene was inserted into the deleted location to rescue viral replication using N gene complementation. (<b>C</b>) Remdesivir significantly inhibited the NLuc activity (relative luminescence units, RLU) of the Rep-N-NLuc replicon. The Rep-N-NLuc replicon plasmid, pcDNA3.1-N, and Renilla luciferase-expressing plasmid were co-transfected into BHK-21 cells. The deep blue bar indicates the DMSO-treated MOCK group, while the orange bar represents the group treated with 10 μM remdesivir.Each treatment group included three biological replicates. *** indicates <span class="html-italic">p</span> &lt; 0.001. (<b>D</b>) Remdesivir significantly inhibited the generation of N-NLuc sgRNAs in Rep-N-NLuc. qRT-PCR detection was performed using the samples from (<b>C</b>). N-NLuc sgRNA was detected using special primers. ** indicates <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 4
<p>An inducible SARS-CoV-2 replicon system co-transfected with N-protein rescued viral replication upon Cre/<span class="html-italic">loxP</span>-mediated site-specific recombination. (<b>A</b>) Structure of the recombinant SARS-CoV-2 replicon (iRep-N-NLuc). A transcription stop cassette integrated between nt 66 and 67 (i.e., CAG66^G67; ^ indicates the potential exon/exon boundary) of the NSP1 gene. The inserted sequence consists of the indicated elements in order. PiggyBac 5′ TR and chicken β-globin insulators are inserted at the 5′ end of the CMV promoter and at the 3′ end of the HDV RZ region, respectively, flanking the transcription unit. “L” stands for NLuc. (<b>B</b>) A schema of the intracellular process of inducible iRep-N-NLuc replication in the cell model. In the absence of Cre recombinase, the CMV promoter only drives the expression of the blasticidin resistance gene, preventing SARS-CoV-2 replication, facilitating the selection of stably integrated cell lines. Following Cre recombinase induction in replicon cell lines, site-specific recombination removes the transcription stop cassette at the DNA level. The CMV promoter then initiates transcription of the viral genome. RNA splicing mechanisms within the nucleus eliminate the remaining single <span class="html-italic">LoxP</span> site. The cell nucleus is depicted in blue, while the cytoplasm is shown in light blue. (<b>C</b>) BHK-21 cells were co-transfected with the iRep-N-NLuc replicon and indicated plasmids plus a Renilla luciferase normalization plasmid. The fold change in NLuc activity was normalized to that of the Renilla luciferase activity. *** indicates <span class="html-italic">p</span> &lt; 0.001. (<b>D</b>) BHK-21 cells were co-transfected with iRep-N-NLuc replicon, pCDNA3.1-N, and pCDH-Cre. At 6 h after transfection, cells were treated with 10 μM remdesivir for an additional 24 h and subjected to analysis of NLuc activity. Cells mock-treated with DMSO were used as a control. ** indicates <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 5
<p>Generation of an inducible stable cell line producing a recombinant SARS-CoV-2 replicon. (<b>A</b>) Schematic illustration of the procedure to generate a cell-based SARS-CoV-2 replicon model. (<b>B</b>,<b>C</b>) The iRep-N-NLuc cell line was transfected with a bicistronic IRES vector (pCDH-N-IRES-Cre). Cells were either treated with 10 μM remdesivir or mock treated with DMSO. The recombinant N-NLuc sgRNAs (<b>B</b>) and NLuc activity (<b>C</b>) were determined at 24 h post-transfection. Error bars indicate the mean ± SEM from three independent experiments. ** indicates <span class="html-italic">p</span> &lt; 0.01, **** indicates <span class="html-italic">p</span> &lt; 0.0001. (<b>D</b>) pCDH-N-IRES-Cre-transfected iRep-N-NLuc cells were treated with different concentrations of remdesivir for 24 h. NLuc activity was determined. Cytotoxicity was evaluated using the CCK-8 assay. Data were expressed as the percentage of the mock-treated controls. Error bars indicate the mean ± SEM from three independent experiments.</p>
Full article ">Figure 6
<p>Antiviral drug screening and validation based on the inducible SARS-CoV-2 replicon cell model. (<b>A</b>) Schematic illustration of the procedure for antiviral screening based on the iRep-N-NLuc cells model. (<b>B</b>) Evaluation of the antiviral activity of candidate compounds in the iRep-N-NLuc-cells model. Cells were transfected with pCDH-N-IRES-Cre and treated with candidate compounds at indicated doses. The NLuc activity was determined at 24 h post-transfection and were expressed as a percentage of the values measured in DMSO-treated cells. Error bars indicate the mean ± SEM from three independent experiments. ** indicates <span class="html-italic">p</span> &lt; 0.01, *** indicates <span class="html-italic">p</span> &lt; 0.001, **** indicates <span class="html-italic">p</span> &lt; 0.0001. (<b>C</b>) iRep-N-NLuc cells were transfected with pCDH-N-IRES-Cre, followed by the treatment with forsythiaside A at indicated doses for 24 h. NLuc activity was determined as depicted in (<b>B</b>). Cell viability was measured using the CCK-8 assay. (<b>D</b>) BHK-21 cells were transfected with Rep-S-EGFP and treated with forsythiaside A at indicated doses. After 48 h, N protein expression was determined using western blotting. (<b>E</b>) Rep-N-NLuc-transfected cells were treated with varying concentrations of verteporfin for 24 h. NLuc activity was measured and expressed as a percentage of the values measured in mock-treated cells. Cell cytotoxicity and CC<sub>50</sub> were assessed using the CCK-8 assay. (<b>F</b>) Rep-S-GFP-transfected cells were treated with verteporfin for 36 h. N sgRNA was determined using qRT-PCR. Data are presented as the mean ± SEM from three independent experiments. ** indicates <span class="html-italic">p</span> &lt; 0.01. (<b>G</b>) iRep-N-NLuc cells were transfected with pCDH-N-IRES-Cre for 6 h and subsequently treated with varying concentrations of nirmatrelvir. NLuc activity was determined as depicted in (<b>B</b>). The cytotoxicity assay is the same as the CCK8 assay in <a href="#viruses-16-00708-f002" class="html-fig">Figure 2</a>G.</p>
Full article ">
24 pages, 2158 KiB  
Review
A Decade of Discovery—Eukaryotic Replisome Disassembly at Replication Termination
by Rebecca M. Jones, Alicja Reynolds-Winczura and Agnieszka Gambus
Biology 2024, 13(4), 233; https://doi.org/10.3390/biology13040233 - 31 Mar 2024
Cited by 1 | Viewed by 1790
Abstract
The eukaryotic replicative helicase (CMG complex) is assembled during DNA replication initiation in a highly regulated manner, which is described in depth by other manuscripts in this Issue. During DNA replication, the replicative helicase moves through the chromatin, unwinding DNA and facilitating nascent [...] Read more.
The eukaryotic replicative helicase (CMG complex) is assembled during DNA replication initiation in a highly regulated manner, which is described in depth by other manuscripts in this Issue. During DNA replication, the replicative helicase moves through the chromatin, unwinding DNA and facilitating nascent DNA synthesis by polymerases. Once the duplication of a replicon is complete, the CMG helicase and the remaining components of the replisome need to be removed from the chromatin. Research carried out over the last ten years has produced a breakthrough in our understanding, revealing that replication termination, and more specifically replisome disassembly, is indeed a highly regulated process. This review brings together our current understanding of these processes and highlights elements of the mechanism that are conserved or have undergone divergence throughout evolution. Finally, we discuss events beyond the classic termination of DNA replication in S-phase and go over the known mechanisms of replicative helicase removal from chromatin in these particular situations. Full article
(This article belongs to the Special Issue The Replication Licensing System)
Show Figures

Figure 1

Figure 1
<p>Steps of DNA replication. (<b>i</b>) Origins are selected along the chromosome and double MCM2-7 hexamers are loaded onto the DNA. C-tier: C-terminal tier of MCM2-7; N: N-terminal tier of MCM2-7. (<b>ii</b>) Following activation by binding of CDC45 and GINS (complex of SLD5, PSF1-3), CMG helicases, each built around a single MCM2-7 hexamer, bypass one-another and replication forks are created. Replication machinery at active replication forks is formed by hundreds of proteins. We depicted a few additional essential ones (e.g., DNA polymerases) in a side box. For the clarity of the figures, we retained only components of CMG helicase in further steps of the mechanism. (<b>iii</b>) CMG helicases progress through the chromatin, unwinding DNA in a 3′-5′ direction during elongation. (<b>iv</b>) Replication forks from neighbouring origins converge towards one-another as the DNA between them is unwound, with the aid of topoisomerases, fork rotation and 5′-3′ helicase enzymes. (<b>v</b>) The terminating CMG meets a ssDNA-dsDNA junction, comprising the RNA primer (red arrow) of the last, immature Okazaki fragment. (<b>vi</b>) The CMGs slide onto the dsDNA and translocate away, prior to disassembly, while the last stretches of DNA are synthesised and the final fragments are ligated together.</p>
Full article ">Figure 2
<p>Initiation and termination zones. In prokaryotic circular mini-chromosomes, replication initiates from just one origin. The helicase (DnaB in bacteria and T-Ag in SV40 virus) first bypasses five ‘permissive’ (grey section head-on) Tus–Ter complexes, before reaching the termination site. From here, helicase progression is halted as it meets ‘non-permissive’ Tus–Ter complexes (black section head-on). In eukaryotic chromosomes, origins initiate at thousands of sights throughout the genome. Termination zones vary greatly as origins fire stochastically and progressing replication forks encounter many problems and obstacles along the way. C-tier: C-terminal tier of MCM2-7; N: N-terminal tier of MCM2-7; DPC: DNA-protein crosslink.</p>
Full article ">Figure 3
<p>Mechanism of Cullin-driven replisome disassembly. (<b>i</b>) MCM7 is specifically polyubiquitylated with K48-linked polyubiquitin chains by Cullin ubiquitin ligase. In budding yeast, this ubiquitylation process engages SCF<sup>Dia2</sup> E3 ligase, which cooperates with Cdc34 E2 conjugation enzyme, while recruitment relies upon Ctf4 and Mrc1. In metazoa, CUL2<sup>LRR1</sup> E3 ligase, together with a range of priming and extending E2 conjugating enzymes, acts on the MCM7 substrate. In this case, TIMELESS and TIPIN replisome factors help in the recruitment of CUL2<sup>LRR1</sup> to CMG and promote ubiquitylation events. (<b>ii</b>) Ubiquitylated CMGs are recognized by p97 unfoldase, which binds to its substrate by cooperating with a range of cofactors. (<b>iii</b>) The NPL4 subunit of p97 starts unfolding the ubiquitin chain on MCM7, which is then transferred through the channel of p97. The last step of this process involves extraction of MCM7 from the CMG helicase and removal/dismantling of the whole CMG from DNA.</p>
Full article ">Figure 4
<p>Mechanism of Cullin ubiquitin ligase interaction with terminated replisome. Schematic model representations depicting interactions between Cullin ubiquitin ligases and the budding yeast and human replisomes adapted from Jenkyn-Bedford et al.’s (2021) cryo-EM structures of complexes (reconstituted from purified proteins and assembled on a short fragment of dsDNA) [<a href="#B74-biology-13-00233" class="html-bibr">74</a>]. Both Cdc53 and CUL2 display elongated configurations and conformational flexibility. (<b>i</b>) Budding yeast replisome interacting with Cdc53<sup>Dia2</sup> (SCF<sup>Dia2</sup>); Dia2 interacts with Mcm3 and Mcm5 ZnF domains, with the Mcm7 N-terminus sandwiched between them. Due to the elongated forms and conformational flexibility of Cullin ligase, the Hrt1 (Rbx) component of the E3 ligase can be positioned at a close proximity to K29 of Mcm7, as indicated. (<b>ii</b>) Human replisome interacting with CUL2<sup>LRR1</sup>; the N-terminal pleckstrin homology domain of LRR1 interacts with the ZnF domains of MCM2 and MCM6, with dsDNA and with TIMELESS while the LRR region of LRR1 interacts with ZnF domains of MCM3 and MCM5 and the HMG-box of AND-1.</p>
Full article ">Figure 5
<p>Mechanism of TRAIP-driven replisome disassembly. Replisomes retained on chromatin until mitosis in metazoa undergo TRAIP-dependent disassembly; these include post-termination replisomes, stalled replisomes or replisomes trapped in regions of under-replicated DNA. (<b>i</b>) Replisomes are ubiquitylated with K6- and K63-linked ubiquitin chains on the MCM7 subunit of CMG. This modification is dependent on the presence of TRAIP E3 ligase, although the E2 is currently unknown. (<b>ii</b>) Ubiquitylated CMGs become substrates for p97 unfoldase, which cooperates with a range of cofactors that aid in recognition and binding to ubiquitylated CMGs. (<b>iii</b>) Dismantling of the replisomes from chromatin starts when the NPL4 subunit of p97 unfolds the ubiquitin chain attached to the MCM7 subunit of CMG and then passes it through the p97 channel. The process ends with MCM7 being extracted from the replisome and disintegration/removal of the whole CMG from chromatin.</p>
Full article ">
19 pages, 2429 KiB  
Review
Starting DNA Synthesis: Initiation Processes during the Replication of Chromosomal DNA in Humans
by Heinz Peter Nasheuer and Anna Marie Meaney
Genes 2024, 15(3), 360; https://doi.org/10.3390/genes15030360 - 14 Mar 2024
Cited by 1 | Viewed by 2930
Abstract
The initiation reactions of DNA synthesis are central processes during human chromosomal DNA replication. They are separated into two main processes: the initiation events at replication origins, the start of the leading strand synthesis for each replicon, and the numerous initiation events taking [...] Read more.
The initiation reactions of DNA synthesis are central processes during human chromosomal DNA replication. They are separated into two main processes: the initiation events at replication origins, the start of the leading strand synthesis for each replicon, and the numerous initiation events taking place during lagging strand DNA synthesis. In addition, a third mechanism is the re-initiation of DNA synthesis after replication fork stalling, which takes place when DNA lesions hinder the progression of DNA synthesis. The initiation of leading strand synthesis at replication origins is regulated at multiple levels, from the origin recognition to the assembly and activation of replicative helicase, the Cdc45–MCM2-7–GINS (CMG) complex. In addition, the multiple interactions of the CMG complex with the eukaryotic replicative DNA polymerases, DNA polymerase α-primase, DNA polymerase δ and ε, at replication forks play pivotal roles in the mechanism of the initiation reactions of leading and lagging strand DNA synthesis. These interactions are also important for the initiation of signalling at unperturbed and stalled replication forks, “replication stress” events, via ATR (ATM–Rad 3-related protein kinase). These processes are essential for the accurate transfer of the cells’ genetic information to their daughters. Thus, failures and dysfunctions in these processes give rise to genome instability causing genetic diseases, including cancer. In their influential review “Hallmarks of Cancer: New Dimensions”, Hanahan and Weinberg (2022) therefore call genome instability a fundamental function in the development process of cancer cells. In recent years, the understanding of the initiation processes and mechanisms of human DNA replication has made substantial progress at all levels, which will be discussed in the review. Full article
(This article belongs to the Special Issue Mechanisms and Regulation of Human DNA Replication)
Show Figures

Figure 1

Figure 1
<p>The initiation process at eukaryotic origins of DNA replication. At the end of mitosis and in the early G1 phase of the cell cycle, ORC, the origin-recognition complex, binds to eukaryotic origins of replication together with CDC6. In G1, CDT1 (chromatin licensing and DNA replication factor 1) together with the MCM2-7 complex then associates with the CDC6–ORC complex, and the MCM2-7 proteins are loaded as helicase-inactive double hexamers (MCM-DHs) onto the chromatin, forming the pre-replicative complex (Pre-RC) and license the origin. The modification of CDC6 and binding of CDT1 to Geminin inactivates the loading activity of these proteins with CDC6 being degraded similarly as free CDT1. In the next step, Treslin-MTBP (SLD3-SLD7 in yeast) interacts with MCM-DHs at the chromatin, and the DBF4/DRF1 CDC7 kinase (DDK) phosphorylates the MCM2-7 proteins. The DDK-dependent phosphorylation can be reversed by RIF1-PP1 making this step reversible. Next, Cyclin-CDKs phosphorylate Treslin and stimulate the formation of Donson–TOPBP1 complexes, which in turn bind to MCM-DHs. Donson–TOPBP1 supports the loading of the GINS complex and its association with MCM-DHs. The binding of CDC45 leads to the formation of the CMG complex and its activation, whereas TOPBP1 and Treslin–MTBP are released from the chromatin. CryoEM data suggest that Donson associates as a dimer with CMG, but only one Donson subunit binds to GINS and MCM2-7 proteins stabilising the CMG complex [<a href="#B28-genes-15-00360" class="html-bibr">28</a>]. In the following step, two replication forks (RFs) are formed and replication protein A (RPA), with the help of the CDC45, binds to and stabilises the resulting ssDNA. The association of AND-1/CTF4/WDHD1 (shown as AND-1 in the diagram) with CMG allows for the loading of an inactive DNA polymerase α (Pol α) (dark green), including its primase subunits, to RFs. The activation of Pol α (light green) permits the primase subunit PRIM1/PRI1, with the help of PRIM2/PRI2 and additional replication factors, to synthesise the first RNA primer in origin sequences, resulting in the completion of the initiation process at origins and the start of the elongation phase. Additional proteins associated with RFs, such as the fork-stabilising proteins Timeless, Tipin, and Claspin plus Pol ε [<a href="#B3-genes-15-00360" class="html-bibr">3</a>,<a href="#B42-genes-15-00360" class="html-bibr">42</a>,<a href="#B43-genes-15-00360" class="html-bibr">43</a>], were omitted in the diagram for simplification and clarity reasons providing a better overview. Adapted using information from [<a href="#B3-genes-15-00360" class="html-bibr">3</a>,<a href="#B26-genes-15-00360" class="html-bibr">26</a>,<a href="#B27-genes-15-00360" class="html-bibr">27</a>,<a href="#B28-genes-15-00360" class="html-bibr">28</a>,<a href="#B44-genes-15-00360" class="html-bibr">44</a>,<a href="#B45-genes-15-00360" class="html-bibr">45</a>,<a href="#B46-genes-15-00360" class="html-bibr">46</a>] and created with <a href="http://BioRender.com" target="_blank">BioRender.com</a>.</p>
Full article ">Figure 2
<p>Leading and lagging strand synthesis at a eukaryotic replication fork. In this RF model, CMG helicase (CDC45-MCM2-7-GINS with CDC45, in dark blue, the MCM2-7 hexamer, in purple, and GINS, in light green) unwinds the parental dsDNA into the leading and lagging strand templates (dark-red and dark blue, respectively). The protein Donson associates with the CMG complex during its formation and remains attached to it during the unwinding reaction. Additionally, the diagram shows the replication proteins that are involved in DNA synthesis and the maturation of Okazaki fragments. As seen in the model, RPA heterotrimers (three shades of blue) bind to the unwound ssDNAs preventing hairpin formation and nuclease-dependent ssDNA degradation. RFC (blue) loads the PCNA ring (red brown) onto the primed template DNA. The latter stabilises Pol ε (light blue) on the template DNA when synthesising the leading strand (light blue DNA). Pol ε also associates with the CMG complex to support its unwinding activity, but this interaction might also be important during replication fork stalling (see <a href="#sec5-genes-15-00360" class="html-sec">Section 5</a>). For lagging strand DNA synthesis, the AND-1/CTF4/WDHD1 homotrimer (named AND-1 in the diagram with one subunit consisting of an HMG (green), SepB (dark blue), and WD (blue) domain) links CMG to the Pol α complex (green). The primase function of Pol α synthesises the RNA primer (light grey), starting Okazaki fragment synthesis during lagging strand synthesis. After the initiation step, primase hands over the RNA primer to the DNA polymerase domain of Pol α on PolA1 (first polymerase transition). The latter extends the RNA primer and synthesises a short RNA–DNA fragment before leaving the template. RFC (blue) replaces Pol α with the help of RPA and loads PCNA, the DNA clamp, on the primed DNA. This RFC–PCNA complex allows Pol δ (pink) to associate with the RNA–DNA primer (2nd polymerase transition). The RFC–PCNA–Pol δ complex elongates this RNA–DNA in a processive manner until it reaches the next Okazaki fragment. Then, Pol δ slows down but continues to elongate the newly synthesised DNA. The polymerase displaces the RNA and parts of the Pol α-synthesised DNA of the Okazaki fragment in front (strand displacement). Thus, Pol δ produces an RNA-DNA flap, which is recognised and cleaved by PCNA-associated FEN1 (blue), creating a perfect product, nicked DNA, for LIG1 (top panel on the left; the two inserted panels provide insights into the different pathways of the Okazaki fragment maturation process). The DNA ligase LIG1, which is also bound to PCNA along with Pol δ and FEN1, then ligates the two DNA fragments, yielding a continuous stretch of DNA. In an alternate pathway, RPA binds the flap structure produced by Pol δ competing with FEN1 (lower inserted panel). RPA recruits the DNA2 helicase/endonuclease to the flap structure. The latter in turn cleaves the ssDNA but leaving an extra nucleotide remaining, which results in a product that LIG1 does not ligate. After RPA and DNA2 have left the DNA, FEN1 cuts off the remaining base and LIG1 ligates the two DNA fragments. Adapted from [<a href="#B2-genes-15-00360" class="html-bibr">2</a>,<a href="#B32-genes-15-00360" class="html-bibr">32</a>,<a href="#B61-genes-15-00360" class="html-bibr">61</a>,<a href="#B67-genes-15-00360" class="html-bibr">67</a>] and created with <a href="http://BioRender.com" target="_blank">BioRender.com</a>.</p>
Full article ">Figure 3
<p>Signalling at unperturbed and perturbed replication forks. In these simplified RF models, ATR/MEC1 and CHK1/RAD53 initiate signalling pathways at RFs prior to and after passing damaged DNA resulting in stalled leading strand synthesis in the case of DNA damage, which is summarised in panels (<b>A</b>) and (<b>B</b>), respectively. In panel A, during normal, unperturbed replication processes at RFs, the Okazaki fragment synthesis on the lagging strand produces a signal via ATR/MEC1 (light red) and CHK1/RAD53 (dark pink) to slow down the RF by modulating, e.g., the CMG helicase, to synchronise DNA synthesis and nucleotide synthesis. This ATR/MEC1 activity requires Pol α and DNA2 (Okazaki fragment initiation and maturation) and signals unperturbed RFs [<a href="#B114-genes-15-00360" class="html-bibr">114</a>,<a href="#B116-genes-15-00360" class="html-bibr">116</a>]. When Pol ε on the leading strand encounters a DNA lesion (pink star) or is exposed to nucleotide depletion, the enzyme stops DNA synthesis and disengages with or modulates the CMG complex. Then, the latter continues to unwind DNA yielding stretches of ssDNA, which are boundby RPA. These RPA–ssDNA structures recruit ATRIP and ATR (dark red) to the leading strand, in addition to ATR’s binding via TOPBP1, its main activator, to the lagging strand with Pol α and 9-1-1 as partners (panel B). Additionally, the binding of Pol ε to TOPBP1, which may occur via a binding site of Pol ε to TOPBP1 hidden when associated with the CMG complex, may enhance this DDR signalling. It is important to note that Pol α and the RNA primer synthesis are key for the initiation of replication stress signals via ATR/MEC1 and that at RFs, Pol α does not synthesise RNA primers on the leading strand [<a href="#B34-genes-15-00360" class="html-bibr">34</a>,<a href="#B72-genes-15-00360" class="html-bibr">72</a>,<a href="#B73-genes-15-00360" class="html-bibr">73</a>,<a href="#B117-genes-15-00360" class="html-bibr">117</a>,<a href="#B118-genes-15-00360" class="html-bibr">118</a>]. The brown, pink, and red arrows indicate ATR/MEC1 and CHK1/RAD53 signalling during unperturbed and perturbed DNA replication. This model suggests that ATR signalling requires multiple key partners located on both template strands of a stalled RF. It is important to mention that human and yeast CHK1 are homologues on the sequence level, whereas human CHK2 is the orthologue of yeast RAD53, but regarding the ATR pathway during replication stress, human CHK1 and RAD53 are functionally equivalent [<a href="#B119-genes-15-00360" class="html-bibr">119</a>]. The figure was created with Bio Render using published results [<a href="#B32-genes-15-00360" class="html-bibr">32</a>,<a href="#B72-genes-15-00360" class="html-bibr">72</a>,<a href="#B73-genes-15-00360" class="html-bibr">73</a>,<a href="#B114-genes-15-00360" class="html-bibr">114</a>,<a href="#B116-genes-15-00360" class="html-bibr">116</a>,<a href="#B117-genes-15-00360" class="html-bibr">117</a>,<a href="#B119-genes-15-00360" class="html-bibr">119</a>,<a href="#B120-genes-15-00360" class="html-bibr">120</a>,<a href="#B121-genes-15-00360" class="html-bibr">121</a>,<a href="#B122-genes-15-00360" class="html-bibr">122</a>,<a href="#B123-genes-15-00360" class="html-bibr">123</a>,<a href="#B124-genes-15-00360" class="html-bibr">124</a>].</p>
Full article ">
26 pages, 9154 KiB  
Article
Developmental Changes in Genome Replication Progression in Pluripotent versus Differentiated Human Cells
by Sunil Kumar Pradhan, Teresa Lozoya, Paulina Prorok, Yue Yuan, Anne Lehmkuhl, Peng Zhang and M. Cristina Cardoso
Genes 2024, 15(3), 305; https://doi.org/10.3390/genes15030305 - 27 Feb 2024
Cited by 1 | Viewed by 2107
Abstract
DNA replication is a fundamental process ensuring the maintenance of the genome each time cells divide. This is particularly relevant early in development when cells divide profusely, later giving rise to entire organs. Here, we analyze and compare the genome replication progression in [...] Read more.
DNA replication is a fundamental process ensuring the maintenance of the genome each time cells divide. This is particularly relevant early in development when cells divide profusely, later giving rise to entire organs. Here, we analyze and compare the genome replication progression in human embryonic stem cells, induced pluripotent stem cells, and differentiated cells. Using single-cell microscopic approaches, we map the spatio-temporal genome replication as a function of chromatin marks/compaction level. Furthermore, we mapped the replication timing of subchromosomal tandem repeat regions and interspersed repeat sequence elements. Albeit the majority of these genomic repeats did not change their replication timing from pluripotent to differentiated cells, we found developmental changes in the replication timing of rDNA repeats. Comparing single-cell super-resolution microscopic data with data from genome-wide sequencing approaches showed comparable numbers of replicons and large overlap in origins numbers and genomic location among developmental states with a generally higher origin variability in pluripotent cells. Using ratiometric analysis of incorporated nucleotides normalized per replisome in single cells, we uncovered differences in fork speed throughout the S phase in pluripotent cells but not in somatic cells. Altogether, our data define similarities and differences on the replication program and characteristics in human cells at different developmental states. Full article
(This article belongs to the Special Issue Mechanisms and Regulation of Human DNA Replication)
Show Figures

Figure 1

Figure 1
<p>Cell cycle and replication dynamics analysis of pluripotent and somatic cells. (<b>A</b>) Schematic representation of how the fraction of cells in S phase was determined based on DAPI and EdU intensity using high-throughput microscopy. (<b>B</b>) Doubling time of hESC H1, hiPSC A4, and hTERT RPE1 was calculated by counting cell numbers at different time points from a defined number of seeded cells. The S phase fraction was calculated by dividing the EdU positive cells with the total number of cells from high-throughput image analysis. The S phase duration was calculated by multiplying the doubling time with the fraction of cells in the S phase. (* = multiplication) (<b>C</b>) A pulse-chase–pulse-chase experiment followed by replication foci (RFi) detection at three time points in the same cell in different cell lines. (<b>D</b>) Illustration shows cell cycle phases and (sub)S phase durations among cell types. The duration of each S phase sub-stage was calculated by multiplying the fraction of cells in each sub-stage by the doubling time of the specific cell. (<b>E</b>) Live-cell time-lapse microscopy of hiPSC A4 and hTERT RPE1 expressing GFP-PCNA showing genome replication progression. PCNA S phase foci are visible from 90 min on. For more details see <a href="#app1-genes-15-00305" class="html-app">Table S1</a>. Scale bar: 10 µm.</p>
Full article ">Figure 2
<p>Feature analysis of the replication foci (RFi) in different S phases. (<b>A</b>) The illustration shows the image analysis approach for characterizing RFi from different time points in the same cell. Nuclear mask was created using the DNA dye DAPI and applied to the other three channels (EdU, BrdU, and PCNA) before the respective channel segmentation. Within this mask, RFi features were quantified. Scale bar: 5 µm. (<b>B</b>) Plots show the number, volume, and distance analysis (inter-RFi and RFi to the nuclear border) of the RFi as the cell progresses through the S phase. (<b>C</b>) Illustration shows the subnuclear distribution of RFi and its features in different S phase stages as indicated. The lower and upper whiskers of the boxplot correspond to the 25th and 75th percentiles, the box to the 50th percentile, and the line depicts the median. Statistical significance was performed using ANOVA, and Tukey’s honest significance test (not significant is given for <span class="html-italic">p</span>-values ≥ 0.05; one star (*) for <span class="html-italic">p</span>-values &lt; 0.05 and ≥ 0.005; two stars (**) is given for values &lt; 0.005 and ≥ 0.0005, and ≥ 0.0005 are given (***); only the significant differences are shown). For more details, see <a href="#app1-genes-15-00305" class="html-app">Table S2</a>. Scale bar: 10 µm.</p>
Full article ">Figure 3
<p>Quantification of the number of replicons and fork speed in S phase stages (<b>A</b>) Super-resolution PCNA (green) images overlaid with DAPI (blue) in three S phase stages in hESC H1, hiPSC A4, and hTERT RPE1 are shown. (<b>B</b>) Comparison between confocal and super-resolution (ASJD) images of RFi of the same cell and region. The plot shows the comparative volume of (nano)RFi detected in confocal and ASJD mode. (<b>C</b>) The plot shows the quantification of the nano-RFi in different S phase stages. (<b>D</b>) An illustration depicts the approach to measure the comparative replication fork speed. The EdU was pulsed for 15 min and detected using click chemistry, and PCNA was detected by antibodies. For measuring nucleotide incorporation rate, the ratio of EdU (incorporated nucleotides) and PCNA (active replication) sum intensities was measured as a marker for the speed of replication forks. If the ratio shows a value ≤ 1, this means a complete overlap or localization of EdU inside PCNA and indicates a slow replication fork speed. If the ratio of both signals is &gt; 1, more DNA was synthesized, indicating faster replication fork speed. The middle plot depicts the fork rates of S phase cells across cell lines measured by high-throughput imaging and analysis without discriminating between S phase stages. The right plot shows the fork rate of individual S phase stages measured from high-resolution images across cell lines. The lower and upper whiskers of the boxplot correspond to the 25th and 75th percentiles, the box to the 50th percentile, and the line depicts the median. Statistical significance was performed using ANOVA and Tukey’s honest significance test (not significant is given for <span class="html-italic">p</span>-values ≥ 0.05; one star (*) for <span class="html-italic">p</span>-values &lt; 0.05 and ≥ 0.005; two stars (**) is given for values &lt; 0.005 and ≥ 0.0005, and 0.0005 to 0 are given (***); only the significant differences are shown). For more details, see <a href="#app1-genes-15-00305" class="html-app">Table S3</a>. Scale bar: 10 µm.</p>
Full article ">Figure 4
<p>Genome-wide replication origins distribution in selected human cell lines based on the SNS-seq origin mapping method. (<b>A</b>) Representative example of replication origins distribution in hESC, hiPSC, and HMEC. The origin profiles correspond to normalized read counts (scale 0–0.7 counts per million). Below the profiles, the origins identified by MACS2 peak callers are shown. (<b>B</b>) Comparison of the origin numbers in the human embryonic cell line (hESC), induced pluripotent cells (hiPSC), and the somatic HMEC cell line. Additionally, for the total identified peaks, the graph represents the origin number after clustering of closely situated origins at the distances of 10, 20, and 30 kb. (<b>C</b>) Comparison of the inter-origin distances (IOD). The IOD distances were also compared after origin clustering at the distances specified. The statistical evaluation of IODs between different cell lines and same clustering distance were significant (<span class="html-italic">p</span>-value &lt; 0.001) with the only exception of the difference between ESC and iPSC, which was not significant <span class="html-italic">p</span>-value = 0.4125770. For more details see <a href="#app1-genes-15-00305" class="html-app">Table S4</a>. (<b>D</b>) Overlap of peaks among the different cell lines.</p>
Full article ">Figure 5
<p>Quantification of chromatin compaction with replication progression across cell lines. RFi in S phase stages were mapped to chromatin compaction classes across cell lines using the statistical tool Nucim on platform R (see <a href="#sec2-genes-15-00305" class="html-sec">Section 2</a> and <a href="#app1-genes-15-00305" class="html-app">Figure S2</a>). Lines connecting data corresponding to the same S phase stages were drawn for easier visualization. Distribution differences on classes <span class="html-italic">p</span>-values &lt; 0.005 for all S phase stages for each cell line).</p>
Full article ">Figure 6
<p>Replication timing of genomic repeat elements. (<b>A</b>) Schematic representation of a chromosome with the tandem and interspersed repeat sequences color-coded. (<b>B</b>) The co-detection of two combinations of probes across cell lines as indicated (red and cyan) with PCNA (green). The line plots depict the fluorescence intensity distribution of the PCNA and the probes along the line (in microns) drawn on the merged image. (<b>C</b>) Heat plot shows the fold change in the sum intensity of each probe replicated in the S phase stages as indicated. The sum intensity of each probe was measured using the segmented RFi as masks in each S phase stage in individual cells and normalized to the median sum intensity of S I for each probe and cell line. For details, see <a href="#app1-genes-15-00305" class="html-app">Figure S1B and Table S5</a>. Scale bar: 10 µm.</p>
Full article ">Figure 7
<p>Developmental difference in replication timing of rDNA repeats. (<b>A</b>) Analysis pipeline to characterize the rDNA replication timing. RFi and rDNA spots were segmented separately. In FiJi, using the logical function “AND”, both segmented spots were processed to obtain the intersected voxels from both RFi and rDNA, which directly represent the colocalizing rDNA and RFi. (<b>B</b>) Images show the PCNA (green), the rDNA (red), and the merged images in different S phase stages as indicated. The contours (yellow) in the enlarged merged image indicate the colocalizing spots. (<b>C</b>) The plot depicts the quantification of the replicating rDNA spots in the S phase stages in hPSCs and hTERT RPE1. (<b>D</b>) Images show the overlap of RNA polymerase I subunit RPA 194 (representing active transcription) with replicating rDNA repeats in the S phase stages indicated. Contours in the enlarged image show the colocalizing RPA 194 and replicating rDNA as measured with the “AND” logic operation. The line plot shows the intensity distribution of RPA 194 (cyan), rDNA (red), and the EdU (green) along the line. (<b>E</b>) The plot shows the number of replicating rDNA spots associated with RPA 194. The lower and upper whiskers of the boxplot correspond to the 25th and 75th percentiles, the box to the 50th percentile, and the line depicts the median. Statistical significance was performed using ANOVA, and Tukey’s honest significance test (not significant is given for <span class="html-italic">p</span>-values ≥ 0.05; one star (*) for <span class="html-italic">p</span>-values &lt; 0.05 and ≥ 0.005; two stars (**) is given for values &lt; 0.005 and ≥ 0.0005, and 0.0005 to 0 are given (***); only the significant differences are shown). For more details, see <a href="#app1-genes-15-00305" class="html-app">Table S6</a>. The scale bar is 10 µm and 2 µm in the enlarged images.</p>
Full article ">Figure 8
<p>A summary of the developmental difference in genome replication features in pluripotent stem cells (PSC) and somatic cells. The late-replicating RFi cluster around and inside nucleolus in PSC but moves away in somatic cells. The late-replicating rDNA in PSC replicates earlier in S II. The increased origin firing increases the possibility of replication and transcription collision in S I of PSCs. (RFi: replication foci, RPA 194: large subunit of RNA pol I marking active transcription.).</p>
Full article ">
9 pages, 1987 KiB  
Communication
Somatic Embryogenesis and Agrobacterium-Mediated Gene Transfer Procedures in Chilean Temperate Japonica Rice Varieties for Precision Breeding
by Marion Barrera, Blanca Olmedo, Carolina Zúñiga, Mario Cepeda, Felipe Olivares, Ricardo Vergara, Karla Cordero-Lara and Humberto Prieto
Plants 2024, 13(3), 416; https://doi.org/10.3390/plants13030416 - 31 Jan 2024
Viewed by 1513
Abstract
Rice (Oryza sativa) varieties are generated through breeding programs focused on local requirements. In Chile, the southernmost rice producer, rice productivity relies on the use and generation of temperate japonica germplasms, which need to be adapted to the intensifying effects of [...] Read more.
Rice (Oryza sativa) varieties are generated through breeding programs focused on local requirements. In Chile, the southernmost rice producer, rice productivity relies on the use and generation of temperate japonica germplasms, which need to be adapted to the intensifying effects of climate change. Advanced biotechnological tools can contribute to these breeding programs; new technologies associated with precision breeding, including gene editing, rely on procedures such as regeneration and gene transfer. In this study, the local rice varieties Platino, Cuarzo, Esmeralda, and Zafiro were evaluated for somatic embryogenesis potential using a process that involved the combined use of auxins and cytokinins. An auxin-based (2,4-D) general medium (2N6) allowed for the induction of embryogenic masses in all the genotypes. After induction, masses required culturing either in N6R (kinetin; Platino) or N6RN (BAP, kinetin, IBA, and 2,4-D; Cuarzo, Esmeralda, and Zafiro) to yield whole plants using regeneration medium (N6F, no hormone). The sprouting rates indicated Platino as the most responsive genotype; for this reason, this variety was evaluated for gene transfer. Fifteen-day-old embryo masses were assayed for Agrobacterium-mediated transformation using the bacterial strain EHA105 harboring pFLC-Myb/HPT/GFP, a modified T-DNA vector harboring a geminivirus-derived replicon. The vector included the green fluorescent protein reporter gene, allowing for continuous traceability. Reporter mRNA was produced as early as 3 d after agroinfiltration, and stable expression of the protein was observed along the complete process. These achievements enable further biotechnological steps in these and other genotypes from our breeding program. Full article
(This article belongs to the Special Issue Application of Biotechnology in Crop Improvement)
Show Figures

Figure 1

Figure 1
<p>Somatic embryogenesis and regeneration in temperate japonica rice varieties Platino, Cuarzo, Esmeralda, and Zafiro. (<b>a</b>) Embryogenic calli induction and proliferation achieved through germinating seeds, isolating the zygotic embryos (arrow), and culturing these structures in 2N6 induction medium. Under these conditions, somatic embryos for all varieties reached a strong proliferative phase. (<b>b</b>) A proliferative loop was achieved between primary embryo cell masses (15 d) and embryogenic callus (30 d) with cultivation in 2N6. Calli germination was achieved via culturing embryogenic masses in N6R medium for Platino or N6RN medium for Cuarzo, Esmeralda, and Zafiro. From these cultures, active budding was obtained (“Germination”, green areas), which were isolated and cultured in N6F medium (“Plantlet”). Completely formed plants from these buds are shown at the bottom of the figure (“Whole Plant”). (<b>c</b>) Embryo germination shown as means plus standard error per genotype, including the significance (95% probability level) of the difference between means using asterisk and ns for positive or negative below-threshold results, respectively. The values were averaged over three replicates. In (<b>a</b>,<b>b</b>), bars represent 1 mm.</p>
Full article ">Figure 2
<p><span class="html-italic">Agrobacterium-</span>mediated gene transfer of rice somatic embryos. (<b>a</b>) Geminivirus-based T-DNA vector (pFLC-<span class="html-italic">Myb</span>/<span class="html-italic">HPT</span>/<span class="html-italic">GFP</span>) expressing green fluorescent protein (GFP), <span class="html-italic">StMyb</span>A1, and hygromycin resistance marker genes was used in these <span class="html-italic">Agrobacterium</span> EHA105-mediated gene transfer experiments. (<b>b</b>) Bacterial infections were performed at 15 d of primary calli and embryo masses being cultured in 2N6 medium. GFP-emitting spots were observed in the calli as soon as 4 d after infection. (<b>c</b>) Means of <span class="html-italic">GFP</span> expressing spots among varieties for the response to the gene transfer process. (<b>d</b>) Molecular detection of the <span class="html-italic">gfp</span> transcript in Platino as early as 3 d after infection. (<b>e</b>) Cell masses showing stable <span class="html-italic">GFP</span> expression were kept in this medium supplemented with cefotaxime and carbenicillin. (<b>f</b>) Calli germination and structure formation were achieved under hygromycin selection, allowing for plantlet generation in N6F medium. In a: LIR, <span class="html-italic">Large intergenic region</span> from the <span class="html-italic">Bean yellow dwarf virus</span> (<span class="html-italic">BeYDV</span>); Rep/RepA, geminivirus replicase gene from <span class="html-italic">BeYDV</span>; SIR, <span class="html-italic">Small intergenic region</span> from <span class="html-italic">BeYDV</span>; e<span class="html-italic">GFP</span> and HygR, <span class="html-italic">Green fluorescent protein</span> gene fused to the <span class="html-italic">hygromycin phophotransferase</span> gene; CmYLCV pro, <span class="html-italic">Cestrum yellow leaf curling virus</span> promoter; CaMV35S pro, <span class="html-italic">Cauliflower mosaic virus</span> 35S promoter; StMybA1, potato <span class="html-italic">StMybA</span>1 gene; Kan, <span class="html-italic">Neomycin phosphotransferase</span> gene. In c, means of <span class="html-italic">GFP</span> expressing spots are plotted as means plus standard error over four replicates. In d, St, 2 log molecular weight ladder (New England Biolabs, Ipswich, MA, USA); arrows indicate 100 (lower) and 500 (upper) bps; dpi, days post-agroinfiltration; +, pFLC-<span class="html-italic">Myb</span>/<span class="html-italic">HPT</span>/<span class="html-italic">GFP</span>. White bars represent 1 mm.</p>
Full article ">
16 pages, 5480 KiB  
Article
Development of Zika Virus Mini-Replicon Based Single-Round Infectious Particles as Gene Delivery Vehicles
by Joh-Sin Wu, Ju-Ying Kan, Hsueh-Chou Lai and Cheng-Wen Lin
Viruses 2023, 15(8), 1762; https://doi.org/10.3390/v15081762 - 18 Aug 2023
Cited by 1 | Viewed by 1679
Abstract
Zika virus (ZIKV) is a type of RNA virus that belongs to the Flaviviridae family. We have reported the construction of a DNA-launched replicon of the Asian-lineage Natal RGN strain and the production of single-round infectious particles (SRIPs) via the combination of prM/E [...] Read more.
Zika virus (ZIKV) is a type of RNA virus that belongs to the Flaviviridae family. We have reported the construction of a DNA-launched replicon of the Asian-lineage Natal RGN strain and the production of single-round infectious particles (SRIPs) via the combination of prM/E virus-like particles with the replicon. The main objective of the study was to engineer the ZIKV replicon as mammalian expression vectors and evaluate the potential of ZIKV mini-replicon-based SRIPs as delivery vehicles for heterologous gene expression in vitro and in vivo. The mini-replicons contained various genetic elements, including NS4B, an NS5 methyltransferase (MTase) domain, and an NS5 RNA-dependent RNA polymerase (RdRp) domain. Among these mini-replicons, only ZIKV mini-replicons 2 and 3, which contained the full NS5 and NS4B-NS5 genetic elements, respectively, exhibited the expression of reporters (green fluorescent protein (GFP) and cyan fluorescent protein–yellow fluorescent fusion protein (CYP)) and generated self-replicating RNAs. When the mini-replicons were transfected into the cells expressing ZIKV prM/E, this led to the production of ZIKV mini-replicon-based SRIPs. ZIKV mini-replicon 3 SRIPs showed a significantly higher yield titer and a greater abundance of self-replicating replicon RNAs when compared to ZIKV mini-replicon 2 SRIPs. Additionally, there were disparities in the dynamics of CYP expression and cytotoxic effects observed in various infected cell types between ZIKV mini-replicon 2-CYP and 3-CYP SRIPs. In particular, ZIKV mini-replicon 3-CYP SRIPs led to a substantial decrease in the survival rates of infected cells at a MOI of 2. An in vivo gene expression assay indicated that hACE2 expression was detected in the lung and brain tissues of mice following the intravenous administration of ZIKV mini-replicon 3-hACE2 SRIPs. Overall, this study highlights the potential of ZIKV mini-replicon-based SRIPs as promising vehicles for gene delivery applications in vitro and in vivo. Full article
(This article belongs to the Special Issue Advances in Alphavirus and Flavivirus Research)
Show Figures

Figure 1

Figure 1
<p>The construction and gel electrophoresis analysis of ZIKV mini-replicons containing the GFP reporter. Fragments F1 and F2 were amplified, digested, ligated, and transformed into <span class="html-italic">E. coli</span> cells, resulting in the creation of ZIKV mini-rep 1-GFP. Additional fragments F3 and F4, encompassing NS5 MTase and partial NS4A, NS4B, and NS5 MTase domain, were also amplified, digested, and ligated with ZIKV mini-rep 1-GFP, leading to the generation of ZIKV mini-rep 2-GFP and ZIKV mini-rep 3-GFP clones.</p>
Full article ">Figure 2
<p>The quantitative and qualitative analysis of ZIKV mini-replicon-based SRIPs containing the GFP reporter. Cytopathic effect, GFP expression, and ZIKV RdRp expression in ZIKV prM and E co-expressing cells transfected with indicated ZIKV mini-replicons were assessed using fluorescence microscopy and immunofluorescent staining with anti-ZIKV RdRp antibodies, followed by labeling with Alexa Fluor 546-conjugated secondary antibodies (<b>A</b>). The titer and protein composition of ZIKV mini-rep 2-GFP and 3-GFP SRIPs obtained from the supernatant of packaging cells transfected with ZIKV mini-rep 2-GFP, and 3-GFP were analyzed using TCID50 assay (<b>B</b>), Western blotting with anti-ZIKV-E and prM antibodies (<b>C</b>,<b>D</b>), and dot blot with anti-ZIKV-C antibodies (<b>E</b>). Lane 1, control media; Lane 2, mini-rep 2-GFP SRIP; Lane 3, mini-rep 3-GFP SRIP. ***, <span class="html-italic">p</span>-value &lt; 0.001. Scale bar, 200 μm.</p>
Full article ">Figure 3
<p>Construction of ZIKV mini-replicons containing the CYP reporter (cyan fluorescent protein-linker-yellow fluorescent protein). ZIKV mini-rep 1-no reporter was created by self-ligating the PCR-amplified fragment F5 using the mini-rep 1-GFP template and specific primers. Subsequently, a CFP/YFP gene fragment (F6) was amplified and inserted into ZIKV mini-rep 1-no reporter, resulting in ZIKV mini-rep 1-CYP. The F3 and F4 fragments were individually cloned into specific sites of ZIKV mini-rep 1-CYP, generating ZIKV mini-rep 2-CYP and ZIKV mini-rep 3-CYP, respectively.</p>
Full article ">Figure 4
<p>The analysis of infectivity and quantification of ZIKV mini-rep 2-CYP and 3-CYP SRIPs. The cytopathic effect, CYP expression, and ZIKV RdRp expression in ZIKV prM and E co-expressing cells infected with ZIKV mini-rep 2-CYP and 3-CYP SRIPs were analyzed using fluorescence microscopy and immunofluorescent staining (<b>A</b>). The titer of ZIKV mini-rep 2-CYP and 3-CYP SRIPs was determined by assessing the supernatant of packaging cells infected with ZIKV mini-rep 2-CYP and 3-CYP SRIPs using the TCID50 assay (<b>B</b>). ***, <span class="html-italic">p</span>-value &lt; 0.001. Scale bar, 200 μm.</p>
Full article ">Figure 5
<p>CYP expression kinetics in the cells infected with ZIKV mini-rep 2-CYP and 3-CYP SRIPs at different MOIs. The evaluation was performed in four types of cell lines: HEK293T (<b>A</b>,<b>B</b>), A549 (<b>C</b>,<b>D</b>), TE671 (<b>E</b>,<b>F</b>), and SF268 (<b>G</b>,<b>H</b>) cells. The MOIs tested were 0.1, 1, and 2. To assess the FRET signal of CYP fluorescence, cell lysates were prepared from the infected cells and transferred to a 96-well black plate on Days 1, 2, and 3 post-infection. The FRET signal of CYP was measured using a SpectraMax Multi-Mode Microplate Reader with excitation and emission wavelengths set at 436–528 nm. ***, <span class="html-italic">p</span>-value &lt; 0.001.</p>
Full article ">Figure 6
<p>The survival rates of cells infected with ZIKV mini-rep 2-CYP and 3-CYP SRIPs were evaluated at different MOIs in four types of cell lines: HEK293T (<b>A</b>), A549 (<b>B</b>), TE671 (<b>C</b>), and SF268 (<b>D</b>) cells. The tested MOIs included 0.1, 1, and 2. To assess the survival rates of infected cells, the MTT assay was performed on Day 2 for SF268 cells and on Day 3 for HEK293T, A549, and TE671 cells after infection. **, <span class="html-italic">p</span>-value &lt; 0.01. ***, <span class="html-italic">p</span>-value &lt; 0.001.</p>
Full article ">Figure 7
<p>The assessment of infectivity and quantification of ZIKV mini-rep 2-hACE2 and 3-hACE2 SRIPs. The hACE2 gene was amplified using PCR and inserted into the KpnI/AscI sites of ZV mini-rep 2-GFP and ZV mini-rep 3-GFP, resulting in the creation of mini-replicons named ZV mini-rep 2-hACE2 and ZV mini-rep 3-hACE2, respectively (<b>A</b>). The mRNA expression levels of hACE2 in cells infected with ZIKV mini-rep 2-hACE2 and 3-hACE2 P0 SRIPs were detected using real-time RT-PCR (<b>B</b>). Viral titers of ZIKV mini-rep 2-hACE2 and 3-hACE2 P1 SRIPs were determined using TCID50 assays (<b>C</b>). ***, <span class="html-italic">p</span>-value &lt; 0.001.</p>
Full article ">Figure 8
<p>Immunohistochemistry analysis and histopathology examination of the lung and brain tissues obtained from two groups of mice. The groups included Group I (PBS injections), and Group II (ZV prME/mini-rep 3-hACE2 SRIPs. Immunohistochemistry was performed with primary antibodies against hACE2 after deparaffinization and rehydration of lung tissue sections (<b>A</b>). Tissue morphology was visualized through staining with SPECTRA H&amp;E Stains (<b>B</b>). Scale bar, 50 μm.</p>
Full article ">
14 pages, 2522 KiB  
Article
Hepatitis C Virus Down-Regulates the Expression of Ribonucleotide Reductases to Promote Its Replication
by Chee-Hing Yang, Cheng-Hao Wu, Shih-Yen Lo, Ahai-Chang Lua, Yu-Ru Chan and Hui-Chun Li
Pathogens 2023, 12(7), 892; https://doi.org/10.3390/pathogens12070892 - 29 Jun 2023
Cited by 1 | Viewed by 1173
Abstract
Ribonucleotide reductases (RRs or RNRs) catalyze the reduction of the OH group on the 2nd carbon of ribose, reducing four ribonucleotides (NTPs) to the corresponding deoxyribonucleotides (dNTPs) to promote DNA synthesis. Large DNA viruses, such as herpesviruses and poxviruses, could benefit their replication [...] Read more.
Ribonucleotide reductases (RRs or RNRs) catalyze the reduction of the OH group on the 2nd carbon of ribose, reducing four ribonucleotides (NTPs) to the corresponding deoxyribonucleotides (dNTPs) to promote DNA synthesis. Large DNA viruses, such as herpesviruses and poxviruses, could benefit their replication through increasing dNTPs via expression of viral RRs. Little is known regarding the relationship between cellular RRs and RNA viruses. Mammalian RRs contain two subunits of ribonucleotide reductase M1 polypeptide (RRM1) and two subunits of ribonucleotide reductase M2 polypeptide (RRM2). In this study, expression of cellular RRMs, including RRM1 and RRM2, is found to be down-regulated in hepatitis C virus (HCV)-infected Huh7.5 cells and Huh7 cells with HCV subgenomic RNAs (HCVr). As expected, the NTP/dNTP ratio is elevated in HCVr cells. Compared with that of the control Huh7 cells with sh-scramble, the NTP/dNTP ratio of the RRM-knockdown cells is elevated. Knockdown of RRM1 or RRM2 increases HCV replication in HCV replicon cells. Moreover, inhibitors to RRMs, including Didox, Trimidox and hydroxyurea, enhance HCV replication. Among various HCV viral proteins, the NS5A and/or NS3/4A proteins suppress the expression of RRMs. When these are taken together, the results suggest that HCV down-regulates the expression of RRMs in cultured cells to promote its replication. Full article
(This article belongs to the Special Issue Advances in HCV Research)
Show Figures

Figure 1

Figure 1
<p>Huh 7.5 cells were either mock-infected or infected with infectious HCV (M.O.I. = 0.5). Three or five days after infection, protein samples derived from these cells were analyzed by Western blotting against RRM1 or RRM2. The presence of HCV in the cells was demonstrated by HCV NS3 protein expression.</p>
Full article ">Figure 2
<p>(<b>A</b>) Western blotting analysis of the expression of RRM1 and RRM2 in Huh7 cells and HCV replicon cells with or without interferon treatment. (<b>B</b>) The mRNA levels of RRM1 and RRM2 in Huh7 cells and HCV replicon cells were determined by real-time RT-PCR. (<b>C</b>) Western blotting analysis demonstrated that the HCV subgenomic RNA level is inversely proportionate to the dosage of interferon-alpha used. The HCV NS3 protein expression level reflects the amount of HCV subgenomic RNAs, and actin was applied as a loading control. *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 3
<p>Western botting analysis showed the protein expression level of RRM1 and RRM2 after shRNA knockdown in the Huh7 cells.</p>
Full article ">Figure 4
<p>(<b>A</b>) Knockdown of either RRM1 or RRM2 enhances HCV replication. Western blotting analysis of the protein expression in HCV replicon cells stably transfected with control sh-scramble or various shRNA clones targeting RRM1 or RRM2. (<b>B</b>) The amount of RRM1 or RRM2 mRNAs (left panels) or the amount of HCV subgenomic RNAs (right panels) determined by real-time RT-PCR in HCV replicon cells with various shRNA clones. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 5
<p>Treatment of RRM inhibitors enhanced hepatitis C viral replication. Western blotting analysis of the HCV NS5A protein to reflect the viral replication in the HCVr cells 48 h after the treatment of various amounts of Didox (<b>A</b>), Trimidox (<b>C</b>), or hydroxyurea (<b>D</b>). (<b>B</b>) Real-time RT-PCR analysis of HCV 5′-UTR to reflect viral RNA amount in HCVr cells 48 hrs after the treatment of various amounts of Didox. * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 6
<p>(<b>A</b>) Western blotting analysis of RRM1 or RRM2 protein expression in Huh7 cells stably transfected with control EGFP or HCV NS5A protein. (<b>B</b>,<b>C</b>) Real-time RT-PCR analysis of the mRNA expression of RRM1 (<b>B</b>) or RRM2 (<b>C</b>) 48 hrs after the transfection of various amounts of vectors and/or plasmids expressing NS5A protein into Huh7 cells. (<b>D</b>) Various luciferase reporter assays were performed 48 h after transfection of reporters for promoter activity of RRM1 (upper) or RRM2 (lower) and various amounts of vectors and/or plasmids expressing NS5A protein into Huh7 cells. *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 7
<p>(<b>A</b>) Western blotting analysis of RM1 or RRM2 in Huh7 cells transfected with various plasmids as indicated; 48 h after transfection, protein samples derived from these cells were analyzed with myc tag, RRM1 or RRM2. (<b>B</b>) Various luciferase reporter assays were performed 48 h after transfection of reporters for promoter activity of RRM2 and various amounts of vectors and/or plasmids expressing NS3/4A protein into Huh7 cells. *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">
14 pages, 1734 KiB  
Article
Masitinib Inhibits Hepatitis A Virus Replication
by Reina Sasaki-Tanaka, Toshikatsu Shibata, Mitsuhiko Moriyama, Hirofumi Kogure, Asuka Hirai-Yuki, Hiroaki Okamoto and Tatsuo Kanda
Int. J. Mol. Sci. 2023, 24(11), 9708; https://doi.org/10.3390/ijms24119708 - 3 Jun 2023
Cited by 2 | Viewed by 2844
Abstract
The hepatitis A virus (HAV) infection causes acute hepatitis. HAV also induces acute liver failure or acute-on-chronic liver failure; however, no potent anti-HAV drugs are currently available in clinical situations. For anti-HAV drug screening, more convenient and useful models that mimic HAV replication [...] Read more.
The hepatitis A virus (HAV) infection causes acute hepatitis. HAV also induces acute liver failure or acute-on-chronic liver failure; however, no potent anti-HAV drugs are currently available in clinical situations. For anti-HAV drug screening, more convenient and useful models that mimic HAV replication are needed. In the present study, we established HuhT7-HAV/Luc cells, which are HuhT7 cells stably expressing the HAV HM175-18f genotype IB subgenomic replicon RNA harboring the firefly luciferase gene. This system was made by using a PiggyBac-based gene transfer system that introduces nonviral transposon DNA into mammalian cells. Then, we investigated whether 1134 US Food and Drug Administration (FDA)-approved drugs exhibited in vitro anti-HAV activity. We further demonstrated that treatment with tyrosine kinase inhibitor masitinib significantly reduced both HAV HM175-18f genotype IB replication and HAV HA11-1299 genotype IIIA replication. Masitinib also significantly inhibited HAV HM175 internal ribosomal entry-site (IRES) activity. In conclusion, HuhT7-HAV/Luc cells are adequate for anti-HAV drug screening, and masitinib may be useful for the treatment of severe HAV infection. Full article
(This article belongs to the Special Issue Liver Diseases: From Bench to Bedside)
Show Figures

Figure 1

Figure 1
<p>Construction of the HuhT7-HAV/Luc cells. (<b>A</b>) Insertion construct of HAV/Luc. (<b>B</b>) HuhT7 cells [<a href="#B14-ijms-24-09708" class="html-bibr">14</a>] and HuhT7-HAV/Luc cells were seeded at a density of 1 × 10<sup>5</sup> cells/well in 12-well plates (AGC Techno Glass, Haibaragun, Shizuoka, Japan). HuhT7-HAV/Luc was treated with interferon-α-2a (IFN-α-2a; Sigma-Aldrich, Saint Louis, MO, USA) at 0 or 0.1 μg/mL. After 48 h of incubation, firefly luciferase (Fluc) activity was measured [<a href="#B15-ijms-24-09708" class="html-bibr">15</a>,<a href="#B17-ijms-24-09708" class="html-bibr">17</a>]. Data are expressed as the means and standard deviations of duplicate determinations from two independent experiments. Statistical significance was determined using a two-tailed Student’s <span class="html-italic">t</span> test. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 (versus untreated control). T7, T7 promoter; P2, HAV P2; P3, HAV P3; 5′UTR, 5′ untranslated region; 3′UTR, 3′ untranslated region; polyA, polyA tail.</p>
Full article ">Figure 2
<p>Masitinib is a candidate drug for inhibition of the hepatitis A virus (HAV) HM175-18f genotype IB subgenomic replicon. Analyses of 1134 drugs screened by luciferase assays and dimethylthiazol carboxymethoxyphenyl sulfophenyl tetrazolium (MTS) assays [<a href="#B15-ijms-24-09708" class="html-bibr">15</a>,<a href="#B17-ijms-24-09708" class="html-bibr">17</a>]. The results of the drug screening are shown. All drugs were plotted on a scattergram in which the <span class="html-italic">Y</span>-axis and <span class="html-italic">X</span>-axis indicate the % cell viability of HuhT7-HAV/Luc cells and the relative luciferase activity (fold) of HAV/Luc, respectively. Masitinib, cetylpyridinium chloride, nebivolol, cyclosporine and thonzonium bromide are indicated as red circles. The red horizontal line indicates 90% cell viability. The orange vertical line indicates 0.33-fold relative luciferase activity.</p>
Full article ">Figure 3
<p>Masitinib inhibits hepatitis A virus (HAV) HA11-1299 genotype IIIA infection. Huh7 cells infected with the HAV HA-11-1299 genotype IIIA strain were treated with 10 μM masitinib; 5 µM cetylpyridinium chloride, nebivolol and cyclosporine; and 1 µM thonzonium bromide for 72 h. HAV RNA levels were examined using real-time RT-PCR [<a href="#B15-ijms-24-09708" class="html-bibr">15</a>,<a href="#B17-ijms-24-09708" class="html-bibr">17</a>,<a href="#B20-ijms-24-09708" class="html-bibr">20</a>]. Actin mRNA was used as an internal control. HAV RNA levels were significantly inhibited in masitinib-treated Huh7 cells. Data are expressed as the means and standard deviations of triplicate determinations from three independent experiments. Statistical significance was determined using a two-tailed Student’s <span class="html-italic">t</span> test. ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 4
<p>Immunofluorescence analysis of hepatitis A virus (HAV) HM175-18f genotype IB-infected Huh7 cells. Antibodies against HAV VP1 are shown in green, and nuclei stained with Hoechst are shown in blue [<a href="#B17-ijms-24-09708" class="html-bibr">17</a>,<a href="#B21-ijms-24-09708" class="html-bibr">21</a>]. Positive immunofluorescence staining observed for HAV VP1 in HAV-infected cells but not in the uninfected cells (Mock) [<a href="#B21-ijms-24-09708" class="html-bibr">21</a>]. HAV HM175-18f genotype IB-infected Huh7 cells treated with 10 µM masitinib reduced HAV VP1 staining. HAV HM175-18f genotype IB-infected Huh7 cells treated with 0.1 μg/mL interferon-α-2a (IFN) stained with HAV VP1 as the control. Scale bar: 25 μm.</p>
Full article ">Figure 5
<p>Masitinib inhibits hepatitis A virus (HAV) internal ribosomal entry site (IRES)-mediated translation. (<b>A</b>) COS7-HAV-IRES cells were seeded at a density of 1 × 10<sup>5</sup> cells/well in 12-well plates (AGC Techno Glass). COS7-HAV-IRES was treated with masitinib at 0, 1 and 5 μM. After 48 h of incubation, a luciferase assay was performed. (<b>B</b>) Huh7 cells were seeded 24 h prior to transfection at a density of 1 × 10<sup>5</sup> cells/well in 12-well plates. Cells were transiently transfected with 0.2 μg of pSV40-HAV-IRES using Effectene Transfection Reagent (Qiagen, Chuo-ku, Tokyo, Japan ). After 24 h of transfection, the cells were treated with masitinib at 0, 1, 5 and 10 μM. After 72 h of transfection, luciferase activities were determined. The cytotoxicity of masitinib on COS7-HAV-IRES and Huh7 cells was determined (<b>C</b>,<b>D</b>). COS7-HAV-IRES and Huh7 cells were treated with masitinib at 0, 5 and 10 μM for 48 h. Cell viability was measured via dimethylthiazol carboxymethoxyphenyl sulfophenyl tetrazolium (MTS) assays. Data are expressed as the means and standard deviations of triplicate determinations from three independent experiments. We compared statistical significance using the Student’s <span class="html-italic">t</span> test in two independent groups: a sample group and a control group. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">
13 pages, 3134 KiB  
Review
Recollections of a Helmstetter Disciple
by Alan C. Leonard
Life 2023, 13(5), 1114; https://doi.org/10.3390/life13051114 - 30 Apr 2023
Cited by 1 | Viewed by 1341
Abstract
Nearly fifty years ago, it became possible to construct E. coli minichromosomes using recombinant DNA technology. These very small replicons, comprising the unique replication origin of the chromosome oriC coupled to a drug resistance marker, provided new opportunities to study the regulation of [...] Read more.
Nearly fifty years ago, it became possible to construct E. coli minichromosomes using recombinant DNA technology. These very small replicons, comprising the unique replication origin of the chromosome oriC coupled to a drug resistance marker, provided new opportunities to study the regulation of bacterial chromosome replication, were key to obtaining the nucleotide sequence information encoded into oriC and were essential for the development of a ground-breaking in vitro replication system. However, true authenticity of the minichromosome model system required that they replicate during the cell cycle with chromosome-like timing specificity. I was fortunate enough to have the opportunity to construct E. coli minichromosomes in the laboratory of Charles Helmstetter and, for the first time, measure minichromosome cell cycle regulation. In this review, I discuss the evolution of this project along with some additional studies from that time related to the DNA topology and segregation properties of minichromosomes. Despite the significant passage of time, it is clear that large gaps in our understanding of oriC regulation still remain. I discuss some specific topics that continue to be worthy of further study. Full article
Show Figures

Figure 1

Figure 1
<p>Scheme for construction of minichromosomes. Red color denotes plasmid DNA. Chromosomes are shown in a fast growth configuration used to increase the relative copies of <span class="html-italic">oriC</span>. See text for a description of the figure.</p>
Full article ">Figure 2
<p><span class="html-italic">OriC</span> cloning reports from the weekly lab newsletter “The Flash”. See text.</p>
Full article ">Figure 3
<p>Baby machine analysis of minichromosome cell cycle replication. (<b>A</b>). Diagram of baby machine apparatus and sample collection. Cells growing exponentially were labeled with [<sup>3</sup>H]thymidine for 4 min, bound to a membrane filter, and eluted with glucose/Casamino acids minimal medium. (<b>B</b>). Electrophoretic separation of labeled chromosome and pAL49 minichromosome DNA from new daughter cells. Whole-cell lysates of new daughter cells in the effluent were subjected to agarose gel electrophoresis and fluorography. The radioactive bands corresponding to chromosomal and pAL49 DNA are shown for consecutive 4 min samples of the effluent. Exposure times to the x-ray films were 3 h for the chromosomal bands and 10 days for the minichromosome bands. Modified from reference [<a href="#B10-life-13-01114" class="html-bibr">10</a>].</p>
Full article ">Figure 4
<p>(Left panel) Timing of chromosome and pAL49 minichromosome replication during the division cycle. Exponential-phase cultures of <span class="html-italic">E. coli</span> B/r F26(pAL49) growing in glucose plus Casamino Acids (<b>a</b>), glucose plus six amino acids (<b>b</b>), glucose (<b>c</b>), or glycerol (<b>d</b>) were pulse-labeled and treated as described in the legend to <a href="#life-13-01114-f003" class="html-fig">Figure 3</a>. The radioactivity per cell in minichromosome DNA (closed circles) and total radioactivity per cell (open circles) in newborn cells collected from the effluents of membrane-bound cultures are plotted at the midpoints of the 4 min collection intervals. Abrupt increases in radiolabel (reading right to left) indicate the time of initiation of chromosomal DNA replication. (Right panel). Minichromosome replication during the division cycle of <span class="html-italic">E. coli</span> B/r F(pAL49) growing at different rates. Cells growing exponentially in minimal medium containing glucose plus Casamino acids (<b>a</b>), glucose plus six amino acids (<b>b</b>), or glucose alone (<b>c</b>) were pulse-labeled with [<sup>3</sup>H]thymidine for 4 min, bound to a membrane filter, and eluted with minimal medium of the same composition. Whole-cell lysates of the newborn cells were treated as in <a href="#life-13-01114-f003" class="html-fig">Figure 3</a>. Radioactivity corresponding to closed circular pAL49 minichromosome DNA is shown for consecutive 4 min samples of the effluent at each growth rate. The cell concentrations are also shown, and the vertical interrupted lines indicate the end of each generation of growth on the membrane. Modified from reference [<a href="#B11-life-13-01114" class="html-bibr">11</a>].</p>
Full article ">Figure 5
<p>Cell cycle replication of various plasmids. (<b>A</b>). Fluorograph of radioactive plasmid DNA in newborn cells from the effluent of a membrane filter-bound culture of <span class="html-italic">E. coli</span> B/rF26 containing pSG21 mini-F, pBR322, and pAL70 simultaneously. Cells were grown, pulse-labeled, and prepared as in <a href="#life-13-01114-f003" class="html-fig">Figure 3</a>. In this experiment, all lanes contained lysate from the same number of newborn cells. (<b>B</b>). Radioactive plasmid DNA in newborn cells from a membrane filter-bound culture of <span class="html-italic">E. coli</span> B/r F26 containing F’ lac, pSClO1, pAL49, and pBR322 simultaneously. Cells growing exponentially were pulse labeled and treated as described in <a href="#life-13-01114-f003" class="html-fig">Figure 3</a>. Modified from [<a href="#B16-life-13-01114" class="html-bibr">16</a>].</p>
Full article ">Figure 6
<p>(Left panel) Dye titrations of pAL2 and pBR322 closed circular DNA. pAL2 (<b>A</b>) and pBR322 (<b>B</b>) DNA was isolated from JTT1 recA grown at 37 °C and electrophoresed in gels containing increasing concentrations of ethidium bromide. pAL2 and pBR322 were electrophoresed through 0.6 and 0.8% agarose, respectively. The concentrations of EtBr (in hundredths of micrograms per milliliter) from left to right are 0, 1, 2, 3, 4, 5, 6, 7, 8, 9, 10 and 11. (I) and (II) are supercoiled and relaxed-nicked circular DNA, respectively. (Right panel) Positions of promoter sequences and the direction of transcription on minichromosomes pAL2, pAL20, pAL22, and pAL220 (indicated by arrows). Only pAL20 and pAL22 were able to replicate in <span class="html-italic">E. coli</span> strains (<span class="html-italic">topA, gyrB</span>) with decreased supercoiling. Modified from [<a href="#B27-life-13-01114" class="html-bibr">27</a>].</p>
Full article ">Figure 7
<p>Segregation of minichromosomes and chromosomes during baby machine analysis. Theoretical segregation patterns of radiolabeled DNA are shown for bound (on the horizontal line) and released cells growing with a generation time of C + D minutes. Four successive generations are shown. Old poles lacking attachment sites are shown by thick lines. (Upper panel) Minichromosome segregation is shown with cells that initially contain 20 copies. Average copy numbers of labeled molecules are shown to the left of each cell and the internal distribution of copies is shown above and below the dotted line representing the division septum. The radiolabeled copies in released cells are also shown. (Lower panel) Chromosome segregation. Radiolabeled chromosomal strands (assuming 2 chromosomes in the initially bound cell) are shown in a similar fashion to panel A. Modified from [<a href="#B37-life-13-01114" class="html-bibr">37</a>].</p>
Full article ">
14 pages, 1979 KiB  
Article
A Therapeutic Vaccine Targeting Rat BORIS (CTCFL) for the Treatment of Rat Breast Cancer Tumors
by Dmitri Loukinov, Amanda Laust Anderson, Mikayel Mkrtichyan, Anahit Ghochikyan, Samuel Rivero-Hinojosa, Jo Tucker, Victor Lobanenkov, Michael G. Agadjanyan and Edward L. Nelson
Int. J. Mol. Sci. 2023, 24(6), 5976; https://doi.org/10.3390/ijms24065976 - 22 Mar 2023
Cited by 2 | Viewed by 2242
Abstract
Cancer testis antigens are ideal for tumor immunotherapy due to their testis-restricted expression. We previously showed that an immunotherapeutic vaccine targeting the germ cell-specific transcription factor BORIS (CTCFL) was highly effective in treating aggressive breast cancer in the 4T1 mouse model. Here, we [...] Read more.
Cancer testis antigens are ideal for tumor immunotherapy due to their testis-restricted expression. We previously showed that an immunotherapeutic vaccine targeting the germ cell-specific transcription factor BORIS (CTCFL) was highly effective in treating aggressive breast cancer in the 4T1 mouse model. Here, we further tested the therapeutic efficacy of BORIS in a rat 13762 breast cancer model. We generated a recombinant VEE-VRP (Venezuelan Equine Encephalitis-derived replicon particle) vector-expressing modified rat BORIS lacking a DNA-binding domain (VRP-mBORIS). Rats were inoculated with the 13762 cells, immunized with VRP-mBORIS 48 h later, and then, subsequently, boosted at 10-day intervals. The Kaplan–Meier method was used for survival analysis. Cured rats were re-challenged with the same 13762 cells. We demonstrated that BORIS was expressed in a small population of the 13762 cells, called cancer stem cells. Treatment of rats with VRP-BORIS suppressed tumor growth leading to its complete disappearance in up to 50% of the rats and significantly improved their survival. This improvement was associated with the induction of BORIS-specific cellular immune responses measured by T-helper cell proliferation and INFγ secretion. The re-challenging of cured rats with the same 13762 cells indicated that the immune response prevented tumor growth. Thus, a therapeutic vaccine against rat BORIS showed high efficacy in treating the rat 13762 carcinoma. These data suggest that targeting BORIS can lead to the elimination of mammary tumors and cure animals even though BORIS expression is detected only in cancer stem cells. Full article
(This article belongs to the Special Issue Tumor Microenvironment and Immune Response in Breast Cancer)
Show Figures

Figure 1

Figure 1
<p>Boris expressed in the side population of rat mammary tumor cell line, the 13762 MAT B II, presenting CSC. (<b>A</b>–<b>C</b>) Flow cytometry and sorting gates for side population (SP) and non-side population (NSP). (<b>D</b>,<b>E</b>) Total RNA from the 13762 MAT B II cells, SP (<b>D</b>), and NSP (<b>E</b>) isolated from the 13762 MAT B II cells, testis, and liver were extracted, and BORIS expression was analyzed by qRT-PCR. The results were normalized to GAPDH expression and related to BORIS expression in liver cells. Error bars represent the mean ± SD (n = 3). Two sets of BORIS primers have been used, set 1 (<b>D</b>) and set 2 (<b>E</b>).</p>
Full article ">Figure 2
<p>Schematic representation of the cloning strategy of rat ZF-deleted modified BORIS (mBORIS). The N-terminus region and C-terminus encoding regions of the <span class="html-italic">BORIS</span> gene were amplified by PCR and linked together through the flexible linker using HindIII and BamHI-introduced restriction sites (<b>A</b>). The amplified gene was cloned into the pET-24a <span class="html-italic">E. coli</span> expression vector. mBORIS recombinant protein was purified from <span class="html-italic">E. coli</span> BL21 (DE3) and analyzed in PAGE (<b>B</b>).</p>
Full article ">Figure 3
<p>VRP-encoded mBORIS are expressed by transduced BHK cells. (<b>A</b>) Transduced cells were lysed with RIPA lysis buffer containing protease inhibitors. Lysates were assessed for VRP-encoded protein expression by Western Blot. All lysates from cells transduced with VRP expressed the VRP non-structural protein nsP2. (<b>B</b>) RNA was isolated and used for first-strand cDNA synthesis. The expression of VRP-encoded mBORIS was determined by real-time qPCR. Expression of a particular VRP-encoded tumor antigen sequence was observed only in BHK cells that had been transduced with VRP-encoding that particular sequence. Data are representative of &gt;3 experiments.</p>
Full article ">Figure 4
<p>Therapeutic vaccination with VRP-mBORIS inhibits tumor growth and cures rat breast cancer, inducing antigen-specific cellular responses. (<b>A</b>) The experimental design of the therapeutic study. Rats were treated with VRP-mBORIS 2 days after injection of the 1 × 10<sup>5</sup> 13762 MAT B III BrCA cells. (<b>B</b>) Depicts the tumor size of individual animals on the designated days for VRP-mBORIS treatment and non-treated animals. (<b>C</b>) Depicts survival of rats in two independent experiments. In one experiment, rats were also treated with irrelevant therapeutics VRP-HA. It was observed that 38–50% of rats were completely cured of the disease, while rats without treatment or treated with irrelevant immunogen died on days 25–34. (<b>D</b>) Survived mice were re-challenged by inoculation of the 1 × 10<sup>5</sup> 13762 MAT B III BrCA cells. It was observed that 100% of rats were resistant and did not develop a new tumor. (<b>E</b>) Proliferation upon in vitro restimulation with recombinant mBORIS protein. Splenocytes labeled with CFSE were restimulated in vitro with 160 µg/mL recombinant rat Boris or recombinant GFP in the presence of 50 U/mL IL-2. Four days later, cells were collected and analyzed for proliferation by dilution of the CFSE signal using flow cytometry. Data are represented here as the % of cells that had undergone at least one round of proliferation in response to stimulation with recombinant mBORIS protein—the % of cells that had undergone at least one round of proliferation in response to stimulation with the control protein, recombinant GFP. Data depicted here were generated from cohorts of three animals. Error bars represent standard error, * indicates <span class="html-italic">p</span> &lt; 0.05. (<b>F</b>) IFNγ producing splenocytes were detected after in vitro restimulation with indicated proteins by ELISPOT assay.</p>
Full article ">
12 pages, 779 KiB  
Review
Application of DNA Replicons in Gene Therapy and Vaccine Development
by Kenneth Lundstrom
Pharmaceutics 2023, 15(3), 947; https://doi.org/10.3390/pharmaceutics15030947 - 15 Mar 2023
Cited by 2 | Viewed by 2552
Abstract
DNA-based gene therapy and vaccine development has received plenty of attention lately. DNA replicons based on self-replicating RNA viruses such as alphaviruses and flaviviruses have been of particular interest due to the amplification of RNA transcripts leading to enhanced transgene expression in transfected [...] Read more.
DNA-based gene therapy and vaccine development has received plenty of attention lately. DNA replicons based on self-replicating RNA viruses such as alphaviruses and flaviviruses have been of particular interest due to the amplification of RNA transcripts leading to enhanced transgene expression in transfected host cells. Moreover, significantly reduced doses of DNA replicons compared to conventional DNA plasmids can elicit equivalent immune responses. DNA replicons have been evaluated in preclinical animal models for cancer immunotherapy and for vaccines against infectious diseases and various cancers. Strong immune responses and tumor regression have been obtained in rodent tumor models. Immunization with DNA replicons has provided robust immune responses and protection against challenges with pathogens and tumor cells. DNA replicon-based COVID-19 vaccines have shown positive results in preclinical animal models. Full article
(This article belongs to the Special Issue Plasmid DNA for Gene Therapy and DNA Vaccine Applications)
Show Figures

Figure 1

Figure 1
<p>Schematic presentation of DNA replicon expression vectors. (<b>A</b>). Alphavirus-based DNA replicon represented by a Semliki Forest virus (SFV) vector. CMV, cytomegalovirus promoter; nsP1-4, non-structural protein genes forming the replicase complex; 26S, SFV 26S subgenomic promoter; GoI, gene of interest; pA, polyadenylation signal. (<b>B</b>). Flavivirus-based DNA replicon represented by a Kunjin virus (KUN) vector. 5′ UTR, 5′ untranslated region; C20, the first 20 amino acids of KUN C protein; U, mouse ubiquitin sequence; F, Foot-and-mouth disease virus 2A autoprotease; E22, the last 22 amino acids of KUN E protein; 3′ UTR, 3′ untranslated region; HDVr, Hepatitis delta virus ribozyme.</p>
Full article ">Figure 2
<p>Schematic presentation of DNA replicon systems. Both naked and lipid nanoparticle (LNP) encapsulate DNA replicons have been used. DNA replicon delivered to the nucleus is in vivo transcribed and mRNA translocated to the cytoplasm for translation of the non-structural (nsP1-4) polyprotein, which is processed to individual nsP1-4 proteins forming the replicase complex. The replicase complex is responsible for massive self-replication of RNA, leading to the translation of high levels of recombinant protein or antigen.</p>
Full article ">
15 pages, 3337 KiB  
Article
Construction of a Chikungunya Virus, Replicon, and Helper Plasmids for Transfection of Mammalian Cells
by Mayra Colunga-Saucedo, Edson I. Rubio-Hernandez, Miguel A. Coronado-Ipiña, Sergio Rosales-Mendoza, Claudia G. Castillo and Mauricio Comas-Garcia
Viruses 2023, 15(1), 132; https://doi.org/10.3390/v15010132 - 31 Dec 2022
Cited by 3 | Viewed by 2620
Abstract
The genome of Alphaviruses can be modified to produce self-replicating RNAs and virus-like particles, which are useful virological tools. In this work, we generated three plasmids for the transfection of mammalian cells: an infectious clone of Chikungunya virus (CHIKV), one that codes for [...] Read more.
The genome of Alphaviruses can be modified to produce self-replicating RNAs and virus-like particles, which are useful virological tools. In this work, we generated three plasmids for the transfection of mammalian cells: an infectious clone of Chikungunya virus (CHIKV), one that codes for the structural proteins (helper plasmid), and another one that codes nonstructural proteins (replicon plasmid). All of these plasmids contain a reporter gene (mKate2). The reporter gene in the replicon RNA and the infectious clone are synthesized from subgenomic RNA. Co-transfection with the helper and replicon plasmids has biotechnological/biomedical applications because they allow for the delivery of self-replicating RNA for the transient expression of one or more genes to the target cells. Full article
(This article belongs to the Special Issue Physical Virology - Viruses at Multiple Levels of Complexity)
Show Figures

Figure 1

Figure 1
<p>Construction and evaluation of an expression plasmid for CHIKV. (<b>A</b>) Schematic representation of the plasmids generated here. The SP6-CHIKV plasmid is used for in vitro transcription with SP6 polymerase and contains the full-length genome of the attenuated strain 181/25 of the Chikungunya virus. The genomic CHIKV was inserted into the pVax-CHIKV vector from which the human cytomegalovirus (CMV) immediate early enhancer and promoter sequences and βGH polyA derived from pVAX. The SV40 polyA and hepatitis delta virus (HDV) were inserted, and the genome was moved from the pVAX to the vector pACNR1811. The asterisk means that mKate2 is a reporter gene. (<b>B</b>) Micrographs of HEK-293T cells transfected with the pACNR-CHIKV plasmid in a fluorescent field (mKate2 channel), phase contrast, and a merge of both. (<b>C</b>) Kinetics of the expression of mKate2 from the plasmid pACNR-CHIKV. Bar, 100 µm.</p>
Full article ">Figure 2
<p>The Chikungunya virus plasmid pACNR-CHIKV produces a viral infection. (<b>A</b>) HEK-293T and Vero E6 cells infected with the supernatant of the producer (transfected) cells in fluorescent-field mKate2 expression (Texas Red filter), nuclei stain (DAPI filter), and merge at MOI of 1 for 24 h.p.i. (<b>B</b>) Cytopathic effects induced by the CHIKV infection in Vero E6 cells at MOI of 1. (<b>C</b>) CHIKV plaques on Vero E6 cells stained with crystal violet and (<b>D</b>) and virus titer measured as plaque-forming units (PFUs). Bar, 200 µm.</p>
Full article ">Figure 3
<p>Chikungunya virus pACNR-CHIKV plasmid induces replication and assembly of viral infectious particles. Vero E6 cells infected with Chikungunya virus at a MOI of 1 were assessed at 24 h post-infection. TEM images show Chikungunya (<b>A</b>) four extracellular virions (▲), (<b>B</b>) a viral factory, (<b>C</b>) four virions budding from a membrane (<b>*</b>), and (<b>D</b>) type-II cytopathic vacuoles (arrow, CPV-II). Bar, 100 nm.</p>
Full article ">Figure 4
<p>Generation and evaluation of replicon and helper vector system. (<b>A</b>) Constructions of pACNR-CHIKV, pACNR-Rep, and pVax-Help plasmids. The genomic CHIKV was inserted into the pACNR1811 vector from which the human cytomegalovirus (CMV) immediate early enhancer and promoter sequences, the SV40 polyA, and hepatitis delta virus (HDV) to generate pACNR-CHIKV. The structural proteins were removed to generate pACNR-Rep and derived from pVax-CHIKV and the nonstructural proteins were removed to generate pVax-Help. The asterisk means that mKate2 is a reporter gene. (<b>B</b>) HEK-293T cells transfected with pACNR-Rep, pACNR-Help, and co-transfected with both vectors after 48 h. Fluorescence micrographs with mKate2 expression (mKate2 channel), the morphology of cells in phase contrast, and a merge. Bar, 100 µm.</p>
Full article ">Figure 5
<p>HEK-293T cells from the entry assay infected with the particles generated with the pACNR-Rep and pVax-Help system in (<b>A</b>) fluorescent-field mKate2 expression (Texas Red filter), nuclei stain (DAPI filter), and (<b>C</b>) overlap of mKate2 and nuclei micrographs, demonstrating that the particles generated by the co-transfection of both plasmids result in the expression of the gene of interest (mKate2) in the target cells. The bar indicates the same scale for all images, 50 μm. HEK-293T cells 48 h post-transfection with replicon and helper system were observed by transmission electron microscopy (TEM). (<b>B</b>) Cytopathic vacuole type 1 (arrow, CPV-I), (<b>C</b>) two virions budding from a membrane (<b>*</b>), and (<b>D</b>) two extracellular virions (▲). Bar B–D, 200 nm.</p>
Full article ">
13 pages, 501 KiB  
Review
Therapeutic Applications for Oncolytic Self-Replicating RNA Viruses
by Kenneth Lundstrom
Int. J. Mol. Sci. 2022, 23(24), 15622; https://doi.org/10.3390/ijms232415622 - 9 Dec 2022
Cited by 8 | Viewed by 2707
Abstract
Self-replicating RNA viruses have become attractive delivery vehicles for therapeutic applications. They are easy to handle, can be rapidly produced in large quantities, and can be delivered as recombinant viral particles, naked or nanoparticle-encapsulated RNA, or plasmid DNA-based vectors. The self-replication of RNA [...] Read more.
Self-replicating RNA viruses have become attractive delivery vehicles for therapeutic applications. They are easy to handle, can be rapidly produced in large quantities, and can be delivered as recombinant viral particles, naked or nanoparticle-encapsulated RNA, or plasmid DNA-based vectors. The self-replication of RNA in infected host cells provides the means for generating much higher transgene expression levels and the possibility to apply substantially reduced amounts of RNA to achieve similar expression levels or immune responses compared to conventional synthetic mRNA. Alphaviruses and flaviviruses, possessing a single-stranded RNA genome of positive polarity, as well as measles viruses and rhabdoviruses with a negative-stranded RNA genome, have frequently been utilized for therapeutic applications. Both naturally and engineered oncolytic self-replicating RNA viruses providing specific replication in tumor cells have been evaluated for cancer therapy. Therapeutic efficacy has been demonstrated in animal models. Furthermore, the safe application of oncolytic viruses has been confirmed in clinical trials. Multiple myeloma patients treated with an oncolytic measles virus (MV-NIS) resulted in increased T-cell responses against the measles virus and several tumor-associated antigen responses and complete remission in one patient. Furthermore, MV-CEA administration to patients with ovarian cancer resulted in a stable disease and more than doubled the median overall survival. Full article
(This article belongs to the Section Molecular Pathology, Diagnostics, and Therapeutics)
Show Figures

Figure 1

Figure 1
<p>Schematic illustration of the delivery of self-replicating RNA viruses. Viral particles, naked RNA replicons, lipid nanoparticle (LNP)-encapsulated RNA, or DNA replicons can be used.</p>
Full article ">
Back to TopTop