Nothing Special   »   [go: up one dir, main page]

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

Journals

Article Types

Countries / Regions

Search Results (128)

Search Parameters:
Keywords = RNase H1

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
16 pages, 19868 KiB  
Article
Kinetic Features of Degradation of R-Loops by RNase H1 from Escherichia coli
by Aleksandra A. Kuznetsova, Iurii A. Kosarev, Nadezhda A. Timofeyeva, Darya S. Novopashina and Nikita A. Kuznetsov
Int. J. Mol. Sci. 2024, 25(22), 12263; https://doi.org/10.3390/ijms252212263 - 15 Nov 2024
Viewed by 168
Abstract
R-loops can act as replication fork barriers, creating transcription–replication collisions and inducing replication stress by arresting DNA synthesis, thereby possibly causing aberrant processing and the formation of DNA strand breaks. RNase H1 (RH1) is one of the enzymes that participates in R-loop degradation [...] Read more.
R-loops can act as replication fork barriers, creating transcription–replication collisions and inducing replication stress by arresting DNA synthesis, thereby possibly causing aberrant processing and the formation of DNA strand breaks. RNase H1 (RH1) is one of the enzymes that participates in R-loop degradation by cleaving the RNA strand within a hybrid RNA–DNA duplex. In this study, the kinetic features of the interaction of RH1 from Escherichia coli with R-loops of various structures were investigated. It was found that the values of the dissociation constants Kd were minimal for complexes of RH1 with model R-loops containing a 10–11-nt RNA–DNA hybrid part, indicating effective binding. Analysis of the kinetics of RNA degradation in the R-loops by RH1 revealed that the rate-limiting step of the process was catalytic-complex formation. In the presence of RNA polymerase, the R-loops containing a ≤16-nt RNA–DNA hybrid part were efficiently protected from cleavage by RH1. In contrast, R-loops containing longer RNA–DNA hybrid parts, as a model of an abnormal transcription process, were not protected by RNA polymerase and were effectively digested by RH1. Full article
(This article belongs to the Special Issue Role of RNA Decay in Bacterial Gene Regulation)
Show Figures

Figure 1

Figure 1
<p>The EMSA of the R-loops. The concentration of the RNA primer and of the DNA template and nontemplate strands was 1 µM. The presence/absence of components in the mixture is indicated by +/– signs.</p>
Full article ">Figure 2
<p>MST curves characterizing the interaction of RH1 D10N with R-loops.</p>
Full article ">Figure 3
<p>The RH1-driven cleavage of an RNA primer contained in the R-loops. The R-loops’ concentrations were 0.5 µM, and the RH1 concentration was 25 nM. Lanes show product accumulation for 0, 10, 20, 30, and 40 s and 1, 2, 5, 10, 20, and 30 min. The presence/absence of components in the mixture is indicated by +/– signs.</p>
Full article ">Figure 4
<p>The dependence of the observed rate constant <span class="html-italic">k</span><sub>obs</sub> of the RNA primer cleavage on the RH1 concentration.</p>
Full article ">Figure 5
<p>The RH1-driven cleavage of an RNA primer contained in the R-loops in the presence of the RNAP. The R-loops’ concentrations were 0.5 µM, the RNAP concentration was 1.0 µM, and the RH1 concentration was 25 nM. The lanes show product accumulation for 0, 10, 20, 30, and 40 s and 1, 2, 5, 10, 20, and 30 min. The presence/absence of components in the mixture is indicated by +/– signs.</p>
Full article ">Figure 6
<p>(<b>A</b>) The extent of the cleavage of an RNA primer by RH1 in the absence (gray) or presence (blue) of the RNAP. (<b>B</b>) Observed rate constant <span class="html-italic">k</span><sub>obs</sub> for RNA primer cleavage induced by RH1 in the absence (gray) or presence (blue) of the RNAP. The extent of the cleavage of an RNA primer was calculated at time point 30 min.</p>
Full article ">Figure 7
<p>(<b>A</b>) The crystal structure of a complex of the catalytic domain of <span class="html-italic">Bacillus halodurans</span> RNase HI with RNA–DNA hybrids (PDB ID 5SWM) [<a href="#B68-ijms-25-12263" class="html-bibr">68</a>]. (<b>B</b>) The structural features of the transcription elongation complex.</p>
Full article ">Scheme 1
<p>The kinetic scheme of the RH1-driven cleavage of an RNA primer contained in R-loops. E: RH1; S: an R-loop; E•S: The catalytic complex of the R-loop with the enzyme; P: a reaction product.</p>
Full article ">
16 pages, 3715 KiB  
Article
Screening for Potential Antiviral Compounds from Cyanobacterial Secondary Metabolites Using Machine Learning
by Tingrui Zhang, Geyao Sun, Xueyu Cheng, Cheng Cao, Zhonghua Cai and Jin Zhou
Mar. Drugs 2024, 22(11), 501; https://doi.org/10.3390/md22110501 - 5 Nov 2024
Viewed by 679
Abstract
The secondary metabolites of seawater and freshwater blue-green algae are a rich natural product pool containing diverse compounds with various functions, including antiviral compounds; however, high-efficiency methods to screen such compounds are lacking. Advanced virtual screening techniques can significantly reduce the time and [...] Read more.
The secondary metabolites of seawater and freshwater blue-green algae are a rich natural product pool containing diverse compounds with various functions, including antiviral compounds; however, high-efficiency methods to screen such compounds are lacking. Advanced virtual screening techniques can significantly reduce the time and cost of novel antiviral drug identification. In this study, we used a cyanobacterial secondary metabolite library as an example and trained three models to identify compounds with potential antiviral activity using a machine learning method based on message-passing neural networks. Using this method, 364 potential antiviral compounds were screened from >2000 cyanobacterial secondary metabolites, with amides predominating (area under the receiver operating characteristic curve value: 0.98). To verify the actual effectiveness of the candidate antiviral compounds, HIV virus reverse transcriptase (HIV-1 RT) was selected as a target to evaluate their antiviral potential. Molecular docking experiments demonstrated that candidate compounds, including kororamide, mollamide E, nostopeptolide A3, anachelin-H, and kasumigamide, produced relatively robust non-covalent bonding interactions with the RNase H active site on HIV-1 RT, supporting the effectiveness of the proposed screening model. Our data demonstrate that artificial intelligence-based screening methods are effective tools for mining potential antiviral compounds, which can facilitate the exploration of various natural product libraries. Full article
(This article belongs to the Special Issue Marine Drug Discovery through Molecular Docking)
Show Figures

Figure 1

Figure 1
<p>Screening workflow and model structure. The three gray squares represent the AN model for determining the presence or absence of antiviral material, the RD model for assessing tendency to have anti-RNA or -DNA virus activity, and the model for determining activity against the three viruses, HIV, IAV, and SAR. Linking lines indicate the datasets used to train the different models, with arrows showing model inputs and outputs. The gray shaded area in the upper left corner explains the internal structure of the model, in which the message-passing network (MPN) learned and extracted atom embeddings and bond embeddings that contain local physicochemical information from molecules using molecule structures as topology graphs, with atoms as nodes and bonds as edges; the two types of embeddings of molecules were spliced together to obtain the structural embedding of the molecule. The ratio between molecular electron–ion interaction potential (EIIP) and average quasi-valence number (AQVN) was introduced into the structural embedding as a form of per-position summation to finally generate the molecular embedding and input it into the classifier.</p>
Full article ">Figure 2
<p>Comparison of receiver operator characteristic (ROC) curves for the AN and RD models using message-passing network (MPN) and Morgan fingerprints: red curve: AN model using an MPN fingerprint; cyan curve: RD model using an MPN fingerprint; blue curve: AN model using a Morgan fingerprint; green curve: RD model using a Morgan fingerprint.</p>
Full article ">Figure 3
<p>AN and RD models prediction probability score distributions. Horizontal coordinates: prediction probability score of the model’s output. Vertical coordinates: normalized probability density. Red curve: distribution of predicted probability scores from the AN model; red shading: selected antiviral candidates with predicted probability &gt; 90%. Cyan curve: distribution of predicted scores from the RD model; cyan shading (left): set of molecules selected with a predicted probability of anti-DNA virus activity &lt; 10% and probability of anti-RNA virus activity &gt; 90%; cyan shading (right): set of molecules with a predicted probability of anti-DNA virus activity &gt; 90%.</p>
Full article ">Figure 4
<p>Comparison of the molecular property distributions of cyanobacterial secondary metabolites before and after AN model screening. Distribution plots of the same color indicate comparisons of the same molecular properties. Horizontal coordinates: molecular property metrics. Vertical coordinates: number of molecules. (<b>A</b>,<b>B</b>) number of hydrogen-bond acceptors; (<b>C</b>,<b>D</b>) number of hydrogen-bond donors; (<b>E</b>,<b>F</b>) alcohol–water partition coefficient; (<b>G</b>,<b>H</b>) relative molecular mass; (<b>I</b>,<b>J</b>) number of rotatable bonds; (<b>K</b>,<b>L</b>) topological polar surface area.</p>
Full article ">Figure 5
<p>HIS model prediction probability score distributions. Red, cyan, and yellow curves: distributions of anti-HIV, anti-IAV, and anti-SAR virus predicted probability scores, respectively. Shaded portions: selected molecules with predicted probability &gt; 90%.</p>
Full article ">Figure 6
<p>MPN feature extraction fingerprint dimensionality reduction analysis. (<b>A</b>) Linear principal component analysis (PCA) dimension reduction. (<b>B</b>) Uniform manifold approximation and projection (UMAP) dimension reduction. Scattered dots represent the 2703 molecules in the cyanobacterial secondary metabolite library; red dots: positive outputs from the AN model; blue and yellow dots: kororamide and nostopeptolide A3, respectively, which were used to perform molecular docking validation analysis.</p>
Full article ">Figure 7
<p>Molecular structures of five potential antiviral compounds. (<b>A</b>) kororamide; (<b>B</b>) mollamide E; (<b>C</b>) nostopeptolide A3; (<b>D</b>) anachelin-H; (<b>E</b>) kasumigamide.</p>
Full article ">Figure 8
<p>Interactions between the active binding site of HIV reverse transcriptase and five different molecules. Cyan models: proteins; green discontinuities: hydrogen bonding between small molecules and proteins; yellow cones: pi–cation activity. The sub-icon numbers and corresponding numerators are as follows: (<b>A</b>) kororamide; (<b>B</b>) mollamide E; (<b>C</b>) nostopeptolide A3; (<b>D</b>) anachelin-H; (<b>E</b>) kasumigamide.</p>
Full article ">
15 pages, 4698 KiB  
Article
AtC3H3, an Arabidopsis Non-TZF Gene, Enhances Salt Tolerance by Increasing the Expression of Both ABA-Dependent and -Independent Stress-Responsive Genes
by Hye-Yeon Seok, Sun-Young Lee, Linh Vu Nguyen, Md Bayzid, Yunseong Jang and Yong-Hwan Moon
Int. J. Mol. Sci. 2024, 25(20), 10943; https://doi.org/10.3390/ijms252010943 - 11 Oct 2024
Viewed by 590
Abstract
Salinity causes widespread crop loss and prompts plants to adapt through changes in gene expression. In this study, we aimed to investigate the function of the non-tandem CCCH zinc-finger (non-TZF) protein gene AtC3H3 in response to salt stress in Arabidopsis. AtC3H3, [...] Read more.
Salinity causes widespread crop loss and prompts plants to adapt through changes in gene expression. In this study, we aimed to investigate the function of the non-tandem CCCH zinc-finger (non-TZF) protein gene AtC3H3 in response to salt stress in Arabidopsis. AtC3H3, a gene from the non-TZF gene family known for its RNA-binding and RNase activities, was up-regulated under osmotic stress, such as high salt and drought. When overexpressed in Arabidopsis, AtC3H3 improved tolerance to salt stress, but not drought stress. The expression of well-known abscisic acid (ABA)-dependent salt stress-responsive genes, namely Responsive to Desiccation 29B (RD29B), RD22, and Responsive to ABA 18 (RAB18), and representative ABA-independent salt stress-responsive genes, namely Dehydration-Responsive Element Binding protein 2A (DREB2A) and DREB2B, was significantly higher in AtC3H3-overexpressing transgenic plants (AtC3H3 OXs) than in wild-type plants (WT) under NaCl treatment, indicating its significance in both ABA-dependent and -independent signal transduction pathways. mRNA-sequencing (mRNA-Seq) analysis using NaCl-treated WT and AtC3H3 OXs revealed no potential target mRNAs for the RNase function of AtC3H3, suggesting that the potential targets of AtC3H3 might be noncoding RNAs and not mRNAs. Through this study, we conclusively demonstrated that AtC3H3 plays a crucial role in salt stress tolerance by influencing the expression of salt stress-responsive genes. These findings offer new insights into plant stress response mechanisms and suggest potential strategies for improving crop resilience to salinity stress. Full article
Show Figures

Figure 1

Figure 1
<p>Domain structure of AtC3H3 and multiple sequence alignments among AtC3H3, its paralog, and its orthologs. (<b>a</b>) Black boxes indicate CCCH zinc-finger motifs in AtC3H3. (<b>b</b>) Multiple sequence alignment with protein sequences of AtC3H3, its paralog, and orthologs. Green boxes indicate CCCH zinc-finger motifs conserved among AtC3H3, its paralog, and its orthologs. Conservation rates of amino acids are represented by shading: black shade for 100%, dark gray shade for 80%, and light gray shade for 60%. GI number of each protein sequence is as follows: AtC3H3, 839351; <span class="html-italic">A. lyrata</span>, 9328320; <span class="html-italic">C. sativa</span>, 104739031; <span class="html-italic">C. sativa</span>, 104754644; <span class="html-italic">C. sativa</span>, 104762827; <span class="html-italic">E. salsugineum</span>, 18994367; <span class="html-italic">C. rubella</span>, 17899004; <span class="html-italic">B. oleracea</span>, 106295348; <span class="html-italic">B. rapa</span>, 103844187; <span class="html-italic">B. napus</span>, 106402310; <span class="html-italic">R. sativus</span>, 108812961; <span class="html-italic">T. hassleriana</span>, 104801012; AtC3H26, 817855.</p>
Full article ">Figure 2
<p><span class="html-italic">AtC3H3</span> expression patterns during seedling development and in mature plant organs using quantitative RT-PCR (qRT-PCR). (<b>a</b>) Relative transcript levels of <span class="html-italic">AtC3H3</span> at different developmental stages in wild-type plant (WT) seedlings grown under short-day (SD) conditions. (<b>b</b>) Relative transcript levels of <span class="html-italic">AtC3H3</span> in organs of 49 days after germination (DAG) WT grown under long-day (LD) conditions. RT, roots; RS, rosette leaves; ST, stems; CA, cauline leaves; FC, floral clusters; SI, siliques. In (<b>a</b>,<b>b</b>), <span class="html-italic">GAPc</span> was used to normalize the relative transcript levels. Data represent the average with standard deviations indicated by error bars (<span class="html-italic">n</span> = 3). Statistical differences (<span class="html-italic">p</span> &lt; 0.05) are denoted by different letters.</p>
Full article ">Figure 3
<p>Promoter activity of <span class="html-italic">AtC3H3</span>. (<b>a</b>) Schematic map of <span class="html-italic">β-glucuronidase</span> (<span class="html-italic">GUS</span>)-fused <span class="html-italic">AtC3H3</span> promoter construct. (<b>b</b>) Histochemical GUS assay conducted using transgenic plants harboring <span class="html-italic">GUS</span>-fused <span class="html-italic">AtC3H3</span> promoter construct grown under SD conditions for indicated times. Scale bars indicate 1 cm.</p>
Full article ">Figure 4
<p>Subcellular AtC3H3 localization. (<b>a</b>) Schematic maps of N-terminal or C-terminal sGFP-fused AtC3H3 constructs. (<b>b</b>) Subcellular AtC3H3 localization investigated by transiently expressing sGFP–AtC3H3 and AtC3H3–sGFP constructs in <span class="html-italic">Arabidopsis</span> protoplasts. Left, GFP signal; middle, 4′,6-diamidino-2-phenylindole (DAPI) staining; right, light microscopic image. Scale bars indicate 10 μm.</p>
Full article ">Figure 5
<p><span class="html-italic">AtC3H3</span> and <span class="html-italic">RD29A</span> expression patterns under osmotic stress conditions. (<b>a</b>–<b>c</b>) Relative <span class="html-italic">AtC3H3</span> transcript levels in 10 DAG WT seedlings treated with NaCl (<b>a</b>), mannitol (<b>b</b>), and ABA (<b>c</b>) for indicated times. (<b>d</b>–<b>f</b>) Relative <span class="html-italic">RD29A</span> transcript levels in 10 DAG WT seedlings treated with NaCl (<b>d</b>), mannitol (<b>e</b>), and ABA (<b>f</b>) for indicated times. <span class="html-italic">GAPc</span> was used to normalize the relative transcript levels. <span class="html-italic">AtC3H3</span> or <span class="html-italic">RD29A</span> transcript levels at 0 h of treatment were designated as 1. Data represent the average with standard deviations indicated by error bars (<span class="html-italic">n</span> = 6). Statistical differences (<span class="html-italic">p</span> &lt; 0.05) are denoted by different letters.</p>
Full article ">Figure 6
<p>Response of <span class="html-italic">AtC3H3</span>-overexpressing transgenic plant (<span class="html-italic">AtC3H3</span> OX) seedlings to salt stress. (<b>a</b>) Photographs of WT and <span class="html-italic">AtC3H3</span> OX seedlings incubated under indicated NaCl concentrations for 23 days. (<b>b</b>) Fresh weight (FW) of WT and <span class="html-italic">AtC3H3</span> OX seedlings assessed after 23-day NaCl treatment. (<b>c</b>) Fluorescent image of photosystem II (PS II) activity (<span class="html-italic">F<sub>v</sub></span>/<span class="html-italic">F<sub>m</sub></span>) of WT and <span class="html-italic">AtC3H3</span> OX seedlings incubated under indicated NaCl concentrations for 23 days. (<b>d</b>) <span class="html-italic">F<sub>v</sub></span>/<span class="html-italic">F<sub>m</sub></span> units of WT and <span class="html-italic">AtC3H3</span> OX seedlings assessed after 23-day NaCl treatment. (<b>e</b>) Superoxide accumulation in cotyledons of 10 DAG WT and <span class="html-italic">AtC3H3</span> OX seedlings. Histochemical nitro blue tetrazolium staining carried out after NaCl treatment with indicated concentrations. In (<b>b</b>,<b>d</b>), data represent the average with standard deviations indicated by error bars (<span class="html-italic">n</span> = 24). Statistical differences (<span class="html-italic">p</span> &lt; 0.05) are denoted by different letters.</p>
Full article ">Figure 7
<p>Response of mature <span class="html-italic">AtC3H3</span> OX plants to salt stress. (<b>a</b>) Photograph of WT and <span class="html-italic">AtC3H3</span> OXs treated with indicated NaCl concentrations for 17 days. (<b>b</b>) Survival ratio of WT and <span class="html-italic">AtC3H3</span> OXs after 17-day NaCl treatment. (<b>c</b>) <span class="html-italic">F<sub>v</sub></span>/<span class="html-italic">F<sub>m</sub></span> units of WT and <span class="html-italic">AtC3H3</span> OXs after 17-day NaCl treatment. (<b>d</b>) SPAD values of WT and <span class="html-italic">AtC3H3</span> OXs after 17-day NaCl treatment. In (<b>b</b>–<b>d</b>), data represent the average with standard deviations indicated by error bars (<span class="html-italic">n</span> = 15). Statistical differences (<span class="html-italic">p</span> &lt; 0.05) are denoted by different letters.</p>
Full article ">Figure 8
<p>Representative ABA-dependent and -independent salt stress-responsive gene expression patterns in <span class="html-italic">AtC3H3</span> OXs. Relative <span class="html-italic">AtC3H3</span> (<b>a</b>), <span class="html-italic">RD29B</span> (<b>b</b>), <span class="html-italic">RD22</span> (<b>c</b>), <span class="html-italic">RAB18</span> (<b>d</b>), <span class="html-italic">DREB2A</span> (<b>e</b>), and <span class="html-italic">DREB2B</span> (<b>f</b>) transcript levels in WT and <span class="html-italic">AtC3H3</span> OX seedlings treated with NaCl for indicated times. <span class="html-italic">GAPc</span> was used to normalize the relative transcript levels. Transcript levels of each gene in WT at 0 h of NaCl treatment were set to 1. Data represent the average with standard deviations indicated by error bars (<span class="html-italic">n</span> = 6). Statistical differences (<span class="html-italic">p</span> &lt; 0.05) are denoted by different letters.</p>
Full article ">
16 pages, 5478 KiB  
Protocol
Simultaneous Visualization of R-Loops/RNA:DNA Hybrids and Replication Forks in a DNA Combing Assay
by Miroslav Penchev Ivanov, Heather Zecchini and Petra Hamerlik
Genes 2024, 15(9), 1161; https://doi.org/10.3390/genes15091161 - 3 Sep 2024
Viewed by 1689
Abstract
R-loops, structures that play a crucial role in various biological processes, are integral to gene expression, the maintenance of genome stability, and the formation of epigenomic signatures. When these R-loops are deregulated, they can contribute to the development of serious health conditions, including [...] Read more.
R-loops, structures that play a crucial role in various biological processes, are integral to gene expression, the maintenance of genome stability, and the formation of epigenomic signatures. When these R-loops are deregulated, they can contribute to the development of serious health conditions, including cancer and neurodegenerative diseases. The detection of R-loops is a complex process that involves several approaches. These include S9.6 antibody- or RNAse H-based immunoprecipitation, non-denaturing bisulfite footprinting, gel electrophoresis, and electron microscopy. Each of these methods offers unique insights into the nature and behavior of R-loops. In our study, we introduce a novel protocol that has been developed based on a single-molecule DNA combing assay. This innovative approach allows for the direct and simultaneous visualization of RNA:DNA hybrids and replication forks, providing a more comprehensive understanding of these structures. Our findings confirm the transcriptional origin of the hybrids, adding to the body of knowledge about their formation. Furthermore, we demonstrate that these hybrids have an inhibitory effect on the progression of replication forks, highlighting their potential impact on DNA replication and cellular function. Full article
(This article belongs to the Special Issue DNA Damage Repair in Cancers)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Simultaneous visualization of replication forks and RNA:DNA hybrids in a DNA combing assay. (<b>A</b>) Example fields of view of combed DNA, stained for single-stranded DNA (green), RNA:DNA hybrids/R-loops (red), and replication forks: CldU (cyan), followed by IdU (yellow). RNA:DNA hybrids and replication forks can appear independently or colocalize. (<b>B</b>) Zoomed-in examples of co-staining of RNA:DNA hybrids and replication forks in combed DNA. The dashed hybrid staining indicates that the structures are beyond the resolution limit. Asterisk: one of the larger intact hybrids observed with a length of about 3.5 kb. (<b>C</b>) Effect of transcription inhibition with actinomycin D on RNA:DNA hybrid abundance. 10 μM Actinomycin D was added to cells for 1 h prior to as well as during cell labelling. The number of hybrid spots normalized to DNA or Background area in a technical triplicate are plotted as Mean +/− SD. Statistical difference between non-treated (−) and actinomycin D-treated (+) cells is calculated by one-way ANOVA, (∗∗) adjusted <span class="html-italic">p</span>-value = 0.0054. Hybrids are identified exclusively in the DNA area and their abundance is reduced upon transcriptional inhibition. Background = non-DNA area. (<b>D</b>) Breakdown of technical triplicates in (<b>C</b>) by repeat and S9.6 antibody source.</p>
Full article ">Figure 2
<p>Direct measurement of the inhibitory effect of RNA:DNA hybrid/R-loop collisions on replication fork progression. (<b>A</b>) Two examples of sister forks of different lengths correlated with RNA:DNA hybrid/R-loop colocalization. Left: The left fork, which interferes with more hybrids, is shorter than its sister, which interferes with fewer hybrids. Right: The left fork, which interferes with hybrids is shorter than its sister, which is unchallenged by hybrids. The brightness and contrast of the entire image are adjusted in Photoshop. (<b>B</b>) An example of a replication fork collapse at a cluster of RNA:DNA hybrids/R-loops. The left sister fork collapses at a stretch of hybrids while the right sister fork progresses unimpeded. The brightness and contrast of the entire image are adjusted in Photoshop. (<b>C</b>) Frequency distribution of replication fork lengths (IdU) in DU145 Wild type grouped based on colocalization with at least one hybrid dot (black) or free of hybrids (grey). Manual measurement with FIJI. Gaussian fitting for visualization purposes. (<b>D</b>) Frequency distribution of replication fork lengths (IdU) in DU145 BRCA1 knockout grouped based on colocalization with at least one hybrid dot (black) or free of hybrids (grey). Manual measurement with FIJI. Gaussian fitting for visualization purposes. (<b>E</b>) Frequency distribution of replication fork lengths (IdU) in DU145 ATM knockout grouped based on colocalization with at least one hybrid dot (black) or free of hybrids (grey). Manual measurement with FIJI. Gaussian fitting for visualization purposes. (<b>F</b>) Violin plot visualisation of the data from (<b>C</b>), based on 586 manually measured forks (201 with hybrids, median = 11.87 AU; 385 without hybrids, median = 14.54 AU). Mann Whitney test—Significant, <span class="html-italic">p</span> value &lt; 0.0001 (∗∗∗∗). (<b>G</b>) Violin plot visualisation of the data from (<b>D</b>), based on 544 manually measured forks (165 with hybrids, median = 13.07 AU; 379 without hybrids, median = 15.74 AU). Mann Whitney test—Significant, <span class="html-italic">p</span> value &lt; 0.0001 (∗∗∗∗). (<b>H</b>) Violin plot visualisation of the data from (E), based on 569 manually measured forks (224 with hybrids, median = 11.21 AU; 345 without hybrids, median = 14.54 AU). Mann Whitney test—Significant, <span class="html-italic">p</span> value &lt; 0.0001 (∗∗∗∗). (<b>I</b>) Ratios between lengths of sister forks, pooled from the three cell lines analysed in (<b>C</b>–<b>E</b>). Only pairs in which one fork colocalizes with hybrid(s) and its sister is free of hybrids are analysed. The median ratio between sisters with hybrids and without hybrids is 0.87. Based on 73 sister pairs. Forks with hybrids tend to be shorter than their sisters without hybrids within similar chromatin context.</p>
Full article ">
13 pages, 1402 KiB  
Article
An Improved Bulk DNA Extraction Method for Detection of Helicoverpa armigera (Lepidoptera: Noctuidae) Using Real-Time PCR
by Kayla A. Mollet, Luke R. Tembrock, Frida A. Zink, Alicia E. Timm and Todd M. Gilligan
Insects 2024, 15(8), 585; https://doi.org/10.3390/insects15080585 - 1 Aug 2024
Viewed by 861
Abstract
Helicoverpa armigera is among the most problematic agricultural pests worldwide due to its polyphagy and ability to evolve pesticide resistance. Molecular detection methods for H. armigera have been developed to track its spread, as such methods allow for rapid and accurate differentiation from [...] Read more.
Helicoverpa armigera is among the most problematic agricultural pests worldwide due to its polyphagy and ability to evolve pesticide resistance. Molecular detection methods for H. armigera have been developed to track its spread, as such methods allow for rapid and accurate differentiation from the native sibling species H. zea. Droplet digital PCR (ddPCR) is a preferred method for bulk screening due to its accuracy and tolerance to PCR inhibitors; however, real-time PCR is less expensive and more widely available in molecular labs. Improvements to DNA extraction yield, purity, and throughput are crucial for real-time PCR assay optimization. Bulk DNA extractions have recently been improved to where real-time PCR sensitivity can equal that of ddPCR, but these new methods require significant time and specialized equipment. In this study, we improve upon previously published bulk DNA extraction methods by reducing bench time and materials. Our results indicate that the addition of caffeine and RNase A improves DNA extraction, resulting in lower Cq values during real-time PCR while reducing the processing time and cost per specimen. Such improvements will enable the use of high throughput screening methods across multiple platforms to improve the probability of detection of H. armigera. Full article
(This article belongs to the Section Insect Molecular Biology and Genomics)
Show Figures

Figure 1

Figure 1
<p>Real-time PCR amplification curves representing 70 samples of the RC-buffer extraction and three positive control samples of the O-buffer extraction with and without BP using a ratio of one <span class="html-italic">H. armigera</span> to 50 <span class="html-italic">H. zea</span> legs, two negative control samples of RC-buffer with 50 <span class="html-italic">H. zea</span> legs and no <span class="html-italic">H. armigera</span> legs, and two NTC samples with water instead of DNA. (<b>A</b>–<b>C</b>) show amplification curves for the first, second, and third technical replicates for the <span class="html-italic">H. armigera</span> probe. (<b>D</b>–<b>F</b>) show amplification curves for three technical replicates for the <span class="html-italic">H. armigera</span> probe.</p>
Full article ">Figure 2
<p>Results from the ddPCR assay using EvaGreen to detect <span class="html-italic">H. armigera</span> in a subset of RC-buffer extractions. Each number-labeled sample contained one <span class="html-italic">H. armigera</span> leg and 50 <span class="html-italic">H. zea</span> legs. Sample names and Cq values (averaged from three technical replicates and rounded to two decimal points) correspond to <a href="#app1-insects-15-00585" class="html-app">Supplementary Table S1</a>. Samples labeled “+” are positive controls consisting of O-buffer extractions with one <span class="html-italic">H. armigera</span> leg and 50 <span class="html-italic">H. zea</span> legs. Samples labeled “-” are negative controls consisting of O-buffer extractions with 50 <span class="html-italic">H. zea</span> legs and no <span class="html-italic">H. armigera</span> legs. Positive droplets are blue and negative droplets are grey. Conc (ng/uL) is the calculated starting target DNA concentration.</p>
Full article ">
10 pages, 4642 KiB  
Article
Rational Design of Chimeric Antisense Oligonucleotides on a Mixed PO–PS Backbone for Splice-Switching Applications
by Bao T. Le, Suxiang Chen and Rakesh N. Veedu
Biomolecules 2024, 14(7), 883; https://doi.org/10.3390/biom14070883 - 22 Jul 2024
Viewed by 1108
Abstract
Synthetic antisense oligonucleotides (ASOs) are emerging as an attractive platform to treat various diseases. By specifically binding to a target mRNA transcript through Watson–Crick base pairing, ASOs can alter gene expression in a desirable fashion to either rescue loss of function or downregulate [...] Read more.
Synthetic antisense oligonucleotides (ASOs) are emerging as an attractive platform to treat various diseases. By specifically binding to a target mRNA transcript through Watson–Crick base pairing, ASOs can alter gene expression in a desirable fashion to either rescue loss of function or downregulate pathogenic protein expression. To be clinically relevant, ASOs are generally synthesized using modified analogs to enhance resistance to enzymatic degradation and pharmacokinetic and dynamic properties. Phosphorothioate (PS) belongs to the first generation of modified analogs and has played a vital role in the majority of approved ASO drugs, mainly based on the RNase H mechanism. In contrast to RNase H-dependent ASOs that bind and cleave target mature mRNA, splice-switching oligonucleotides (SSOs) mainly bind and alter precursor mRNA splicing in the cell nucleus. To date, only one approved SSO (Nusinersen) possesses a PS backbone. Typically, the synthesis of PS oligonucleotides generates two types of stereoisomers that could potentially impact the ASO’s pharmaco-properties. This can be limited by introducing the naturally occurring phosphodiester (PO) linkage to the ASO sequence. In this study, towards fine-tuning the current strategy in designing SSOs, we reported the design, synthesis, and evaluation of several stereo-random SSOs on a mixed PO–PS backbone for their binding affinity, biological potency, and nuclease stability. Based on the results, we propose that a combination of PO and PS linkages could represent a promising approach toward limiting undesirable stereoisomers while not largely compromising the efficacy of SSOs. Full article
(This article belongs to the Special Issue RNA Therapeutics)
Show Figures

Figure 1

Figure 1
<p>Structural representations of 2′-O-methyl (2′-OMe) RNA analog on phosphodiester (PO) or phosphorothioate (PS) backbone. (<span class="html-italic">R</span>p) and (<span class="html-italic">S</span>p) represent R and S stereoisomers, respectively.</p>
Full article ">Figure 2
<p>Representative agarose gel images and densitometry analysis (in duplicates) of RT-PCR products amplified from RNA extracted from cells transfected with ASO concentrations at 400, 200, and 100 nM. The original gel images are shown in <a href="#app1-biomolecules-14-00883" class="html-app">Figure S2 (Supplementary Information)</a>. Ne Ctrl: negative control sequence; UT: untreated; NC: PCR sample with no added RNA template.</p>
Full article ">Figure 3
<p>Polyacrylamide gel images of the nuclease stability assay. Time points include 0, 0.25, 0.5, 1, 2, 4, 8, 16, and 24 h. The original gel images are shown in <a href="#app1-biomolecules-14-00883" class="html-app">Figure S3 (Supplementary Information)</a>.</p>
Full article ">
10 pages, 1292 KiB  
Article
Discovery of Benzisothiazolone Derivatives as Bifunctional Inhibitors of HIV-1 Reverse Transcriptase DNA Polymerase and Ribonuclease H Activities
by Alondra Vázquez Rivera, Heather Donald, Mounia Alaoui-El-Azher, John J. Skoko, John S. Lazo, Michael A. Parniak, Paul A. Johnston and Nicolas Sluis-Cremer
Biomolecules 2024, 14(7), 819; https://doi.org/10.3390/biom14070819 - 9 Jul 2024
Viewed by 1179
Abstract
The ribonuclease H (RNase H) active site of HIV-1 reverse transcriptase (RT) is the only viral enzyme not targeted by approved antiretroviral drugs. Using a fluorescence-based in vitro assay, we screened 65,239 compounds at a final concentration of 10 µM to identify inhibitors [...] Read more.
The ribonuclease H (RNase H) active site of HIV-1 reverse transcriptase (RT) is the only viral enzyme not targeted by approved antiretroviral drugs. Using a fluorescence-based in vitro assay, we screened 65,239 compounds at a final concentration of 10 µM to identify inhibitors of RT RNase H activity. We identified 41 compounds that exhibited 50% inhibitory concentration (i.e., IC50) values < 1.0 µM. Two of these compounds, 2-(4-methyl-3-(piperidin-1-ylsulfonyl)phenyl)benzo[d]isothiazol-3(2H)-one (1) and ethyl 2-(2-(3-oxobenzo[d]isothiazol-2(3H)-yl)thiazol-4-yl)acetate (2), which both share the same benzisothiazolone pharmacophore, demonstrate robust antiviral activity (50% effective concentrations of 1.68 ± 0.94 µM and 2.68 ± 0.54, respectively) in the absence of cellular toxicity. A limited structure–activity relationship analysis identified two additional benzisothiazolone analogs, 2-methylbenzo[d]isothiazol-3(2H)-one (3) and N,N-diethyl-3-(3-oxobenzo[d]isothiazol-2(3H)-yl)benzenesulfonamide (4), which also resulted in the inhibition of RT RNase H activity and virus replication. Compounds 1, 2 and 4, but not 3, inhibited the DNA polymerase activity of RT (IC50 values~1 to 6 µM). In conclusion, benzisothiazolone derivatives represent a new class of multifunctional RT inhibitors that warrants further assessment for the treatment of HIV-1 infection. Full article
(This article belongs to the Section Chemical Biology)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Structures of the NNRTI nevirapine and the RNase H active site inhibitors β-thujaplicinol, a hydroxytroplone analog, and 19619. (<b>b</b>) Three dimensional structure of the p66 kDa subunit of HIV-1 RT in complex with β-thujaplicinol (pdb 3IG1).</p>
Full article ">Figure 2
<p>Overview of the HTS campaign to identify the RT RNase H inhibitors.</p>
Full article ">Figure 3
<p>Chemical structures of compounds <b>1</b> and <b>2</b>.</p>
Full article ">Figure 4
<p>Chemical structures of compounds <b>3</b>–<b>11</b>.</p>
Full article ">
31 pages, 8543 KiB  
Article
N-Hydroxypiridinedione: A Privileged Heterocycle for Targeting the HBV RNase H
by Dimitrios Moianos, Maria Makri, Georgia-Myrto Prifti, Aristeidis Chiotellis, Alexandros Pappas, Molly E. Woodson, Razia Tajwar, John E. Tavis and Grigoris Zoidis
Molecules 2024, 29(12), 2942; https://doi.org/10.3390/molecules29122942 - 20 Jun 2024
Viewed by 1479
Abstract
Hepatitis B virus (HBV) remains a global health threat. Ribonuclease H (RNase H), part of the virus polymerase protein, cleaves the pgRNA template during viral genome replication. Inhibition of RNase H activity prevents (+) DNA strand synthesis and results in the accumulation of [...] Read more.
Hepatitis B virus (HBV) remains a global health threat. Ribonuclease H (RNase H), part of the virus polymerase protein, cleaves the pgRNA template during viral genome replication. Inhibition of RNase H activity prevents (+) DNA strand synthesis and results in the accumulation of non-functional genomes, terminating the viral replication cycle. RNase H, though promising, remains an under-explored drug target against HBV. We previously reported the identification of a series of N-hydroxypyridinedione (HPD) imines that effectively inhibit the HBV RNase H. In our effort to further explore the HPD scaffold, we designed, synthesized, and evaluated 18 novel HPD oximes, as well as 4 structurally related minoxidil derivatives and 2 barbituric acid counterparts. The new analogs were docked on the RNase H active site and all proved able to coordinate the two Mg2+ ions in the catalytic site. All of the new HPDs effectively inhibited the viral replication in cell assays exhibiting EC50 values in the low μM range (1.1–7.7 μM) with low cytotoxicity, resulting in selectivity indexes (SI) of up to 92, one of the highest reported to date among HBV RNase H inhibitors. Our findings expand the structure–activity relationships on the HPD scaffold, facilitating the development of even more potent anti-HBV agents. Full article
(This article belongs to the Special Issue Design, Synthesis and Biological Evaluation of Heterocyclic Compounds)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Chemical structures and biological activity of potent example HBV RNase H inhibitors from each chemotype. α-HT: α-Hydroxytropolone; HID: <span class="html-italic">N</span>-Hydroxyisoquinolinedione; HNO: <span class="html-italic">N</span>-Hydroxynapthyridinone; HPD: <span class="html-italic">N</span>-Hydroxypyridinedione [<a href="#B21-molecules-29-02942" class="html-bibr">21</a>,<a href="#B23-molecules-29-02942" class="html-bibr">23</a>,<a href="#B28-molecules-29-02942" class="html-bibr">28</a>,<a href="#B29-molecules-29-02942" class="html-bibr">29</a>]. <b>A24</b> is an approximately equimolar mixture of the E/Z isomers. EC<sub>50</sub>, effective concentration 50%; CC<sub>50</sub>, cytotoxic concentration 50%; SI, selectivity index (CC<sub>50</sub>/EC<sub>50</sub>).</p>
Full article ">Figure 2
<p>Design of <span class="html-italic">N</span>-hydroxypyridinediones (HPDs) oximes: optimization of the oxime side substitution.</p>
Full article ">Figure 3
<p>HPDs chelating Mg<sup>2+</sup> ions. (<b>A</b>) The compound <b>39</b> (green, pKa = 9.609) docked into the active site of HBV P RNase H domain chelating both Mg<sup>2+</sup> ions. The ligand interaction diagram on the right shows two deprotonated hydroxyl groups on the HPD core making 7 salt bridge bonds (<b><span style="color:red">-</span><span style="color:#0070C0">-</span></b>) and 1 metal coordination interaction (<b><span style="color:#A5A5A5">--</span></b>). In (<b>B</b>,<b>C</b>), the compounds <b>56</b> (yellow, pKa = 7.509) and <b>59</b> (pink) make 3 salt bridge (<b><span style="color:red">-</span><span style="color:#0070C0">-</span></b>) interactions with Mg<sup>2+</sup> ions. <b>Right</b>; ligand interaction diagram; <b>Left</b>, surface diagram. A PDB file containing for the HBV RNase H used for docking can be found in [<a href="#B18-molecules-29-02942" class="html-bibr">18</a>]. Docking scores are in kCal/mol.</p>
Full article ">Figure 4
<p>Docking studies of HPDs into the active site of RNase H domain of HBV P. (<b>A</b>) The R-group of the compound <b>49</b> (white) docked into the active site of the HBV P RNase H domain makes a pi–pi cation interaction with H726 in the S3 binding pocket. (<b>B</b>) The compound <b>36</b> (green) R-group makes an h-bond with S750, whereas (<b>C</b>) <b>45</b> (cyan) makes an h-bond with N749 in the S3 binding pocket. <b>Right panel</b>, ligand interaction map; <b>Left panel</b>, surface diagram of docked compound into the active site of HBV P RNase H domain. A PDB file containing for the HBV RNase H used for docking can be found in [<a href="#B18-molecules-29-02942" class="html-bibr">18</a>]. Docking scores are in kCal/mol.</p>
Full article ">Scheme 1
<p>Synthesis of the <span class="html-italic">N</span>-hydroxypyridinediones 33–50, minoxidil analogs <b>55</b>–<b>58</b>, and barbituric acid analogs <b>59</b>–<b>60</b>. Reagents and conditions: (a1) NaH, DMF, 0 °C to RT, overnight; (a2) PPh3, DIAD, dry THF, Ar, 0 °C to RT, overnight–4 days (b) H<sub>2</sub>NNH<sub>2</sub>, DCM, or MeOH, RT, 1–16 h; (c) Et<sub>3</sub>N, dry toluene, 65 °C, 4 h; (d) H<sub>2</sub>, Pd/C 10%, RT, MeOH, 30 min; (e) EtOH, RT, overnight; (f) NaH 60% <span class="html-italic">w</span>/<span class="html-italic">w</span>, neat, 150–180 °C, 3 h–overnight; (g) mCPBA, MeOH, 0 °C, 3 h–overnight; (h) (Ac)<sub>2</sub>O, H<sub>2</sub>SO<sub>4</sub>, 110 °C (reflux), 1.5 h; (i) EtOH, RT, molecular sieves, reflux, 3 days.</p>
Full article ">Scheme 2
<p>Synthesis of hydroxylamines <b>29</b>–<b>31.</b> Reagents and conditions: (a) Br<sub>2</sub>, CHCl<sub>3</sub>, RT, 2 h, 92%; (b) urotropine, CHCl<sub>3</sub>, RT, 4 h, then HCl 37%, EtOH, RT, overnight, 69%; (c) Boc<sub>2</sub>O, NaHCO<sub>3</sub>, MeOH/H<sub>2</sub>O (1:1), RT, 90 min, quant.; (d) NaBH<sub>4</sub>, EtOH, 0 °C, 1 h, 78%; (e) <span class="html-italic">N</span>-hydroxyphthalimide, DEAD, triphenylphosphine, THF, −10 °C to RT, overnight, 94%; (f) 3M HCl in AcOEt, RT, 90 min, 94%; (g) heteroarylacid, TBTU, DIPEA, DMF, RT, overnight, then aq. NH<sub>2</sub>NH<sub>2</sub> (55%) THF, RT, 60–90 min, 67–85%, or aq. MeNH<sub>2</sub> (40%), EtOH/THF 3:1, RT, overnight, 92%.</p>
Full article ">
17 pages, 3790 KiB  
Article
Structure-Based Design of Novel Thiazolone[3,2-a]pyrimidine Derivatives as Potent RNase H Inhibitors for HIV Therapy
by Xuan-De Zhu, Angela Corona, Stefania Maloccu, Enzo Tramontano, Shuai Wang, Christophe Pannecouque, Erik De Clercq, Ge Meng and Fen-Er Chen
Molecules 2024, 29(9), 2120; https://doi.org/10.3390/molecules29092120 - 3 May 2024
Viewed by 1098
Abstract
Ribonuclease H (RNase H) was identified as an important target for HIV therapy. Currently, no RNase H inhibitors have reached clinical status. Herein, a series of novel thiazolone[3,2-a]pyrimidine-containing RNase H inhibitors were developed, based on the hit compound 10i, identified [...] Read more.
Ribonuclease H (RNase H) was identified as an important target for HIV therapy. Currently, no RNase H inhibitors have reached clinical status. Herein, a series of novel thiazolone[3,2-a]pyrimidine-containing RNase H inhibitors were developed, based on the hit compound 10i, identified from screening our in-house compound library. Some of these derivatives exhibited low micromolar inhibitory activity. Among them, compound 12b was identified as the most potent inhibitor of RNase H (IC50 = 2.98 μM). The experiment of magnesium ion coordination was performed to verify that this ligand could coordinate with magnesium ions, indicating its binding ability to the catalytic site of RNase H. Docking studies revealed the main interactions of this ligand with RNase H. A quantitative structure activity relationship (QSAR) was also conducted to disclose several predictive mathematic models. A molecular dynamics simulation was also conducted to determine the stability of the complex. Taken together, thiazolone[3,2-a]pyrimidine can be regarded as a potential scaffold for the further development of RNase H inhibitors. Full article
(This article belongs to the Special Issue Synthesis and Evaluation of Bioactivity of Enzyme Inhibitors)
Show Figures

Figure 1

Figure 1
<p>Previously disclosed representative HIV RNase H inhibitors, chelated with Mg<sup>2+</sup>. The dotted lines in the diagram showed the interaction of the molecule with the Mg<sup>2+</sup>.</p>
Full article ">Figure 2
<p>Docking poses of compounds <b>10i</b> (<b>A</b>,<b>B</b>) at the binding pocket of the HIV-1 RT (PDB code: 3QIP). (<b>C</b>) The discovery of novel thiazolone[3,2-<span class="html-italic">a</span>]pyrimidine derivatives.</p>
Full article ">Figure 3
<p>Scatter plot of the QSAR model. The blue dots in the figure are from the training set and the red dots are from the independent test set. pIC<sub>50</sub> (observer) is the experimental value and pIC<sub>50</sub> (predicted) is the model predicted value. The IC<sub>50</sub> values used for training are from <a href="#molecules-29-02120-t001" class="html-table">Table 1</a>, converted to moles.</p>
Full article ">Figure 4
<p>(<b>A</b>–<b>C</b>) Molecular docking of compound <b>12b</b> with RNase H (PDB code: 3QIP); (<b>D</b>) overlay of <b>12b</b> and <b>10i</b> within the binding pocket of RNase H; (<b>E</b>) the crystal structure of 5,6-dihydroxy-2-[(2-phenyl-1H-indol-3-yl)methyl]pyrimidine-4-carboxylic acid with RNase H (PDB code: 3QIP); (<b>F</b>) overlay of <b>12b</b> and 5,6-dihydroxy-2-[(2-phenyl-1H-indol-3-yl)methyl]pyrimidine-4-carboxylic acid within the binding pocket of RNase H; the Mg<sup>2+</sup> ions are represented by the purple sphere. The interactions between residues, ligands, and Mg<sup>2+</sup> ions are indicated by the yellow lines. The molecular scaffold of 5,6-dihydroxy-2-[(2-phenyl-1H-indol-3-yl)methyl]pyrimidine-4-carboxylic acid is green and the molecular scaffold of compound <b>12b</b> is orange. Important amino acid residues are represented in cyan.</p>
Full article ">Figure 5
<p>(<b>A</b>,<b>B</b>): Protein-RMSF during the process of the molecular dynamic simulation, ChainA: p66, ChainB: p51. (<b>C</b>): Ligand-RMSD. (<b>D</b>): The number of hydrogen bonds with a time range from 0 ps to 50,000 ps.</p>
Full article ">Figure 6
<p>(<b>A</b>) UV spectra of <b>12b</b> in EtOH:dioxane:water = 1:1:1 = 5.21 × 10<sup>−5</sup> M (orange trace) and <b>12b</b> = 5.21 × 10<sup>−5</sup> M + MgCl<sub>2</sub> = 10<sup>−3</sup> M (purple trace). (<b>B</b>) UV spectra of <b>12a</b> in EtOH:dioxane:water = 1:1:1 = 1.24 × 10<sup>−4</sup> M (orange trace) and <b>12a</b> = 1.24 × 10<sup>−4</sup> M + MgCl<sub>2</sub> = 10<sup>−3</sup> M (purple trace).</p>
Full article ">Scheme 1
<p>Synthesis of compounds <b>10a</b>–<b>n</b>, <b>11a</b>–<b>k</b>, and <b>12a</b>–<b>f</b>. Reagents and conditions: (<b>a</b>) SO<sub>2</sub>(NH<sub>2</sub>)<sub>2</sub>, EtOH, 80 °C, 2 h, yield: 70–85%; (<b>b</b>) methyl bromoacetate, pyridine, EtOH, reflux, 3 h, yield: 55–80%; (<b>c</b>) 1. ArNH<sub>2</sub>, NaNO<sub>2</sub>, HCl, H<sub>2</sub>O, 0 °C, 30 min; 2. pyridine, EtOH, 0 to 25 °C, 3 h, yield: 30–90%.</p>
Full article ">
18 pages, 11185 KiB  
Article
Localization of S-Locus-Related Self-Incompatibility in Lycium barbarum Based on BSA Analysis
by Cuiping Wang, Jiali Wu, Yan Gao, Guoli Dai, Xiaohui Shang, Haijun Ma, Xin Zhang, Wendi Xu and Ken Qin
Horticulturae 2024, 10(2), 190; https://doi.org/10.3390/horticulturae10020190 - 18 Feb 2024
Cited by 1 | Viewed by 1307
Abstract
The recognition of pollen and pistil in the self-incompatibility process is generally determined by the interaction between the pollen S gene and pistil S gene located at the S locus. However, the regulatory mechanism of self-incompatibility in goji remains unknown. In this study, [...] Read more.
The recognition of pollen and pistil in the self-incompatibility process is generally determined by the interaction between the pollen S gene and pistil S gene located at the S locus. However, the regulatory mechanism of self-incompatibility in goji remains unknown. In this study, we used the self-compatible strain ‘13–19’ and self-incompatible strain ‘xin9’ from Ningxia as parents to create an F1 hybrid population. Reciprocal cross-pollination was performed within the same plant to evaluate the self-compatibility of the parents and F1 progeny. The parents and progeny were subjected to whole-genome resequencing, and mixed pools of DNA were constructed using 30 self-compatible and 30 self-incompatible individuals. Association analysis using the SNP-index method and Euclidean distance was employed to identify the key candidate region of the S locus. The candidate region was further annotated using the Swiss-Prot database to identify genes within the region. Additionally, transcriptome sequencing data from different organs/tissues, as well as from pistils of self-compatible and self-incompatible strains at control (0 h), short (0.5 h), medium (8 h), and long (48 h) time points after self-pollination and cross-pollination, were analyzed to assess differential gene expression and screen for self-compatibility-related loci. Specific primers were designed for PCR amplification to determine the S-RNase genotypes of the extreme parents. The results revealed that the S locus in goji is located within a 32.2 Mb region on chromosome 2 that contains a total of 108 annotated genes. Differential expression analysis showed that ten genes, including Lba02g01064, were specifically expressed in stamens, with four of them annotated as F-box genes, potentially serving as determinants of self-compatibility in stamens. Lba02g01102 was exclusively expressed in pistils and annotated as an S-RNase gene, likely involved in self-compatibility. The expression of Lba02g01102 in pistils decreased after self-pollination and cross-pollination. Six candidate genes exhibited significant changes after self-pollination and cross-pollination. Both parents and progeny carried two S-RNase alleles, and the S-RNase genotypes showed a significant correlation with self-compatibility, with the self-compatible progeny containing the S8-RNase allele. The identification of the S locus in goji provides molecular markers for future marker-assisted breeding and offers genetic resources for studying the mechanism of self-incompatibility in goji, thus contributing to the improvement of goji varieties. Full article
(This article belongs to the Section Genetics, Genomics, Breeding, and Biotechnology (G2B2))
Show Figures

Figure 1

Figure 1
<p>Breeding flowchart for the plant materials used in this study.</p>
Full article ">Figure 2
<p>Frequency of self-compatibility-related phenotypes. FR: fruit set rate; AFW: average fruit weight; CI: compatibility index; CCI: compared compatibility index.</p>
Full article ">Figure 3
<p>Distribution of ED correlation values on chromosomes. The vertical coordinate is the association value (ED<sup>4</sup>) and the horizontal coordinate is the chromosome position. The colored scatter is the raw association value (ED<sup>4</sup>) for each SNP, the black curve is the association value after sliding window fitting, and the red dashed line is the threshold line (0.27).</p>
Full article ">Figure 4
<p>Distribution of SNP-index association values on chromosomes. Each point corresponds to a SNP site, with the x-axis indicating the chromosomal location of the SNP and the y-axis representing the SNP-Index value. The red and green lines serve as threshold indicators, with red signifying a significance level of 0.01 and green indicating a significance level of 0.05. The black line represents the fitted curve for the ΔSNP-index.</p>
Full article ">Figure 5
<p>Gene differential expression map in the candidate interval. Pi: pistil; St: stamen; Sh: shoot; OL: old leaf; NL: new leaf; GF: green fruit; RF: red fruit.</p>
Full article ">Figure 6
<p>Differential expression analysis of the genes in the positioning interval in the style of self-compatible lines and self-incompatible lines at different times after selfing and cross hybridization. A: unpollinated styles of self-compatible line; B: unpollinated styles of self-incompatible line; AcE: 0.5 h styles of cross hybridization of self-compatible line; AcM: 8 h styles of cross hybridization of self-compatible line; AcL: 48 h styles of cross hybridization of self-compatible line; AsE: 0.5 h styles of selfing of self-compatible line; AsM: 8 h styles of selfing of self-compatible line; AsL: 48 h styles of selfing of self-compatible line; BcE: 0.5 h styles of cross hybridization of self-incompatible line; BcM: 8 h styles of cross hybridization of self-incompatible line; BcL: 48 h styles of cross hybridization of self-incompatible line; BsE: 0.5 h styles of selfing of self-incompatible line; BsM: 8 h styles of selfing of self-incompatible line; BsL: 48 h styles of selfing of self-incompatible line.</p>
Full article ">Figure 7
<p>Identification of S-RNase genotypes of parents and offspring. The x-axis represents the four different genotypes of S-RNase detected (S1: S1-RNase; S2: S2-RNase; S8: S8-RNase; S11: S11-RNase), and the y-axis represents the different strains tested. Light yellow indicates the absence of that genotype, while other colors indicate the presence of that genotype.</p>
Full article ">
17 pages, 4257 KiB  
Article
Modeling and Analysis of HIV-1 Pol Polyprotein as a Case Study for Predicting Large Polyprotein Structures
by Ming Hao, Tomozumi Imamichi and Weizhong Chang
Int. J. Mol. Sci. 2024, 25(3), 1809; https://doi.org/10.3390/ijms25031809 - 2 Feb 2024
Viewed by 1291
Abstract
Acquired immunodeficiency syndrome (AIDS) is caused by human immunodeficiency virus (HIV). HIV protease, reverse transcriptase, and integrase are targets of current drugs to treat the disease. However, anti-viral drug-resistant strains have emerged quickly due to the high mutation rate of the virus, leading [...] Read more.
Acquired immunodeficiency syndrome (AIDS) is caused by human immunodeficiency virus (HIV). HIV protease, reverse transcriptase, and integrase are targets of current drugs to treat the disease. However, anti-viral drug-resistant strains have emerged quickly due to the high mutation rate of the virus, leading to the demand for the development of new drugs. One attractive target is Gag-Pol polyprotein, which plays a key role in the life cycle of HIV. Recently, we found that a combination of M50I and V151I mutations in HIV-1 integrase can suppress virus release and inhibit the initiation of Gag-Pol autoprocessing and maturation without interfering with the dimerization of Gag-Pol. Additional mutations in integrase or RNase H domain in reverse transcriptase can compensate for the defect. However, the molecular mechanism is unknown. There is no tertiary structure of the full-length HIV-1 Pol protein available for further study. Therefore, we developed a workflow to predict the tertiary structure of HIV-1 NL4.3 Pol polyprotein. The modeled structure has comparable quality compared with the recently published partial HIV-1 Pol structure (PDB ID: 7SJX). Our HIV-1 NL4.3 Pol dimer model is the first full-length Pol tertiary structure. It can provide a structural platform for studying the autoprocessing mechanism of HIV-1 Pol and for developing new potent drugs. Moreover, the workflow can be used to predict other large protein structures that cannot be resolved via conventional experimental methods. Full article
(This article belongs to the Section Biochemistry)
Show Figures

Figure 1

Figure 1
<p>Flowchart of HIV-1 NL4.3 Pol dimer structure prediction. Step 1: part of the Pol structure with missing fragments for both chain A and chain B can be predicted based on the reference structure of 7SJX. Steps 2a and 2b: the missing fragments from step 1 can be fixed based on the template-based modeling methods and added to the structure from step 1. Step 3: part of the Pol dimer can be predicted based on the anchor-based method. Steps 4s(1) and 4s(2): IN dimer can be predicted based on the anchor-based methods and the dimer is further optimized with SWISS-MODEL. Steps 4a and 4b: one IN subunit can be assembled into chain A of Pol based on the known IN NTD position. Step 5: RH domain can be assembled into chain B of Pol. Steps 6a and 6b: IN domain can be assembled into chain B of Pol based on the empirical strategy. Step 7: the final NL4.3 Pol dimer can be modeled after large-scale optimization and refinement.</p>
Full article ">Figure 2
<p>Structure comparison between mature and immature proteins. (<b>A</b>) Mature PR structure (PDB ID: 2HB4) aligned to 7SJX, where 2HB4 is colored in orange and the PR domain of 7SJX is colored in marine with rest of the structure colored in gray. (<b>B</b>) Mature RT structure (PDB ID: 1DLO) aligned to 7SJX, where 1DLO is colored in limon and the RT domain of 7SJX is colored in purple with rest of the structure colored in gray.</p>
Full article ">Figure 3
<p>The predicted structure of the PR-RT component in HIV-1 NL4.3 Pol polyprotein. (<b>A</b>) Sequence comparison between HIV-1 NL4.3 Pol and those extracted from 7SJX (chain A: 7sjxA; chain B: 7sjxB), where mutations and missing fragments in the middle are highlighted with a yellow bar, and the last determined positions for both 7sjxA and 7sjxB are highlighted with a blue bar. The PR domain, RT domain, RH domain and IN domain are colored in orange, magenta, red, and green, respectively. (<b>B</b>) Modeled part of NL4.3 Pol dimer consisting of PR + RT + RH in chain A and PR + RT in chain B where chain A is colored in gray and chain B in pale cyan. The modeled mutations, missing fragments from either the middle or the end of sequence are colored in blue, red, and orange, separately.</p>
Full article ">Figure 4
<p>The predicted structure of the full-length HIV-1 NL4.3 Pol chain A. (<b>A</b>) Sequence comparison between RT + RH + IN of HIV-1 NL4.3 Pol and RT + RH of 7sjxA, where mutations and missing fragments in the middle are highlighted with a yellow bar, the last determined position for 7sjxA is highlighted with a blue bar, and the tested starting positions for assembly from the IN domain are highlighted with a purple bar. The RT domain, RH domain, and IN domain are colored in magenta, red, and green, respectively. (<b>B</b>) Modeled chain A of NL4.3 HIV-1 Pol where the PR domain is colored in red, RT domain in green, RH domain in light blue, and IN domain in magenta with the first 49 residues are highlighted with a deep olive color.</p>
Full article ">Figure 5
<p>The predicted structure of the full-length HIV-1 NL4.3 Pol chain B. (<b>A</b>) Sequence comparison between RT + RH of NL4.3 HIV-1 Pol and RT of 7sjxB, where mutations and missing fragments in the middle are highlighted with a yellow bar, the last determined position for 7sjxB is highlighted with a blue bar, and the tested starting positions for assembly from the RH domain are highlighted with a purple bar. The RT domain, RH domain, and IN domain are colored in magenta, red, and green, respectively. (<b>B</b>) Modeled part of chain B in NL4.3 HIV-1 Pol where the PR domain is colored in salmon pink, the RT domain in lemon, and the RH domain in blue.</p>
Full article ">Figure 6
<p>Predicted HIV-1 NL4.3 Pol dimer structure. (<b>A</b>) Front view; (<b>B</b>) 45° rotation along y-axis from (<b>A</b>) where the PR domain is colored in red, the RT domain is green, RH is light blue and IN is magenta for chain A, while for chain B, the same domains as chain A are colored in salmon pink, lemon, blue, and hot pink, respectively.</p>
Full article ">Figure 7
<p>Comparison of PISA radar chart between HIV-1 NL4.3 Pol dimer and published partial HIV Pol dimer. (<b>A</b>) Radar chart from modeled NL4.3 HIV-1 Pol dimer, where NSB refers to the number of salt bridges, NHB refers to the number of hydrogen bonds, HYP refers to hydrophobic <span class="html-italic">p</span>-value, TBE refers to total binding energy, SOE refers to solvation energy, INA refers to interface area, and NDB refers to the number of disulfide bonds. If most points exceed the 50% threshold, it indicates that the interface is likely to be significant for biological assembly. Otherwise, interfaces with radar area, fitting within the 50% threshold, are more likely to be artefacts of crystal packing. (<b>B</b>) Radar chart from reference structure of 7SJX.</p>
Full article ">
13 pages, 1704 KiB  
Article
Complement Suppresses the Initial Type 1 Interferon Response to Ocular Herpes Simplex Virus Type 1 Infection in Mice
by Daniel J. J. Carr, Adrian Filiberti and Grzegorz B. Gmyrek
Pathogens 2024, 13(1), 74; https://doi.org/10.3390/pathogens13010074 - 13 Jan 2024
Viewed by 2058
Abstract
The complement system (CS) contributes to the initial containment of viral and bacterial pathogens and clearance of dying cells in circulation. We previously reported mice deficient in complement component 3 (C3KO mice) were more sensitive than wild-type (WT) mice to ocular HSV-1 infection, [...] Read more.
The complement system (CS) contributes to the initial containment of viral and bacterial pathogens and clearance of dying cells in circulation. We previously reported mice deficient in complement component 3 (C3KO mice) were more sensitive than wild-type (WT) mice to ocular HSV-1 infection, as measured by a reduction in cumulative survival and elevated viral titers in the nervous system but not the cornea between days three and seven post infection (pi). The present study was undertaken to determine if complement deficiency impacted virus replication and associated changes in inflammation at earlier time points in the cornea. C3KO mice were found to possess significantly (p < 0.05) less infectious virus in the cornea at 24 h pi that corresponded with a decrease in HSV-1 lytic gene expression at 12 and 24 h pi compared to WT animals. Flow cytometry acquisition found no differences in the myeloid cell populations residing in the cornea including total macrophage and neutrophil populations at 24 h pi with minimal infiltrating cell populations detected at the 12 h pi time point. Analysis of cytokine and chemokine content in the cornea measured at 12 and 24 h pi revealed that only CCL3 (MIP-1α) was found to be different between WT and C3KO mice with >2-fold increased levels (p < 0.05, ANOVA and Tukey’s post hoc t-test) in the cornea of WT mice at 12 h pi. C3KO mouse resistance to HSV-1 infection at the early time points correlated with a significant increase in type I interferon (IFN) gene expression including IFN-α1 and IFN-β and downstream effector genes including tetherin and RNase L (p < 0.05, Mann–Whitney rank order test). These results suggest early activation of the CS interferes with the induction of the type I IFN response and leads to a transient increase in virus replication following corneal HSV-1 infection. Full article
Show Figures

Figure 1

Figure 1
<p><b>HSV-1 replication is reduced in the cornea of C3KO mice</b>. (<b>A</b>) Titers of HSV-1 in the cornea of WT and C3KO mice at 24 h PI as determined by plaque assay. (<b>B</b>–<b>D</b>) HSV-1 lytic gene expression including (<b>B</b>) ICP27, (<b>C</b>) TK, and (<b>D</b>) gB at the indicated time point PI as determined by real-time RT-PCR. The mean ± SEM (n = 6–8 samples/time point) is plotted for each time point. Each graph is a summary of the results from 2–3 experiments. **** <span class="html-italic">p</span> &lt; 0.0001, ** <span class="html-italic">p</span> &lt; 0.01, * <span class="html-italic">p</span> &lt; 0.05 comparing the two groups/time point as determined by Bonferroni–Dunn <span class="html-italic">t</span>-test method.</p>
Full article ">Figure 2
<p><b>C3 levels peak in the cornea of WT mice at 12 h PI.</b> WT and C3KO mice were infected with HSV-1 (500 PFU/cornea). At 12 or 24 h PI, the mice were exsanguinated and the corneas were removed and processed for C3 content by ELISA. The results are depicted as mean + SEM (n = 4–8 samples/time point) and plotted for each time point. The 0 h PI time point represents uninfected mice to serve as background control. The graph is a summary of the results from 2–3 experiments. ** <span class="html-italic">p</span> &lt; 0.01, * <span class="html-italic">p</span> &lt; 0.05 comparing the indicated groups as determined by the Holm–Sidak multiple <span class="html-italic">t</span>-test method.</p>
Full article ">Figure 3
<p><b>Myeloid cell infiltration into the cornea of WT and C3KO mice does not differ at 24 h PI.</b> WT and C3KO mice were infected with HSV-1 (500 PFU/cornea). At 24 h PI, the mice were exsanguinated, and the cornea was removed and processed for myeloid cell content by flow cytometry. The gating strategy is shown for monocyte/macrophage and neutrophil populations in panel (<b>A</b>). Panel (<b>B</b>) shows the summary of myeloid cell types, n = 5–6/group, * <span class="html-italic">p</span> &lt; 0.05 comparing the indicated group as determined by Mann–Whitney rank order test.</p>
Full article ">Figure 4
<p><b>MIP-1α/CCL3 is elevated early in the cornea of WT mice in response to HSV-1 infection.</b> WT and C3KO mice were infected with HSV-1 (500 PFU/cornea). At 12 or 24 h PI, the mice were exsanguinated, and the cornea was removed, homogenized, and clarified supernatant was assessed for analyte content by suspension array. Bars represent mean ± SEM, n = 8–10/group/time point from 3 experiments. * <span class="html-italic">p</span> &lt; 0.05 comparing the indicated group as determined by ANOVA and Tukey’s post-hoc <span class="html-italic">t</span>-test.</p>
Full article ">Figure 5
<p><b>IFN-β and RNase L expression are elevated in the cornea of C3KO mice early post HSV-1 infection.</b> C57BL/6 WT and C3KO male and female mice (n = 6–10/group/time point) were infected with HSV-1 McKrae (500–1000 PFU/cornea). At the indicated time PI, the mice were exsanguinated and the corneas were processed for mRNA analysis by real-time RT-PCR for relative expression of (<b>A</b>) IFN-α1, (<b>B</b>) IFN-β, (<b>C</b>) PKR, (<b>D</b>) RNase L. The results are expressed as the mean ± SEM relative value as determined using the ∆∆C<sub>t</sub> method. ** <span class="html-italic">p</span> &lt; 0.01, * <span class="html-italic">p</span> &lt; 0.05 comparing the indicated groups at the indicated time point as determined by Holm–Sidak multiple <span class="html-italic">t</span>-test method.</p>
Full article ">
14 pages, 2399 KiB  
Article
Biochemical Analysis to Understand the Flooding Tolerance of Mutant Soybean Irradiated with Gamma Rays
by Setsuko Komatsu, Tiantian Zhou and Yuhi Kono
Int. J. Mol. Sci. 2024, 25(1), 517; https://doi.org/10.3390/ijms25010517 - 30 Dec 2023
Viewed by 1088
Abstract
Flooding stress, which reduces plant growth and seed yield, is a serious problem for soybean. To improve the productivity of flooded soybean, flooding-tolerant soybean was produced by gamma-ray irradiation. Three-day-old wild-type and mutant-line plants were flooded for 2 days. Protein, RNA, and genomic [...] Read more.
Flooding stress, which reduces plant growth and seed yield, is a serious problem for soybean. To improve the productivity of flooded soybean, flooding-tolerant soybean was produced by gamma-ray irradiation. Three-day-old wild-type and mutant-line plants were flooded for 2 days. Protein, RNA, and genomic DNA were then analyzed based on oppositely changed proteins between the wild type and the mutant line under flooding stress. They were associated with cell organization, RNA metabolism, and protein degradation according to proteomic analysis. Immunoblot analysis confirmed that the accumulation of beta-tubulin/beta-actin increased in the wild type under flooding stress and recovered to the control level in the mutant line; however, alpha-tubulin increased in both the wild type and the mutant line under stress. Ubiquitin was accumulated and genomic DNA was degraded by flooding stress in the wild type; however, they were almost the same as control levels in the mutant line. On the other hand, the gene expression level of RNase H and 60S ribosomal protein did not change in either the wild type or the mutant line under flooding stress. Furthermore, chlorophyll a/b decreased and increased in the wild type and the mutant line, respectively, under flooding stress. These results suggest that the regulation of cell organization and protein degradation might be an important factor in the acquisition of flooding tolerance in soybean. Full article
(This article belongs to the Collection Feature Papers in Molecular Plant Sciences)
Show Figures

Figure 1

Figure 1
<p>Functional categories of proteins with differential abundance in the mutant line compared with the wild type under flooding stress or non-flooding conditions. Functional categories of changed proteins were determined using MapMan bin codes (<a href="#app1-ijms-25-00517" class="html-app">Supplemental Tables S1 and S2</a>). (<b>A</b>) Changed proteins in mutant-line compared with wild-type soybean under non-flooding conditions. (<b>B</b>) Changed proteins in mutant-line compared with wild-type soybean under flooding stress. Red and blue columns show increased and decreased proteins. Abbreviations: TCA, tricarboxylic acid cycle; PTM, posttranslational modification.</p>
Full article ">Figure 2
<p>Functional categories of proteins with differential abundance in the mutant line or wild type under flooding stress compared with non-flooding conditions. Functional categories of changed proteins were determined using MapMan bin codes (<a href="#app1-ijms-25-00517" class="html-app">Supplemental Tables S1 and S2</a>). (<b>A</b>) Changed proteins in wild-type soybean under flooding stress compared with non-flooding conditions. (<b>B</b>) Changed proteins in mutant-line soybean under flooding stress compared with non-flooding conditions. Red and blue columns show increased and decreased proteins. Abbreviations: TCA, tricarboxylic acid cycle; OPP, oxidative pentose phosphate; mitoETC, mitochondrial electron transport chain; PTM, posttranslational modification; and aa met, amino acid metabolism.</p>
Full article ">Figure 3
<p>Immunoblot analysis of ubiquitin in the mutant line under flooding stress. Proteins were extracted from the roots, including hypocotyl, separated on SDS-polyacrylamide gel by electrophoresis and transferred onto a membrane. The membrane was cross-reacted with an anti-ubiquitin antibody. The Coomassie brilliant blue staining pattern was used as a loading control (<b>A</b>) (<a href="#app1-ijms-25-00517" class="html-app">Figure S2A</a>). The integrated densities of the bands were calculated using ImageJ software (<b>B</b>). Data are shown as the means ± SD from 3 independent biological replicates (<a href="#app1-ijms-25-00517" class="html-app">Figure S3</a>). Student’s <span class="html-italic">t</span>-test was used to compare values between control and treatment as well as wild type and mutant line under flooding stress. Asterisks indicate a significant change (* <span class="html-italic">p</span> ≤ 0.05, ** <span class="html-italic">p</span> ≤ 0.01).</p>
Full article ">Figure 4
<p>Analysis of genomic DNA degradation in the roots of the mutant line under flooding stress. After flooding stress, genomic DNA was extracted from the roots of the wild type and the mutant line. (<b>A</b>) The concentration of genomic DNA. (<b>B</b>) The pattern of agarose gel electrophoresis of extracted genomic DNA. Data are shown as the means ± SD from 3 independent biological replicates. Student’s <span class="html-italic">t</span>-test was used to compare values between control and treatment as well as wild type and mutant line under flooding stress. Asterisks indicate a significant change (* <span class="html-italic">p</span> ≤ 0.05, ** <span class="html-italic">p</span> ≤ 0.01).</p>
Full article ">Figure 5
<p>Immunoblot analysis of proteins involved in cell organization in the mutant line under flooding stress. Proteins were extracted from the root and hypocotyl, separated on SDS-polyacrylamide gel by electrophoresis and transferred onto a membrane. The membrane was cross-reacted with anti-alpha-tubulin, beta-tubulin, and beta-actin antibodies. The Coomassie brilliant blue staining pattern was used as a loading control (<a href="#app1-ijms-25-00517" class="html-app">Supplemental Figure S2B</a>). The integrated densities of the bands were calculated using ImageJ software with 3 independent biological replicates (<a href="#app1-ijms-25-00517" class="html-app">Figures S4–S6</a>). Data are shown as the means ± SD from 3 independent biological replicates. Student’s <span class="html-italic">t</span>-test was used to compare values between control and treatment as well as wild type and mutant line under flooding stress. Asterisks indicate a significant change (* <span class="html-italic">p</span> ≤ 0.05).</p>
Full article ">Figure 6
<p>Gene expression of proteins involved in RNA metabolism in the mutant line under flooding stress. Gene expression analysis of <span class="html-italic">60S ribosome protein</span> and <span class="html-italic">RNase H</span> in the wild type and the mutant line with or without flooding stress was performed. (<b>A</b>) <span class="html-italic">60S ribosome protein</span>- and (<b>B</b>) <span class="html-italic">RNase H</span>-specific oligonucleotides were used to amplify transcripts from total RNA isolated from the roots and hypocotyl. <span class="html-italic">18S rRNA</span> was used as an internal control (<b>C</b>). Data are shown as the means ± SD from 3 independent biological replicates. Student’s <span class="html-italic">t</span>-test was used to compare values between control and treatment as well as wild type and mutant line under flooding stress (** <span class="html-italic">p</span> ≤ 0.01).</p>
Full article ">Figure 7
<p>The contents of chlorophylls <span class="html-italic">a</span> and <span class="html-italic">b</span> in the hypocotyl of the mutant line under flooding stress. Photograph of the upper part of soybean plants with and without flooding stress (<b>A</b>). Size bar indicates 1 cm. Contents of chlorophylls a and b of the hypocotyl of soybean with and without flooding stress (<b>B</b>). Chlorophylls <span class="html-italic">a</span> and <span class="html-italic">b</span> extracted from the hypocotyl of the wild type and the mutant line were measured. Data are shown as the means ± SD from 3 independent biological replicates (<a href="#app1-ijms-25-00517" class="html-app">Figures S7–S9</a>). Student’s <span class="html-italic">t</span>-test was used to compare values between control and treatment as well as wild type and mutant line under flooding stress. Asterisks indicate a significant change (** <span class="html-italic">p</span> ≤ 0.01).</p>
Full article ">
17 pages, 4085 KiB  
Article
Charge-Complementary Polymersomes for Enhanced mRNA Delivery
by HakSeon Kim, Yu-Rim Ahn, Minse Kim, Jaewon Choi, SoJin Shin and Hyun-Ouk Kim
Pharmaceutics 2023, 15(12), 2781; https://doi.org/10.3390/pharmaceutics15122781 - 15 Dec 2023
Cited by 1 | Viewed by 1705
Abstract
Messenger RNA (mRNA) therapies have emerged as potent and personalized alternatives to conventional DNA-based therapies. However, their therapeutic potential is frequently constrained by their molecular instability, susceptibility to degradation, and inefficient cellular delivery. This study presents the nanoparticle “ChargeSome” as a novel solution. [...] Read more.
Messenger RNA (mRNA) therapies have emerged as potent and personalized alternatives to conventional DNA-based therapies. However, their therapeutic potential is frequently constrained by their molecular instability, susceptibility to degradation, and inefficient cellular delivery. This study presents the nanoparticle “ChargeSome” as a novel solution. ChargeSomes are designed to protect mRNAs from degradation by ribonucleases (RNases) and enable cell uptake, allowing mRNAs to reach the cytoplasm for protein expression via endosome escape. We evaluated the physicochemical properties of ChargeSomes using 1H nuclear magnetic resonance, Fourier-transform infrared, and dynamic light scattering. ChargeSomes formulated with a 9:1 ratio of mPEG-b-PLL to mPEG-b-PLL-SA demonstrated superior cell uptake and mRNA delivery efficiency. These ChargeSomes demonstrated minimal cytotoxicity in various in vitro structures, suggesting their potential safety for therapeutic applications. Inherent pH sensitivity enables precise mRNA release in acidic environments and structurally protects the encapsulated mRNA from external threats. Their design led to endosome rupture and efficient mRNA release into the cytoplasm by the proton sponge effect in acidic endosome environments. In conclusion, ChargeSomes have the potential to serve as effective secure mRNA delivery systems. Their combination of stability, protection, and delivery efficiency makes them promising tools for the advancement of mRNA-based therapeutics and vaccines. Full article
Show Figures

Figure 1

Figure 1
<p>Morphological analysis and composition ratios of ChargeSomes at various ratios. (<b>a</b>) ChargeSomes with varying mPEG-b-PLL to mPEG-b-PLL-SA ratios (1:9, 3:7, 5:5, 7:3, and 9:1) were analyzed using TEM and negative staining. The aforementioned ratios serve as indicators of the hierarchical organization and interrelationships among the various components of ChargeSomes. The scale bar corresponds to 200 nm. (<b>b</b>) Size distribution of ChargeSomes determined by DLS. (<b>c</b>) Zeta-potential analysis of ChargeSomes using a zeta-potential particle size analyzer. I, II, III, IV, and V represent the zeta potential of the respective formation ratios of ChargeSomes (1:9, 3:7, 5:5, 7:3, and 9:1).</p>
Full article ">Figure 2
<p>Effects of ChargeSomes on RAW 264.7 cell viability. (<b>a</b>) The cell viability of RAW 264.7 cells at 12 h (green), and 24 h (yellow) post-treatment with ChargeSome particles at various concentrations was evaluated using the EZ-Cytox assay. Student’s <span class="html-italic">t</span>-test and one-way ANOVA were used for statistical analysis. Data are presented as means ± S.D. (n = 5; *** <span class="html-italic">p</span> &lt; 0.001). (<b>b</b>) RAW 264.7 cells underwent treatment with lipofectamine and distinct fractions of ChargeSome particles to enable a comparison of cell viability. I, II, III, IV, and V correspond to the mPEG-b-PLL/mPEG-b-PLL-SA ratios of 1:9, 3:7, 5:5, 7:3, and 9:1, respectively, while VI denotes lipofectamine.</p>
Full article ">Figure 3
<p>Cell uptake of OVA–FITC-encapsulated ChargeSomes in RAW 264.7 cells observed via CLSM. RAW 264.7 cells were incubated with OVA–FITC-encapsulated ChargeSomes (green) for 6 h, and lysosomes and nuclei were stained using Lysotracker-DND-99 (red) and Hoechst 33342 (blue), respectively. Scale bars represent 10 um.</p>
Full article ">Figure 4
<p>Comparative analysis of cell fluorescence uptake conducted using a flow cytometer. RAW 264.7 cells were exposed to OVA–FITC-only and OVA–FITC-encapsulated ChargeSomes for 6 h. (<b>a</b>) Histograms of the samples (ratio of mPEG-b-PLL to mPEG-b-PLL-SA). (<b>b</b>) Median fluorescence intensity (MFI) values for the samples. Data are presented as means ± S.D. (n = 3).</p>
Full article ">Figure 5
<p>Cell uptake and endosomal escape of ChargeSomes. Negative control and OVA–FITC-encapsulated ChargeSome samples were administered and cultured for 1, 2, 4, and 6 h. Subsequently, endosomal escape of the particles was confirmed by employing a confocal laser-scanning microscope. RAW 264.7 cells were incubated with OVA–FITC-encapsulated ChargeSomes (green), lysosomes, and nuclei were stained using Lysotracker-DND-99 (red) and Hoechst 33342 (blue), respectively. Scale bars represent 10 um.</p>
Full article ">Figure 6
<p>Cell uptake and EGFP mRNA expression. (<b>a</b>) RAW 264.7 cells were treated with EGFP mRNA-encapsulated ChargeSomes and naked mRNA separately, followed by CLSM analysis after 24 h of incubation. The efficiency of antigen cell absorption was determined by flow cytometry to assess EGFP expression. RAW 264.7 cells were expressed with EGFP mRNA-encapsulated ChargeSomes (green) and nuclei were stained using Hoechst 33342 (blue), respectively. (<b>b</b>) Histogram representing the negative control (N.C) and each sample. (<b>c</b>) Median fluorescence intensity (MFI) for both the negative control and each sample. Student’s <span class="html-italic">t</span>-test and one-way ANOVA were used for statistical analysis. Data are presented as means ± S.D. (n = 3; *** <span class="html-italic">p</span> &lt; 0.001). Scale bars represent 10 um.</p>
Full article ">Scheme 1
<p>ChargeSome nanoparticles are created using a charged polymer. These nanoparticles are pH-sensitive, enabling them to disintegrate at a pH of 5.5 or lower, leading to the release of encapsulated genes. Upon cell entry, these nanoparticles are taken up in the form of endosomes. Once inside the endosomes with a pH below 5.5, the nanoparticles disintegrate, causing destabilization of the endosome membrane and facilitating their release into the cytoplasm through endosomal escape. The mechanism of this process is depicted in a scheme illustrating particle behavior within a cell.</p>
Full article ">
17 pages, 2741 KiB  
Article
Effective Antiviral Application of Antisense in Plants by Exploiting Accessible Sites in the Target RNA
by Cornelia Gruber, Torsten Gursinsky, Selma Gago-Zachert, Vitantonio Pantaleo and Sven-Erik Behrens
Int. J. Mol. Sci. 2023, 24(24), 17153; https://doi.org/10.3390/ijms242417153 - 5 Dec 2023
Cited by 4 | Viewed by 3496
Abstract
Antisense oligodeoxynucleotides (ASOs) have long been used to selectively inhibit or modulate gene expression at the RNA level, and some ASOs are approved for clinical use. However, the practicability of antisense technologies remains limited by the difficulty of reliably predicting the sites accessible [...] Read more.
Antisense oligodeoxynucleotides (ASOs) have long been used to selectively inhibit or modulate gene expression at the RNA level, and some ASOs are approved for clinical use. However, the practicability of antisense technologies remains limited by the difficulty of reliably predicting the sites accessible to ASOs in complex folded RNAs. Recently, we applied a plant-based method that reproduces RNA-induced RNA silencing in vitro to reliably identify sites in target RNAs that are accessible to small interfering RNA (siRNA)-guided Argonaute endonucleases. Here, we show that this method is also suitable for identifying ASOs that are effective in DNA-induced RNA silencing by RNases H. We show that ASOs identified in this way that target a viral genome are comparably effective in protecting plants from infection as siRNAs with the corresponding sequence. The antiviral activity of the ASOs could be further enhanced by chemical modification. This led to two important conclusions: siRNAs and ASOs that can effectively knock down complex RNA molecules can be identified using the same approach, and ASOs optimized in this way could find application in crop protection. The technology developed here could be useful not only for effective RNA silencing in plants but also in other organisms. Full article
Show Figures

Figure 1

Figure 1
<p>DNA-directed cleavage of a target RNA in <span class="html-italic">Nicotiana tabacum</span> BYL. (<b>A</b>) Schematic representation of cleavage assays that were performed in the presence or absence of heterologous AGO protein. RNA-directed cleavage was performed with the siRNA duplex (siR gf698) targeting GFP mRNA or with the corresponding siRNA guide strand (gs) indicated in red. DNA-directed cleavage of the target RNA was performed with an antisense oligonucleotide (ASO gf698) whose sequence corresponds to that of the siRNA guide strand (indicated in green) or, alternatively, with a double-stranded oligonucleotide (ASO duplex) whose sequence corresponds to that of the siRNA duplex (not shown). (<b>B</b>) The assays shown were performed in the presence (<b>left</b>) or absence (right) of <span class="html-italic">N. benthamiana</span> AGO1 protein generated by in vitro translation of the corresponding mRNA (see <b>A</b>) in BYL and with different concentrations of the added nucleic acids (0.1 and 1 µM, respectively). Cleavage of the <sup>32</sup>P-labeled template was monitored by denaturing PAGE [<a href="#B30-ijms-24-17153" class="html-bibr">30</a>]; the cleavage products (cp) are indicated.</p>
Full article ">Figure 2
<p>Characteristics of the DNA-directed hydrolysis activity in <span class="html-italic">Nicotiana tabacum</span> BYL. (<b>A</b>) The cleavage pattern of ASO-directed hydrolysis of target RNA in BYL is similar to the cleavage pattern generated by <span class="html-italic">E. coli</span> RNase H. ASO gf698, and the corresponding target RNA was hybridized and tested in cleavage assays in BYL, or, under analogous buffer conditions, with purified RNase H protein from <span class="html-italic">E. coli</span> (New England Biolabs, Frankfurt, Germany). For comparison, siRNA (siR gf698)-directed cleavage of the same target RNA was performed with reconstituted AGO1/RISC. The analyses were performed in the same way as described in <a href="#ijms-24-17153-f001" class="html-fig">Figure 1</a>; cleavage products (cp) are indicated. (<b>B</b>) DNA-directed cleavage activity in BYL exhibits features of both RNase H1 and RNase H2. Different 27 nt DNA oligonucleotides containing one, two, or four ribonucleotides (highlighted in red) were used as targets for cleavage mediated by a complementary ASO (<b>left</b>). The design of the heteroduplexes was based on the report of Eder et al. [<a href="#B34-ijms-24-17153" class="html-bibr">34</a>]; sequences were adapted to hybridize with ASO gf698 (see <a href="#ijms-24-17153-f001" class="html-fig">Figure 1</a> and <a href="#sec4-ijms-24-17153" class="html-sec">Section 4</a>). The assay was performed as previously described (<b>upper</b> panel); as controls, all target nucleic acids were incubated in BYL in the absence of ASO (<b>lower</b> panel). Red arrows indicate the putative cleavage sites inferred from denaturing PAGE of total RNA isolated from the samples and subsequent autoradiography (<b>right</b>).</p>
Full article ">Figure 3
<p>ASOs derived from <span class="html-italic">e</span>siRNAs targeting the TBSV genome show antiviral activity. (<b>A</b>) ASOs derived from <span class="html-italic">e</span>siRNAs mediate efficient hydrolysis of viral RNA in vitro. Cleavage assays with TBSV-targeting siRNAs siR179, siR3722, and siR3939 that were previously shown to function efficiently in AGO1/RISC and siRNAs siR209, siR3243, and siR3701 that were previously shown to function efficiently in AGO2/RISC [<a href="#B20-ijms-24-17153" class="html-bibr">20</a>]. <span class="html-italic">N. benthamiana</span> AGO1 or AGO2 were generated by translation in BYL in the presence of the associated siRNA duplexes, thus generating RISCs that were programmed with these siRNAs. Subsequently, labeled full-length TBSV genomic RNA was added and analyzed for siRNA-mediated cleavage by denaturing agarose gel and autoradiography (<b>left</b>, <b>middle</b>). ASOs corresponding in sequence to the guide (antisense) strand of the aforementioned siRNAs were used in an analogous assay designed to analyze RNase H-mediated hydrolysis of TBSV RNA (<b>right</b>). The assay was performed in the same manner as described above, except that no AGO mRNAs were added to BYL (see also <a href="#ijms-24-17153-f001" class="html-fig">Figure 1</a>). Asterisks indicate cleavage products. (<b>B</b>) ASOs corresponding in sequence to the guide strand of TBSV-targeting <span class="html-italic">e</span>siRNAs protect plants against TBSV infections. <span class="html-italic">N. benthamiana</span> plants were co-inoculated with genomic TBSV RNA and with TBSV-targeting siRNAs or ASOs using the <span class="html-italic">rub-inoculation</span> procedure (see <a href="#sec4-ijms-24-17153" class="html-sec">Section 4</a>). SiR gf698 and the corresponding ASO gf698 were used as negative controls; siR209 was applied as a positive control. Additional controls involved ASO3701 and ASO3722; their sequences were derived from siRNAs directed against the TBSV genome but were nonfunctional in vitro (see <b>A</b>). The plants were monitored for 21 days post inoculation (dpi) for the appearance of typical symptoms of virus infection (representative pictures on the left). Except for ASO179, minimally, two independent infection experiments were performed for each nucleic acid.</p>
Full article ">Figure 4
<p>Chemical modifications increase the antiviral activity of ASO. (<b>A</b>) Cleavage assays with chemically modified ASOs. Variants of ASO209 with 2′-O-methoxyethyl modifications at the four terminal nucleotides (two at the 5′ end, as well as at the 3′ end, referred to here as MOE4), locked nucleic acid modifications at the two terminal nucleotides (LNA2), phosphorothioate modification at the two terminal phosphodiester bonds (PS2), or at all phosphodiester bonds (PS20) were used to analyze the RNase H-mediated hydrolysis of the 5′ part of TBSV genomic RNA containing the ASO binding site. The ASO and the <sup>32</sup>P-labeled target RNA (encompassing the 5′-terminal 740 nucleotides of the TBSV genome) were combined before being added to BYL. Cleavage was analyzed by denaturing PAGE. Cleavage products (cp) are indicated by arrows. (<b>B</b>) Chemical modifications improve the <span class="html-italic">e</span>ASO-mediated protection of plants against a TBSV infection. <span class="html-italic">N. benthamiana</span> plants were co-inoculated with TBSV genomic RNA and with ASO209 or ASO209 variants by rubbing the nucleic acids onto the leaves. ASO gf698, targeting GFP mRNA, was used as a negative control, and siR209 was applied as a positive control. The plants were monitored for 21 days post-inoculation (dpi) for the appearance of typical symptoms. Two independent infection experiments were performed. (<b>C</b>) Sequential inoculation of ASOs and viral RNA also protects plants from viral infection. <span class="html-italic">N. benthamiana</span> plants were first inoculated with ASO209, chemically modified variants of ASO209 or the aforementioned controls. Three minutes later, TBSV genomic RNA was rubbed onto the same leaves. The plants were monitored for 21 days post-inoculation (dpi) for the appearance of disease symptoms. Two independent infection experiments were performed.</p>
Full article ">
Back to TopTop