[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
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

Journals

Article Types

Countries / Regions

Search Results (30)

Search Parameters:
Keywords = Cajal body

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
18 pages, 4796 KiB  
Article
Mutations of Key Functional Residues in CRM1/XPO1 Differently Alter Its Intranuclear Localization and the Nuclear Export of Endogenous Cargos
by Miren Josu Omaetxebarria, Maria Sendino, Liher Arrizabalaga, Irune Mota, Ana Maria Zubiaga and José Antonio Rodríguez
Biomolecules 2024, 14(12), 1578; https://doi.org/10.3390/biom14121578 - 10 Dec 2024
Viewed by 541
Abstract
CRM1 (XPO1) has been well-characterized as a shuttling receptor that mediates the export of protein and RNA cargos to the cytoplasm, and previous analyses have pinpointed several key residues (A541, F572, K568, S1055, and Q742) that modulate CRM1 export activity. CRM1 also has [...] Read more.
CRM1 (XPO1) has been well-characterized as a shuttling receptor that mediates the export of protein and RNA cargos to the cytoplasm, and previous analyses have pinpointed several key residues (A541, F572, K568, S1055, and Q742) that modulate CRM1 export activity. CRM1 also has a less studied nuclear function in RNA biogenesis, which is reflected by its localization to the Cajal body and the nucleolus. Here, we have investigated how the mutation of these key residues affects the intranuclear localization of CRM1 and its ability to mediate export of endogenous cargos. We identify A541K as a separation-of-function mutant that reveals the independent nature of the Cajal body and nucleolar localizations of CRM1. We also show that the F572A mutation may have strikingly opposite effects on the export of specific cargos. Importantly, and in contrast to previous claims, our findings indicate that S1055 phosphorylation is not generally required for CRM1 function and that the Q742 is not a function-defining residue in human CRM1. Collectively, our findings provide new insights into an understudied aspect of CRM1 biology and highlight several important issues related to CRM1 function and regulation that need to be re-evaluated and addressed in more detail. Full article
(This article belongs to the Collection Feature Papers in 'Biomacromolecules: Proteins')
Show Figures

Figure 1

Figure 1
<p>CRM1 residues mutated in this study and experimental design. (<b>A</b>) Schematic representation of human CRM1 protein showing the position of the key residues and their mutations analyzed in this study. The aspects of CRM1 function and regulation where these residues are reportedly involved (NES binding, RanBP1 binding, acetylation, phosphorylation) are also indicated. (<b>B</b>) Workflow diagram illustrating the experimental design of the study. (*) Some experiments were also carried out in HEK293T cells.</p>
Full article ">Figure 2
<p>An LMB resistant, YFP-tagged version of CRM1 recapitulates the intranuclear localization of endogenous CRM1 in HeLa cells. (<b>A</b>) Confocal images showing representative examples of the localization of endogenous CRM1 (upper set of panels), YFP-CRM1 (middle set of panels) and LMB-resistant YFP-CRM1* (lower set of panels) in HeLa cells untreated (UT) or treated with the indicated drugs for 3 h. LMB was used at 6 ng/mL and ActD at 100 ng/mL. (<b>B</b>) Confocal images showing representative examples of the co-localization of YFP-CRM1 and YFP-CRM1* with the Cajal body (CB) marker coilin in the nucleus of HeLa cells untreated (UT) or treated with LMB. Zoom images show magnification of one selected CB (white square). (<b>C</b>) Confocal images showing representative examples of the localization of YFP-CRM1, YFP-CRM1*, and endogenous NMD3 in the nucleus of HeLa cells. YFP-CRM1 and YFP-CRM1* co-localize with NMD3 in the nucleoli of ActD-treated cells but not in untreated (UT) cells. Zoom images show magnification of one selected nucleolus (white square). In all the panels DAPI was used to stain the nuclei, and the scale bar represents 10 μm.</p>
Full article ">Figure 3
<p>CRM1 mutations differently alter the localization of the receptor to the Cajal body in HeLa cells. (<b>A</b>) Confocal images showing representative examples of the co-localization of the different YFP-CRM1* mutants with the CB marker coilin in the nucleus of untreated HeLa cells. “Wider field” panels are included to show a larger area of the cell cytoplasm. Arrowheads indicate the nucleus depicted in more detail in the panels below. Zoom images show magnification of one selected CB (white square). DAPI was used to stain the nuclei, and the scale bar represents 10 μm. (<b>B</b>) Graph showing the percentage of HeLa cells with CB localization of each YFP-CRM1* variant. Bars represent the mean of five independent experiments and error bars indicate standard deviation (SD). Student’s <span class="html-italic">t</span> test was used to compare each mutant to the WT. n.s.: non-significant; (**) <span class="html-italic">p</span> &lt; 0.01. (<b>C</b>) Results of a representative experiment where the fluorescence intensity of each YFP-CRM1* variant at the CB and the nucleoplasm was quantified by image analysis using Fiji. The graph shows the CB/nucleoplasm intensity ratio. Each dot represents a single cell and the mean (+/− SD) is also shown. Student’s <span class="html-italic">t</span> test was used to compare each mutant to the WT. n.s.: non-significant; (*) <span class="html-italic">p</span> &lt; 0.05; (****) <span class="html-italic">p</span> &lt; 0.0001. (<b>D</b>) Table summarizing the normalized Cajal body score (Normalized CBscore) for each variant. This score was calculated from five independent experiments (at least 25 cells per condition were scored in each experiment), by multiplying the mean percentage of cells with CB localization by the mean CB/nucleoplasm fluorescence intensity ratio (see <a href="#app1-biomolecules-14-01578" class="html-app">Supplementary Table S1</a>). The score of each mutant was normalized to the score of wild type YFP-CRM1*, set at 100.</p>
Full article ">Figure 4
<p>CRM1 mutations differently alter the relocation of the receptor to the nucleolus in actinomycin D-treated HeLa cells. (<b>A</b>) Confocal images showing representative examples of the co-localization of the different YFP-CRM1* mutants with endogenous NMD3 in the nucleus of HeLa cells treated with ActD (100 ng/mL for 3 h). “Wider field” panels are included to show a larger area of the cell cytoplasm. Zoom images show magnification of one selected nucleolus (white square). DAPI was used to stain the nuclei, and the scale bar represents 10 μm. (<b>B</b>) Graph showing the percentage of HeLa cells with nucleolar localization of each YFP-CRM1* variant. Bars represent the mean of four independent experiments and error bars indicate standard deviation (SD). Student’s <span class="html-italic">t</span> test was used to compare each mutant to the WT. n.s.: non-significant; (*) <span class="html-italic">p</span> &lt; 0.05. (<b>C</b>) Results of a representative experiment where the fluorescence intensity of each YFP-CRM1* variant at the nucleolus and the nucleoplasm was quantified by image analysis using Fiji. The graph shows the nucleolus/nucleoplasm intensity ratio. Each dot represents a single cell and the mean (+/− SD) is also shown. Student’s <span class="html-italic">t</span> test was used to compare each mutant to the WT. n.a: not assessed; n.s.: non-significant; (***) <span class="html-italic">p</span> &lt; 0.001; (****) <span class="html-italic">p</span> &lt; 0.0001. (<b>D</b>) Table summarizing the normalized nucleolar relocation score (Normalized NOLscore) for each variant. This score was calculated from four independent experiments (at least 25 cells per condition were scored in each experiment) by multiplying the mean percentage of cells with nucleolar localization by the mean nucleolar/nucleoplasm fluorescence intensity ratio (see <a href="#app1-biomolecules-14-01578" class="html-app">Supplementary Table S1</a>). The score of each mutant was normalized to the score of wild type YFP-CRM1*, set at 100.</p>
Full article ">Figure 5
<p>Selection of endogenous cargos as markers to evaluate the nuclear export activity of CRM1 mutants. (<b>A</b>) Confocal images showing representative examples of the localization of four endogenous CRM1 cargos (NMD3, RanBP1, SQSTM1, and p65) in HeLa cells untreated (UT) or treated with LMB (6 ng/mL for 3 h). (<b>B</b>) Confocal images showing representative examples of the localization of these cargos in untransfected HeLa cells untreated (UT) or treated with LMB, as well as in LMB-treated transfected cells expressing low to moderate levels of YFP-CRM1*. DAPI was used to stain the nuclei, and the scale bar represents 10 μm.</p>
Full article ">Figure 6
<p>CRM1 mutations differently alter nuclear export of RanBP1. (<b>A</b>) Confocal images showing representative examples of the localization of endogenous RanBP1 in LMB-treated HeLa cells expressing low to moderate levels of each YFP-CRM1* variant. DAPI was used to stain the nuclei, and the scale bar represents 10 μm. (<b>B</b>) Graphs showing the nuclear to cytoplasmic (N/C) ratio of RanBP1 (upper section) and the intensity of the YFP fluorescence (lower section) in HeLa cells expressing the indicated YFP-CRM1* mutant. (<b>C</b>) Graphs showing the nuclear to total (N/total) ratio of RanBP1 (upper section) and the intensity of the YFP fluorescence (lower section) in HEK293T cells expressing the indicated YFP-CRM1* mutant. In B and C panels, each dot represents a single cell where both RanBP1 and YFP-CRM1* fluorescence intensity were determined by image analysis using Fiji. The horizontal lines indicate the mean, and error bars represent SD. Student’s <span class="html-italic">t</span> test was used to compare each mutant to the WT. n.s.: non-significant; (*) <span class="html-italic">p</span> &lt; 0.05; (***) <span class="html-italic">p</span> &lt; 0.001; (****) <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 7
<p>CRM1 mutations differently alter the nuclear export of p65. (<b>A</b>) Confocal images showing representative examples of the localization of endogenous p65 in LMB-treated HeLa cells expressing low to moderate levels of each YFP-CRM1* variant. DAPI was used to stain the nuclei, and the scale bar represents 10 μm. (<b>B</b>) Graphs showing the nuclear to cytoplasmic (N/C) ratio of p65 (upper section) and the intensity of the YFP fluorescence (lower section) in HeLa cells expressing the indicated YFP-CRM1* mutant. (<b>C</b>) Graphs showing the nuclear to total (N/total) ratio of p65 (upper section) and the intensity of the YFP fluorescence (lower section) in HEK293T cells expressing the indicated YFP-CRM1* mutant. In B and C panels, each dot represents a single cell where both p65 and YFP-CRM1* fluorescence intensity were determined by image analysis using Fiji. The horizontal lines indicate the mean, and error bars represent SD. Student’s <span class="html-italic">t</span> test was used to compare each mutant to the WT. n.s.: non-significant; (**) <span class="html-italic">p</span> &lt; 0.01; (***) <span class="html-italic">p</span> &lt; 0.001; (****) <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">
17 pages, 2545 KiB  
Article
miR-30c-5p Gain and Loss of Function Modulate Sciatic Nerve Injury-Induced Nucleolar Stress Response in Dorsal Root Ganglia Neurons
by Raquel Francés, Jorge Mata-Garrido, Miguel Lafarga, María A. Hurlé and Mónica Tramullas
Int. J. Mol. Sci. 2024, 25(21), 11427; https://doi.org/10.3390/ijms252111427 - 24 Oct 2024
Viewed by 2394
Abstract
Neuropathic pain is a prevalent and debilitating chronic syndrome that is often resistant to treatment. It frequently arises as a consequence of damage to first-order nociceptive neurons in the lumbar dorsal root ganglia (DRG), with chromatolysis being the primary neuropathological response following sciatic [...] Read more.
Neuropathic pain is a prevalent and debilitating chronic syndrome that is often resistant to treatment. It frequently arises as a consequence of damage to first-order nociceptive neurons in the lumbar dorsal root ganglia (DRG), with chromatolysis being the primary neuropathological response following sciatic nerve injury (SNI). Nevertheless, the function of miRNAs in modulating this chromatolytic response in the context of neuropathic pain remains unexplored. Our previous research demonstrated that the intracisternal administration of a miR-30c mimic accelerates the development of neuropathic pain, whereas the inhibition of miR-30c prevents pain onset and reverses established allodynia. In the present study, we sought to elucidate the role of miR-30c-5p in the pathogenesis of neuropathic pain, with a particular focus on its impact on DRG neurons following SNI. The organisation and ultrastructural changes in DRG neurons, particularly in the protein synthesis machinery, nucleolus, and Cajal bodies (CBs), were analysed. The results demonstrated that the administration of a miR-30c-5p mimic exacerbates chromatolytic damage and nucleolar stress and induces CB depletion in DRG neurons following SNI, whereas the administration of a miR-30c-5p inhibitor alleviates these effects. We proposed that three essential cellular responses—nucleolar stress, CB depletion, and chromatolysis—are the pathological mechanisms in stressed DRG neurons underlying neuropathic pain. Moreover, miR-30c-5p inhibition has a neuroprotective effect by reducing the stress response in DRG neurons, which supports its potential as a therapeutic target for neuropathic pain management. This study emphasises the importance of miR-30c-5p in neuropathic pain pathogenesis and supports further exploration of miRNA-based treatments. Full article
(This article belongs to the Special Issue Molecular Mechanisms of mRNA Transcriptional Regulation: 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>miR-30c-5p modulation effects on the chromatolysis developed by dorsal root ganglia neurons after spared nerve injury. (<b>A</b>–<b>H</b>) Dissociated dorsal root ganglia (DRG) neurons double stained with propidium iodide (PI, red) and Lamin B1 (green). Note the prominent NBs and round nuclei in sham rats treated with vehicle (<b>A</b>,<b>E</b>), miR-30c-5p mimic (<b>B</b>), or miR-30c-5p inhibitor (<b>F</b>), reflecting a normal distribution of the protein synthesis machinery and nuclear location. DRG neurons from day-5 (<b>C</b>) or day-10 SNI rats (<b>G</b>) exhibited central chromatolysis with dispersion and severe loss of NBs in the centre of the neuronal body, accumulations of Nissl substance at the marginal cytoplasm, and peripheral displacement of the nucleus, which were aggravated by treatment with miR-30c-5p mimic (<b>D</b>). Administration of miR-30c-5p inhibitor reduced the chromatolytic response observed after SNI (<b>H</b>). (<b>I</b>,<b>J</b>) Percentage of neurons showing chromatolysis. (<b>K</b>,<b>L</b>) Percentage of neurons showing eccentricity of the nucleus. The percentage of damaged neurons and eccentric nuclei was determined in 1000 neurons per rat (n = 3 rats per group). ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 vs. Sham; # <span class="html-italic">p</span> &lt; 0.05, ## <span class="html-italic">p</span> &lt; 0.01, ### <span class="html-italic">p</span> &lt; 0.001 vs. SNI (Two-way ANOVA followed by the Bonferroni post hoc test). Scale bar: 5 µm.</p>
Full article ">Figure 2
<p>Electron micrographs illustrating the ultrastructural characteristics of dorsal root ganglia neurons after administration of miR-30c-5p mimic or inhibitor to SNI rats. In dorsal root ganglia (DRG) neurons from sham (<b>A</b>) and SNI rats treated with miR-30c-5p inhibitor (<b>C</b>), the most prominent organelles are the NBs, composed of RER cisterns (<b>C</b>, arrow) and rosettes of free polyribosomes (<b>A</b>, arrow). Bundles of neurofilaments (NF) interspersed between NBs, profiles of Golgi complexes, and mitochondria are also apparent. In DRGs from SNI rats treated with vehicle (<b>B</b>) or miR-30c-5p mimic (<b>D</b>), the NBs disaggregated, leaving an extensive cleared chromatolytic area in the centre of the cell body, free of NBs. The increased number of NFs and the abundance of mitochondria (M)—some of which are very small (&lt;0.5 µm)—in chromatolytic areas are also noteworthy. Scale bar: 5 µm.</p>
Full article ">Figure 3
<p>miR-30c-5p modulation effects on the nucleolar organisation of dorsal root ganglion neurons after spared nerve injury. (<b>A</b>–<b>H</b>) Dissociated dorsal root ganglia (DRG) neurons double immunostained for upstream binding factor (UBF, green) and Lamin B1 (red). DRG neurons from sham rats treated with vehicle (<b>A</b>,<b>E</b>), miR-30c-5p mimic (<b>B</b>), or miR-30c-5p inhibitor (<b>F</b>), and day-10 SNI rats treated with miR-30c-5p inhibitor (<b>H</b>) presented a normal UBF distribution as small dots corresponding to FCs. In contrast, DRG neurons from day-5 SNI rats treated with vehicle (<b>C</b>) or miR-30c-5p mimic (<b>D</b>) and day-10 SNI rats treated with vehicle (<b>G</b>) showed segregation of UBF nucleolar staining into one or a few giant FCs. (<b>I</b>,<b>J</b>) The percentage of neurons showing UBF-positive giant FCs was determined in 1000 neurons per rat (n = 3 rats per group); ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 vs. Sham; ### <span class="html-italic">p</span> &lt; 0.001 vs. SNI (Two-way ANOVA followed by the Bonferroni post hoc test). Scale bar: 5 µm.</p>
Full article ">Figure 4
<p>Representative electron micrographs illustrating the ultrastructural nucleolar characteristics of dorsal root ganglia neurons after administration of miR-30c-5p mimic or inhibitor to SNI rats. Sham (<b>A</b>) and SNI rats treated with miR-30c-5p inhibitor (<b>C</b>) exhibit the typical nucleolar organisation of DRG neurons, characterised by the presence of numerous small-sized fibrillar centres (*, FCs), surrounded by a ring of dense fibrillar component (DFC), and areas of granular component (GC), preferentially at the nucleolar periphery. SNI rats treated with vehicle (<b>B</b>) or with miR-30c-5p mimic (<b>D</b>) present severe nucleolar alterations, including the formation of enlarged FCs and segregation of large masses of GC and DFC at the nucleolar periphery. Scale bar: 2 µm.</p>
Full article ">Figure 5
<p>miR-30c-5p modulation effects on the number of Cajal bodies in dorsal root ganglion neurons after spared nerve injury. Representative images of dissociated DRG neurons immunolabeled for coilin (green) and counterstained with propidium iodide ((PI), red). Example of neurons showing 0 (<b>A</b>), 1 (<b>B</b>), and 2 (<b>C</b>) CBs. (<b>D</b>,<b>E</b>) Quantitative analysis of the percentage of neurons carrying 0, 1, or more than 2 CBs in each of our experimental groups. The number of CBs per neuron was determined in 1000 neurons per rat, in 3 rats of each group (sham; SNI + vehicle; SNI + miR-30c-5p inhibitor; SNI + miR-30c-5p mimic). The quantification analysis indicates that, regardless of the experimental condition, most neurons present 1 CB. There is a significant increase in the percentage of neurons showing more than 2 CBs in SNI rats treated with miR-30c-5p inhibitor. The proportion of neurons without CBs is significantly increased in SNI rats treated with vehicle or miR-30c-5p mimic. (<b>F</b>,<b>G</b>) Electron microscopy of CBs in DRG neurons from SNI rats treated with miR-30c-5p inhibitor showing 3 CBs (<b>F</b>) and a hypertrophic CB physically close to the nucleolus (<b>G</b>). * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.01 vs. Sham; ### <span class="html-italic">p</span> &lt; 0.001 vs. SNI). (Two-way ANOVA followed by the Bonferroni post hoc test). Scale bar: 5 µm. Scale bar: 2 µm.</p>
Full article ">
26 pages, 3701 KiB  
Review
Modulatory Impact of Oxidative Stress on Action Potentials in Pathophysiological States: A Comprehensive Review
by Chitaranjan Mahapatra, Ravindra Thakkar and Ravinder Kumar
Antioxidants 2024, 13(10), 1172; https://doi.org/10.3390/antiox13101172 - 26 Sep 2024
Viewed by 2678
Abstract
Oxidative stress, characterized by an imbalance between the production of reactive oxygen species (ROS) and the body’s antioxidant defenses, significantly affects cellular function and viability. It plays a pivotal role in modulating membrane potentials, particularly action potentials (APs), essential for properly functioning excitable [...] Read more.
Oxidative stress, characterized by an imbalance between the production of reactive oxygen species (ROS) and the body’s antioxidant defenses, significantly affects cellular function and viability. It plays a pivotal role in modulating membrane potentials, particularly action potentials (APs), essential for properly functioning excitable cells such as neurons, smooth muscles, pancreatic beta cells, and myocytes. The interaction between oxidative stress and AP dynamics is crucial for understanding the pathophysiology of various conditions, including neurodegenerative diseases, cardiac arrhythmias, and ischemia-reperfusion injuries. This review explores how oxidative stress influences APs, focusing on alterations in ion channel biophysics, gap junction, calcium dynamics, mitochondria, and Interstitial Cells of Cajal functions. By integrating current research, we aim to elucidate how oxidative stress contributes to disease progression and discuss potential therapeutic interventions targeting this interaction. Full article
(This article belongs to the Special Issue Novel Antioxidant Mechanisms for Health and Diseases)
Show Figures

Figure 1

Figure 1
<p>Illustration of the processes by which oxidative stress disrupts normal cells through the induction of reactive oxygen species. The red positive signs indicate ROS enhancers, while the red negative signs indicate ROS inhibitors. Oxidative stress radicals that modulate the shape of the cell are depicted by red stars and circles.</p>
Full article ">Figure 2
<p>(<b>a</b>) Illustration of the simulated membrane depolarization (black solid line), AP (red solid line), depolarization phase, repolarization phase, threshold potential (star mark), and resting membrane potential, which is maintained at −52 mV. (<b>b</b>–<b>d</b>) show simulated cardiac AP, slow wave with a burst, and series of neuronal Aps, respectively. The <span class="html-italic">X</span>-axis represents unscaled time, while the <span class="html-italic">Y</span>-axis represents unscaled membrane potential.</p>
Full article ">Figure 3
<p>(<b>a</b>) shows how connexins from Cell 1 and Cell 2 form a gap junction, enabling signal transfer between them, as indicated by the red bidirectional arrow. (<b>b</b>) depicts six cells connected in a linear arrangement through gap junctions (red arrow), illustrating signal transmission along this network. (<b>c</b>) is a schematic of the gap junction between Cell 1 and Cell 2, where V<sub>1</sub> and V<sub>2</sub> represent their membrane potentials, and r<sub>j</sub> indicates the gap junction resistance.</p>
Full article ">Figure 4
<p>A schematic representation of the ICC cell among smooth muscle cells (SM cells). An ICC cell consists of several ion channels and is connected to neighboring cells via a gap junction.</p>
Full article ">Figure 5
<p>Illustrates Ca<sup>2+</sup> dynamics processes with all cellular and sub-cellular compartments described in the previous paragraph.</p>
Full article ">Figure 6
<p>Illustrates protein-mediating ion fluxes in the outer mitochondria membrane described in the previous paragraph. The putative channel acetylcholine receptor is also illustrated.</p>
Full article ">Figure 7
<p>A schematic diagram of the representation of the redox modulation on membrane potential via several pathways that can modulate the AP parameter and cellular excitability.</p>
Full article ">
22 pages, 1655 KiB  
Review
The Nucleolus and Its Interactions with Viral Proteins Required for Successful Infection
by José Manuel Ulloa-Aguilar, Luis Herrera Moro Huitron, Rocío Yazmin Benítez-Zeferino, Jorge Francisco Cerna-Cortes, Julio García-Cordero, Guadalupe León-Reyes, Edgar Rodrigo Guzman-Bautista, Carlos Noe Farfan-Morales, José Manuel Reyes-Ruiz, Roxana U. Miranda-Labra, Luis Adrián De Jesús-González and Moises León-Juárez
Cells 2024, 13(18), 1591; https://doi.org/10.3390/cells13181591 - 21 Sep 2024
Viewed by 1773
Abstract
Nuclear bodies are structures in eukaryotic cells that lack a plasma membrane and are considered protein condensates, DNA, or RNA molecules. Known nuclear bodies include the nucleolus, Cajal bodies, and promyelocytic leukemia nuclear bodies. These bodies are involved in the concentration, exclusion, sequestration, [...] Read more.
Nuclear bodies are structures in eukaryotic cells that lack a plasma membrane and are considered protein condensates, DNA, or RNA molecules. Known nuclear bodies include the nucleolus, Cajal bodies, and promyelocytic leukemia nuclear bodies. These bodies are involved in the concentration, exclusion, sequestration, assembly, modification, and recycling of specific components involved in the regulation of ribosome biogenesis, RNA transcription, and RNA processing. Additionally, nuclear bodies have been shown to participate in cellular processes such as the regulation of transcription of the cell cycle, mitosis, apoptosis, and the cellular stress response. The dynamics and functions of these bodies depend on the state of the cell. It is now known that both DNA and RNA viruses can direct their proteins to nuclear bodies, causing alterations in their composition, dynamics, and functions. Although many of these mechanisms are still under investigation, it is well known that the interaction between viral and nuclear body proteins is necessary for the success of the viral infection cycle. In this review, we concisely describe the interaction between viral and nuclear body proteins. Furthermore, we focus on the role of the nucleolus in RNA virus infections. Finally, we discuss the possible implications of the interaction of viral proteins on cellular transcription and the formation/degradation of non-coding RNAs. Full article
Show Figures

Figure 1

Figure 1
<p>Subnuclear structures and their interaction with viruses.</p>
Full article ">Figure 2
<p>The ole of non-coding RNAs during viral infections. (<b>A</b>) RNA molecules play various roles in cells, and they are broadly classified into two categories: coding RNA and non-coding RNA. Non-coding RNAs do not code for proteins but have crucial roles in regulating gene expression and maintaining cellular functions. Here is an overview of the main types of ncRNA: microRNA (miRNA), small interfering RNA (siRNA), piwi-Interacting RNA (piRNA), small nuclear RNA (snRNA), small nucleolar RNA (snoRNA), long non-coding RNA (lncRNA), circular RNA (circRNA), and long intergenic non-coding RNA (lincRNA). (<b>B</b>) miRNAs are small, non-coding, single-stranded RNAs ~23 nt (ranging from 19 to 25 nt). The majority of mammalian miRNAs genes are located in intergenic regions or in antisense orientation and are transcribed by RNA polymerase II (Pol II) as primary miRNA transcripts (pri-miRNAs). (<b>C</b>) pri-miRNAs are capped, polyadenylated, and contain a local stem–loop structure that encodes miRNA sequences in the arm of the stem. This stem–loop structure is cleaved by the nuclear RNase III type enzyme Drosha in a process known as ‘cropping’. In the nucleus, the RNA hairpin structure is excised by the RNAse III-like enzyme Drosha and its co-factor DGCR8 to form the precursor miRNA (pre-miRNA). (<b>D</b>) pre-miRNA is translocated to the cytosol by exportin5, where it is processed by the Dicer protein complex, resulting in an miRNA duplex (miRNA/miRNA*), which is made up of a guide chain (miRNA) and a passenger chain (miR-NA*). (<b>E</b>) The miRNA/miRNA* is then loaded into the Argonaute (AGO), promoting the expulsion and degradation of the miRNA and the formation of the RNA-induced silencing complex (RISC). The RISC recognizes the targeted mRNA through base-pairing with miRNA. (<b>F</b>) miRNAs function as key regulators of gene expression in many different cellular pathways and systems, including immune response. So, several viruses with the purpose of carrying out an efficient replication or a persistent infection are able to modify their biogenesis, such as in the case of HIV. In this sense, several studies report an increase in miRNAs that facilitate its replication while inhibiting the Dicer–TRBP–PACT complex.</p>
Full article ">Figure 3
<p>(<b>A</b>) Atorvastatin and Ivermectin: Both drugs are known to disrupt the nuclear–cytoplasmic transport of proteins. Specifically, they impair the trafficking of viral proteins, which is crucial for the assembly and maturation of viral particles. This action has been noted in viruses like Dengue and Zika virus (ZIKV), where effective viral replication relies on the proper localization of viral proteins within the host cell. (<b>B</b>) 5-Fluorouracil (5-FU): A pyrimidine analog that interferes with nucleotide metabolism and RNA function. 5-FU targets nucleolar structures, disrupting the organization and function of the nucleolus. This disruption impairs the transport and processing of rRNA and other molecules, crucial for ribosome assembly and function. (<b>C</b>) Quarfloxin (CX-3543): Its main mechanism is the inhibition of the interaction between the nucleolar protein nucleophosmin (NPM1) and DNA containing regions rich in G-quadruplexes, secondary structures present in promoter regions of ribosomal DNA. This drug destabilizes ribosome assembly by blocking the transcription of ribosomal RNA, which reduces protein production in the cell. (<b>D</b>) Leptomycin B: Inhibits CRM1 (also known as exportin 1), a key protein in the nuclear export of proteins and RNAs. By blocking the nuclear export of viral proteins and RNAs, Leptomycin B effectively prevents the replication of various viruses, including HIV and Influenza. This inhibition disrupts the life cycle of these viruses, which rely on the export of viral components for replication and assembly. Selinexor: Another inhibitor of CRM1/exportin 1, like Leptomycin B. Used in the treatment of certain cancers and viral infections, Selinexor blocks the nuclear export of viral and cellular components, thereby disrupting viral replication and cancer cell proliferation by affecting cellular stress responses and apoptotic pathways. (<b>E</b>) Cisplatin: Forms covalent adducts with DNA, including ribosomal DNA (rDNA), and proteins within the nucleolus. These adducts create steric hindrances that prevent the proper assembly and function of nucleolar components. This action blocks the synthesis and maturation of rRNA, thereby hindering viral access to the nucleolar machinery necessary for replication. (<b>F</b>) CDK inhibitors: Target cyclin-dependent kinases (CDKs), which are critical regulators of cell cycle progression and nucleolar function. These inhibitors disrupt the nucleolar scaffold, leading to nucleolar dissolution. This disruption affects rRNA transcription and processing, impairing the nucleolus’s ability to produce ribosomes, which are necessary for protein synthesis, including viral proteins. (<b>G</b>) Camptothecin and Doxorubicin: Inhibit RNA polymerase I (Pol I), which is responsible for the transcription of rRNA genes. These drugs reduce the synthesis of rRNA, leading to decreased ribosome production. Since ribosomes are essential for the translation of viral proteins, their reduced availability impairs viral replication. BMH-21: Exerts its action by binding to DNA in rRNA gene regions, which leads to inhibition of RNA polymerase I and degradation of the enzyme. This inhibition specifically affects cells with a high rate of rRNA synthesis, such as tumor cells, without severely impacting normal cells.</p>
Full article ">
18 pages, 12284 KiB  
Article
Defense Responses Induced by Viral Movement Protein and Its Nuclear Localization Modulate Virus Cell-to-Cell Transport
by Anastasia K. Atabekova, Ekaterina A. Lazareva, Alexander A. Lezzhov, Sergei A. Golyshev, Boris I. Skulachev, Sergey Y. Morozov and Andrey G. Solovyev
Plants 2024, 13(18), 2550; https://doi.org/10.3390/plants13182550 - 11 Sep 2024
Viewed by 985
Abstract
Movement proteins (MPs) encoded by plant viruses are essential for cell-to-cell transport of viral genomes through plasmodesmata. The genome of hibiscus green spot virus contains a module of two MP genes termed ‘binary movement block’ (BMB), encoding the proteins BMB1 and BMB2. Here, [...] Read more.
Movement proteins (MPs) encoded by plant viruses are essential for cell-to-cell transport of viral genomes through plasmodesmata. The genome of hibiscus green spot virus contains a module of two MP genes termed ‘binary movement block’ (BMB), encoding the proteins BMB1 and BMB2. Here, BMB1 is shown to induce a defense response in Nicotiana benthamiana plants that inhibits BMB-dependent virus transport. This response is characterized by the accumulation of reactive oxygen species, callose deposition in the cell wall, and upregulation of 9-LOX expression. However, the BMB1-induced response is inhibited by coexpression with BMB2. Furthermore, BMB1 is found to localize to subnuclear structures, in particular to Cajal bodies, in addition to the cytoplasm. As shown in experiments with a BMB1 mutant, the localization of BMB1 to nuclear substructures enhances BMB-dependent virus transport. Thus, the virus transport mediated by BMB proteins is modulated by (i) a BMB1-induced defense response that inhibits transport, (ii) suppression of the BMB1-induced response by BMB2, and (iii) the nuclear localization of BMB1 that promotes virus transport. Collectively, the data presented demonstrate multiple levels of interactions between viral pathogens and their plant hosts during virus cell-to-cell transport. Full article
(This article belongs to the Section Plant Protection and Biotic Interactions)
Show Figures

Figure 1

Figure 1
<p>BMB1 induces a defense response in <span class="html-italic">N. benthamiana</span>. (<b>A</b>) A PCD-like defense response in a leaf region agroinfiltrated for expression of BMB1 and coexpression of BMB1 and BMB2. An empty vector was used as a control. The leaf was imaged at 5 dpi. (<b>B</b>) ROS accumulation in leaves agroinfiltrated for expression of BMB1. DAB staining was carried out at 3 dpi. Scale bars in (<b>A</b>,<b>B</b>), 1 cm. (<b>C</b>,<b>D</b>) Staining of callose depositions with aniline blue in cell walls of leaves agroinfiltrated for expression of empty vector (<b>C</b>) and BMB1 (<b>D</b>) at 2 dpi. Typical images are shown. Scale bar in (<b>C</b>,<b>D</b>), 20 μm. (<b>E</b>) Quantification of callose staining data. Average integrated intensities of signal calculated for individual callose deposition spots are shown; error bars indicate the standard error. More than 500 individual callose deposition spots were measured on five agroinfiltrated leaves to calculate the values shown. Asterisks indicate a statistically significant difference (***, <span class="html-italic">p</span> &lt; 0.001) according to the Student’s <span class="html-italic">t</span>-test. (<b>F</b>) The expression level of 9-LOX in leaves agroinfiltrated for expression of empty vector and BMB1. Samples were collected at 2 dpi. Average expression levels determined by qPCR are shown; error bars indicate the standard error. Ten biological replicates were used to calculate each value shown. The asterisk indicates a statistically significant difference (*, <span class="html-italic">p</span> &lt; 0.5) according to the Student’s <span class="html-italic">t</span>-test.</p>
Full article ">Figure 2
<p>BMB2 suppresses the defense response induced by BMB1. (<b>A</b>,<b>B</b>) The efficiency of virus transport depends on the ratio of BMB1 to BMB2. In the virus transport complementation assay, <span class="html-italic">N. benthamiana</span> leaves were agroinfiltrated for coexpression of PVX-POL-GFP with a combination of BMB1 and BMB2 at a BMB1:BMB2 ratio of either 1:1 ((<b>A</b>,<b>B</b>), left leaf halves), 10:1 ((<b>A</b>), right leaf half), or 1:10 ((<b>B</b>), right leaf half). Leaves were imaged under UV light at 4 dpi. (<b>C</b>) BMB2 suppresses callose deposition induced by BMB1 expression, as determined by quantification of callose staining data obtained for leaves agroinfiltrated for expression of vector, BMB1, BMb1 + BMB2, or BMB1 + BMB2mutN. Average integrated signal intensities calculated for individual callose deposition spots are shown; error bars indicate the standard error. More than 500 individual callose deposition spots were measured on five agroinfiltrated leaves to calculate the values shown. Asterisks indicate a statistically significant difference (***, <span class="html-italic">p</span> &lt; 0.001) according to Student’s <span class="html-italic">t</span>-test. NS, not statistically significant (<span class="html-italic">p</span> &gt; 0.05).</p>
Full article ">Figure 3
<p>Localization of BMB1 to nuclear substructures. (<b>A</b>,<b>B</b>) Coexpression of GFP-BMB1 with mRFP-Fib2. In different individual cells, moderate (<b>A</b>) or high (<b>B</b>) levels of GFP-BMB1 accumulation in the nucleolus were observed upon coexpression with mRFP-Fib2. (<b>C</b>) Coexpression of GFP-BMB1d22 with mRFP-Fib2. (<b>D</b>) Coexpression of GFP-BMB1 with mRFP. (<b>E</b>,<b>F</b>) Treatment with leptomycin B results in the accumulation of GFP-BMB1 in the nucleus (<b>F</b>) compared to an untreated cell expressing GFP-BMB1 (<b>E</b>). In (<b>A</b>–<b>D</b>), the left images represent the GFP channel, the center images represent the mRFP channel, and the right images are a superposition of the images for the GFP and mRFP channels. Arrows point to Cajal bodies. Scale bars: 5 μm in (<b>A</b>–<b>D</b>,<b>F</b>); 10 μm in (<b>E</b>,<b>F</b>).</p>
Full article ">Figure 4
<p>Subcellular localization of BMB1 with artificially added NLS, NES, or FLAG. Confocal images show cells of <span class="html-italic">N. benthamiana</span> leaves agroinfiltrated for coexpression of GFP-BMB1 and mRFP (<b>A</b>), GFP-NLS-BMB1 and mRFP (<b>B</b>), GFP-FLAG-BMB1 and mRFP (<b>C</b>), GFP-NES-BMB1 and mRFP (<b>D</b>), GFP-NES2-BMB1 and mRFP (<b>E</b>), GFP-BMB1 and BMB2-mRFP (<b>F</b>), and GFP-NLS- BMB1 and BMB2-mRFP (<b>G</b>). The arrow in (<b>G</b>) points to the nucleus. The left images represent the GFP channel, the center images represent the mRFP channel, and the right images are a superposition of the images for the GFP and mRFP channels. Scale bars: 10 μm.</p>
Full article ">Figure 5
<p>Influence of NLS and FLAG added to the N-terminus of BMB1 on protein functions. (<b>A</b>) Influence of NLS and FLAG on virus cell-to-cell transport in the complementation assay. <span class="html-italic">N. benthamiana</span> leaves were agroinfiltrated for coexpression of PVX-POL-GFP and BMB2 with either empty vector, BMB1, FLAG-BMB1, or NLS-BMB1 as indicated. The leaf was imaged under UV light at 4 dpi. (<b>B</b>) Quantification of the diameter of infection foci in the complementation experiment. At least 30 loci were measured on four agroinfiltrated leaves to calculate each average value. Error bars indicate the standard error. (<b>C</b>) Quantification of callose deposition staining data for leaf areas agroinfiltrated for expression of empty vector, BMB1, FLAG-BMB1, or NLS-BMB1. More than 500 individual callose deposition spots were measured on five agroinfiltrated leaves to calculate the plotted average values. Error bars indicate the standard error. Differences between all pairs of values plotted in (<b>B</b>,<b>C</b>) are statistically significant (<span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 6
<p>Dysfunctional BMB1 localized to the nucleus enhances virus transport. (<b>A</b>) Competence of BMB1-T71N and NLS-BMB1-T71N in virus transport. In a virus transport complementation assay, <span class="html-italic">N. benthamiana</span> leaves were agroinfiltrated for coexpression of PVX-POL-GFP with BMB2, and either empty vector, BMB1, BMB1-T71N, or NLS-BMB1-T71N. Leaves were imaged under UV light at 4 dpi. (<b>B</b>) Influence of dysfunctional BMB1 derivatives on virus cell-to-cell transport. In a virus transport complementation assay, <span class="html-italic">N. benthamiana</span> leaves were agroinfiltrated for coexpression of PVX-POL-GFP with BMB1, BMB2, and either BMB1-T71N or NLS-BMB1-T71N. Leaves were imaged under UV light at 4 dpi. (<b>C</b>) Quantification of the diameter of infection foci in the complementation experiment. At least 30 loci were measured on four agroinfiltrated leaves to calculate each average value. Error bars indicate the standard error. Differences between all pairs of values plotted in (<b>C</b>) are statistically significant (<span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">
22 pages, 12727 KiB  
Article
Small Cajal Body-Specific RNA12 Promotes Carcinogenesis through Modulating Extracellular Matrix Signaling in Bladder Cancer
by Qinchen Lu, Jiandong Wang, Yuting Tao, Jialing Zhong, Zhao Zhang, Chao Feng, Xi Wang, Tianyu Li, Rongquan He, Qiuyan Wang and Yuanliang Xie
Cancers 2024, 16(3), 483; https://doi.org/10.3390/cancers16030483 - 23 Jan 2024
Cited by 1 | Viewed by 2017
Abstract
Background: Small Cajal body-specific RNAs (scaRNAs) are a specific subset of small nucleolar RNAs (snoRNAs) that have recently emerged as pivotal contributors in diverse physiological and pathological processes. However, their defined roles in carcinogenesis remain largely elusive. This study aims to explore the [...] Read more.
Background: Small Cajal body-specific RNAs (scaRNAs) are a specific subset of small nucleolar RNAs (snoRNAs) that have recently emerged as pivotal contributors in diverse physiological and pathological processes. However, their defined roles in carcinogenesis remain largely elusive. This study aims to explore the potential function and mechanism of SCARNA12 in bladder cancer (BLCA) and to provide a theoretical basis for further investigations into the biological functionalities of scaRNAs. Materials and Methods: TCGA, GEO and GTEx data sets were used to analyze the expression of SCARNA12 and its clinicopathological significance in BLCA. Quantitative real-time PCR (qPCR) and in situ hybridization were applied to validate the expression of SCARNA12 in both BLCA cell lines and tissues. RNA sequencing (RNA-seq) combined with bioinformatics analyses were conducted to reveal the changes in gene expression patterns and functional pathways in BLCA patients with different expressions of SCARNA12 and T24 cell lines upon SCARNA12 knockdown. Single-cell mass cytometry (CyTOF) was then used to evaluate the tumor-related cell cluster affected by SCARNA12. Moreover, SCARNA12 was stably knocked down in T24 and UMUC3 cell lines by lentivirus-mediated CRISPR/Cas9 approach. The biological effects of SCARNA12 on the proliferation, clonogenic, migration, invasion, cell apoptosis, cell cycle, and tumor growth were assessed by in vitro MTT, colony formation, wound healing, transwell, flow cytometry assays, and in vivo nude mice xenograft models, respectively. Finally, a chromatin isolation by RNA purification (ChIRP) experiment was further conducted to delineate the potential mechanisms of SCARNA12 in BLCA. Results: The expression of SCARNA12 was significantly up-regulated in both BLCA tissues and cell lines. RNA-seq data elucidated that SCARAN12 may play a potential role in cell adhesion and extracellular matrix (ECM) related signaling pathways. CyTOF results further showed that an ECM-related cell cluster with vimentin+, CD13+, CD44+, and CD47+ was enriched in BLCA patients with high SCARNA12 expression. Additionally, SCARNA12 knockdown significantly inhibited the proliferation, colony formation, migration, and invasion abilities in T24 and UMUC3 cell lines. SCARNA12 knockdown prompted cell arrest in the G0/G1 and G2/M phase and promoted apoptosis in T24 and UMUC3 cell lines. Furthermore, SCARNA12 knockdown could suppress the in vivo tumor growth in nude mice. A ChIRP experiment further suggested that SCARNA12 may combine transcription factors H2AFZ to modulate the transcription program and then affect BLCA progression. Conclusions: Our study is the first to propose aberrant alteration of SCARNA12 and elucidate its potential oncogenic roles in BLCA via the modulation of ECM signaling. The interaction of SCARNA12 with the transcriptional factor H2AFZ emerges as a key contributor to the carcinogenesis and progression of BLCA. These findings suggest SCARNA12 may serve as a diagnostic biomarker and potential therapeutic target for the treatment of BLCA. Full article
(This article belongs to the Special Issue Genetic and Epigenetic Regulation of Tissue Homeostasis in Cancer)
Show Figures

Figure 1

Figure 1
<p>SCARNA12 expression is upregulated in bladder cancer. (<b>A</b>) Boxplot shows the expression of SCARNA12 across 23 tumor types using RNA sequencing (RNA-seq) data from GEPIA. The <span class="html-italic">X</span>-axis represents tumor types, and the <span class="html-italic">Y</span>-axis represents expression value in log2(TPM + 1). <span class="html-italic">p</span> values are calculated using <span class="html-italic">t</span> tests. (<b>B</b>) Boxplots display the expression of SCARNA12 between bladder cancer (BLCA) tissues and normal tissues in SNORic data sets (396 tumor samples and 16 normal samples). (<b>C</b>) GSE160693&amp;GTEx data sets (52 tumor samples and 9 normal samples). (<b>D</b>) Our own cohort (52 tumor samples and 39 adjacent tissue samples). <span class="html-italic">p</span> values are calculated using <span class="html-italic">t</span> tests. (<b>E</b>) Boxplot displays the expression of SCARNA12 across BLCA cell lines (T24, UMUC3, SW780, and J82) and normal bladder epithelial cell line (SV-HUC-1). <span class="html-italic">p</span> values are calculated using the analysis of variance (ANOVA). (<b>F</b>) qRT-PCR depicts the expression level of SCARNA12 in 26 tumor tissues and 12 normal tissue samples. <span class="html-italic">p</span> values are calculated using <span class="html-italic">t</span> tests. (<b>G</b>) Survival analysis reveals the effect of aberrant SCARNA12 expression on BLCA patients with smoking history. <span class="html-italic">p</span> values are calculated using log-rank tests. (<b>H</b>) In situ hybridization (ISH) verifies the different expression levels of SCARNA12 in BLCA tissues and normal adjacent tissue. (<b>I</b>) Violin plot shows the ISH score of SCARNA12 in 140 tumor tissues and 51 normal tissues. <span class="html-italic">p</span> values are calculated using <span class="html-italic">t</span> tests. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 2
<p>Functional annotations of SCARNA12 in BLCA patients. (<b>A</b>) The volcano plot displays differentially expressed genes (DEGs) in BLCA patients between high and low SCARNA12 expression. (<b>B</b>) Gene Set Enrichment Analysis (GSEA) plot shows the enrichment of hallmark gene sets among up-regulated DEGs in BLCA with high expression of SCARNA12. (<b>C</b>,<b>D</b>) Bubble plots visualize representative Gene Ontology (GO) terms and Kyoto Encyclopedia of Genes and Genomes (KEGG) pathways among up-regulated DEGs in BLCA with high expression of SCARNA12. (<b>E</b>) The correlation heatmap displays the correlation between SCARNA12 and genes related to cell cycle and cell adhesion. <span class="html-italic">p</span> values are calculated using Spearman rank correlation. (<b>F</b>) Boxplots show the expression levels of cell cycle- and cell adhesion-related genes in BLCA with high and low expression of SCARNA12. Sample sizes: <span class="html-italic">n</span> = 26 for each group. <span class="html-italic">p</span> values are calculated using <span class="html-italic">t</span> tests. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 3
<p>Single-cell Mass Cytometry analysis illustrates tumor microenvironment heterogeneity. (<b>A</b>) t-SNE maps display the expression of tumor microenvironment-related markers in BLCA between high and low SCARNA12 expression. Sample sizes: <span class="html-italic">n</span> = 10 for each group. (<b>B</b>) Heatmap illustrates the expression profiles of tumor microenvironment-related markers across 17 cell clusters. The proportion of cells in each cell cluster relative to the total cell count are shown as bar plots on the right side. (<b>C</b>) t-SNE maps display the profiles of cell clusters in BLCA between high and low SCARNA12 expression. (<b>D</b>) Histogram indicates the proportions of 17 cell clusters in each BLCA sample. (<b>E</b>) Volcano plot depicts the differential abundance of cell clusters in BLCA between high and low SCARNA12 expression. (<b>F</b>) Bar plot and boxplot display the proportions of distinct cell clusters and cluster 2 in BLCA between high and low SCARNA12 expression. <span class="html-italic">p</span> values are calculated using <span class="html-italic">t</span> tests. (<b>G</b>) Scatter plot shows the correlation between the abundance of cluster 2 and the expression of stemness-related transcription factors. <span class="html-italic">p</span> values are calculated using Spearman rank correlation. * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 4
<p>Knockdown of SCARNA12 alters cellular functions in BLCA cell line. (<b>A</b>) Knockdown efficiencies of SCARNA12 in T24 and UMUC3 cell lines are assessed by qRT-PCR. <span class="html-italic">p</span> values are calculated using <span class="html-italic">t</span> tests. Three replicate samples are analyzed for each group, and the same sample sizes are used below. (<b>B</b>) Cell proliferation abilities are determined in T24 and UMUC3 WT and SCARNA12-KD cell lines by MTT assays. WT: wild type, KD: knockdown. <span class="html-italic">p</span> values are calculated using repeated-measures ANOVA. (<b>C</b>,<b>D</b>) Clone formation abilities are assessed in T24 and UMUC3 WT and SCARNA12-KD cell lines. Colonies with ≥50 cells and diameter ≥0.5 mm are defined as positive clones. <span class="html-italic">p</span> values are calculated using <span class="html-italic">t</span> tests. (<b>E</b>,<b>F</b>) Scratch wound-healing assays reveal the cell-migration abilities in T24 and UMUC3 WT and SCARNA12-KD cell lines. <span class="html-italic">p</span> values are calculated using <span class="html-italic">t</span> tests. (<b>G</b>,<b>H</b>) Transwell assays indicate migration and invasion abilities in T24 and UMUC3 WT and SCARNA12-KD cell lines. <span class="html-italic">p</span> values are calculated using <span class="html-italic">t</span> tests. ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 5
<p>Knockdown of SCARNA12 alters the capabilities of cellular apoptosis, cell cycle arrest, and tumor growth. (<b>A</b>,<b>E</b>) The effects of SCARNA12 knockdown on cell apoptosis in T24 and UMUC3 cell lines are examined by flow cytometry, and histograms show the percentage of apoptotic cells (<b>C</b>,<b>G</b>). <span class="html-italic">p</span> values are calculated using <span class="html-italic">t</span> tests. Three replicate samples are analyzed for each group, and the same sample sizes are used below. (<b>B</b>,<b>F</b>) The effects of SCARNA12 knockdown on cell cycle arrest in T24 and UMUC3 cell lines are examined by flow cytometry, and histograms show the distribution of cells across different phases of the cell cycle. (<b>D</b>,<b>H</b>) <span class="html-italic">p</span> values are calculated using <span class="html-italic">t</span> tests. (<b>I</b>,<b>J</b>) Photographic images show the xenograft tumors from T24, UMUC3 WT, and SCARNA12-KD groups. Tumor weight and volumes are measured and shown in violin plots. <span class="html-italic">p</span> values are calculated using <span class="html-italic">t</span> tests. ns: no significant; * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 6
<p>Functional enrichment analysis of target genes and potential interacting proteins associated with SCARNA12. (<b>A</b>) PCA plot illustrates the clustering of RNA-seq samples. Three replicate samples are analyzed for each group, and the same sample sizes are used below. (<b>B</b>) Bar plot visualizes representative GO terms among down-regulated DEGs in T24 cell lines following SCARNA12 knockdown. The plot categorizes the enriched terms into Biological Process (BP), Cellular Component (CC), and Molecular Function (MF). (<b>C</b>) Bubble plots visualize representative KEGG pathways among down-regulated DEGs in T24 cell lines following SCARNA12 knockdown. (<b>D</b>) GSEA plot shows the enrichment of extracellular matrix (ECM)-related gene sets among down-regulated DEGs in T24 cell lines following SCARNA12 knockdown. (<b>E</b>) Heatmap shows the expression changes of ECM-related genes in T24 cell lines following SCARNA12 knockdown. (<b>F</b>) Silver staining shows the proteins potentially interacted with SCARNA12. (<b>G</b>) Pie chart shows the components of SCARNA12-interacting proteins. (<b>H</b>) Binding analysis for regulation of transcription (BART) for the down-regulated genes in T24 cell lines following SCARNA12 knockdown. (<b>I</b>) Circle plot shows the GO terms enriched by H2AFZ-related DEGs. (<b>J</b>,<b>K</b>) qRT-PCR experiments evaluate the knockdown efficiency of the transcription factor H2AFZ and the expression levels of SCARNA12 after H2AFZ knockdown. <span class="html-italic">p</span> values are calculated using ANOVA. ns: no significant; * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">
19 pages, 6065 KiB  
Article
Nicotiana benthamiana Methanol-Inducible Gene (MIG) 21 Encodes a Nucleolus-Localized Protein That Stimulates Viral Intercellular Transport and Downregulates Nuclear Import
by Ekaterina V. Sheshukova, Kamila A. Kamarova, Natalia M. Ershova and Tatiana V. Komarova
Plants 2024, 13(2), 279; https://doi.org/10.3390/plants13020279 - 17 Jan 2024
Viewed by 1554
Abstract
The mechanical damage of plant tissues leads to the activation of methanol production and its release into the atmosphere. The gaseous methanol or vapors emitted by the damaged plant induce resistance in neighboring intact plants to bacterial pathogens but create favorable conditions for [...] Read more.
The mechanical damage of plant tissues leads to the activation of methanol production and its release into the atmosphere. The gaseous methanol or vapors emitted by the damaged plant induce resistance in neighboring intact plants to bacterial pathogens but create favorable conditions for viral infection spread. Among the Nicotiana benthamiana methanol-inducible genes (MIGs), most are associated with plant defense and intercellular transport. Here, we characterize NbMIG21, which encodes a 209 aa protein (NbMIG21p) that does not share any homology with annotated proteins. NbMIG21p was demonstrated to contain a nucleolus localization signal (NoLS). Colocalization studies with fibrillarin and coilin, nucleolus and Cajal body marker proteins, revealed that NbMIG21p is distributed among these subnuclear structures. Our results show that recombinant NbMIG21 possesses DNA-binding properties. Similar to a gaseous methanol effect, an increased NbMIG21 expression leads to downregulation of the nuclear import of proteins with nuclear localization signals (NLSs), as was demonstrated with the GFP-NLS model protein. Moreover, upregulated NbMIG21 expression facilitates tobacco mosaic virus (TMV) intercellular transport and reproduction. We identified an NbMIG21 promoter (PrMIG21) and showed that it is methanol sensitive; thus, the induction of NbMIG21 mRNA accumulation occurs at the level of transcription. Our findings suggest that methanol-activated NbMIG21 might participate in creating favorable conditions for viral reproduction and spread. Full article
(This article belongs to the Special Issue Plant Volatile Organic Compounds: Revealing the Hidden Interactions)
Show Figures

Figure 1

Figure 1
<p>Quantitative RT-PCR analysis of <span class="html-italic">NbMIG21</span> mRNA levels in different parts of <span class="html-italic">N. benthamiana</span> plants (<b>A</b>) and in leaves upon methanol treatment (<b>B</b>). (<b>A</b>) Seedlings were grown for 10 days after germination. Leaves (V) were harvested from 3-week-old plants (vegetative stage); leaves (R), stems and flowers were harvested from same plants 5 weeks later (reproductive stage). Seeds were collected from 10-week-old plants. (<b>B</b>) Relative amount of <span class="html-italic">NbMIG21</span> mRNA isolated from leaves (V) of <span class="html-italic">N. benthamiana</span> plants incubated in a sealed desiccator with an elevated methanol (+MeOH) concentration for 18 h. The difference between samples from intact leaves and after methanol treatment is significant at <span class="html-italic">p</span> &lt; 0.001 (Student’s <span class="html-italic">t</span>-test). The levels of expression for both (<b>A</b>) and (<b>B</b>) are normalized to the <span class="html-italic">PP2A</span> gene. The <span class="html-italic">NbMIG21</span> expression level in leaves (V) was taken as 1. The plot represents the mean ± SE of three technical repeats from three biological replicates each.</p>
Full article ">Figure 2
<p>Fluorescent microscopy image of GFP accumulation in epidermal cells of leaves infiltrated with PrMIG21-GFP at 3 dpi. Mixture for infiltration was supplemented with agrobacteria containing plasmid for expression of p19 silencing suppressor of tomato bushy stunt virus.</p>
Full article ">Figure 3
<p>PrMIG21 is methanol sensitive. (<b>A</b>) Schematic representation of experimental workflow, with sample collecting time points indicated. (<b>B</b>) Relative amount of <span class="html-italic">GFP</span> mRNA in leaves of <span class="html-italic">N. benthamiana</span> plants agroinfiltrated with PrMIG21-GFP and treated with gaseous methanol. Amount of <span class="html-italic">GFP</span> mRNA at time point “I” is taken as 1. Student’s <span class="html-italic">t</span>-test was applied to assess statistical significance of difference between control plants and plants incubated with methanol. *: <span class="html-italic">p</span> &lt; 0.05, ***: <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 4
<p>NbMIG21p sequence analysis. (<b>A</b>) NbMIG21p sequence. Five 7-amino-acid long repeats are in bold. 11-amino-acid stretches are underlined: colored double lines mark perfect repeats, single or waived—imperfect. Letters in red—NLS predicted by LOCALIZER (<a href="https://localizer.csiro.au/" target="_blank">https://localizer.csiro.au/</a>, accessed on 10 June 2022); letters in a box—NoLS predicted using NOD <a href="http://www.compbio.dundee.ac.uk/www-nod/index.jsp" target="_blank">http://www.compbio.dundee.ac.uk/www-nod/index.jsp</a> (accessed on 10 June 2022). (<b>B</b>) Multiple alignment of MIG21p homologues from <span class="html-italic">Solanaceae</span> species performed using MEGA 11 software [<a href="#B34-plants-13-00279" class="html-bibr">34</a>] with ClustalW algorithm. Asterisk indicates conservative residues. 01—NbMIG21 (ACY74744.2), 02—<span class="html-italic">N. attenuata</span> hypothetical protein (OIT27799.1), 03—<span class="html-italic">N. attenuata</span> uncharacterized protein (XP_019232815.1), 04—<span class="html-italic">N. sylvestris</span> protein SSUH2 homolog (XP_009768095.1), 05—<span class="html-italic">N. tabacum</span> uncharacterized protein (XP_016494112.1), 06—<span class="html-italic">N. tomentosiformis</span> hornerin-like (XP_009592421.1), 07—<span class="html-italic">N. attenuata</span> uncharacterized protein (XP_019264939.1), 08—<span class="html-italic">D. stramonium</span> hypothetical protein (MCD7459440.1), 09—<span class="html-italic">S. stenotomum</span> uncharacterized protein (XP_049376514.1), 10—<span class="html-italic">S. verrucosum</span> hypothetical protein (WMV44005.1), 11—<span class="html-italic">S. tuberosum</span> hornerin-like (XP_006343871.1), 12—<span class="html-italic">S. commersonii</span> hypothetical protein (KAG5593335.1), 13—<span class="html-italic">S. verrucosum</span> uncharacterized protein (XP_049361851.1), 14—<span class="html-italic">S. lycopersicum</span> hornerin (XP_004245528.1), 15—<span class="html-italic">S. chilense</span> hypothetical protein (TMW85354.1), 16—<span class="html-italic">S. pennellii</span> hornerin-like (XP_015085049.1), 17—<span class="html-italic">S. tuberosum</span> hypothetical protein (CAA12355.1), 18—<span class="html-italic">N. tabacum</span> hornerin-like (XP_016480098.1), and 19—<span class="html-italic">N. sylvestris</span> hornerin-like (XP_009763094.1).</p>
Full article ">Figure 5
<p>NbMIG21p intracellular localization. Images of 35S-GFP:NbMIG21 (<b>A</b>) or 35S-NbMIG21:GFP (<b>B</b>) expressing epidermal cells of <span class="html-italic">N. benthamiana</span> leaves at 3 dpi, obtained using confocal fluorescence microscopy. Projection of several confocal sections (top) superimposed on a bright field image of the same cell (bottom). Bars = 20 μm. 35S: cauliflower mosaic virus 35S promoter; T: 35S terminator of transcription.</p>
Full article ">Figure 6
<p>NbMIG21p co-localizes with fibrillarin and coilin. Schematic representation of genetic constructs encoding RFP-tagged NbMIG21p (<b>left</b>) and GFP-tagged fibrillarin or coilin (<b>top</b>). Fluorescent images of <span class="html-italic">N. benthamiana</span> cells 3 days after co-agroinfiltration with either 35S-NbMIG21:RFP or 35S-RFP:NbMIG21 and 35S-NbFib2:GFP or 35S-NbCoil:GFP. Nucleus is marked with a dashed line. Bar = 5 µm.</p>
Full article ">Figure 7
<p>Mutagenesis of predicted NbMIG21p NoLS. (<b>A</b>) Schematic representation of mutations introduced in potential NbMIG21p NoLS sequence. (<b>B</b>) Fluorescent images of <span class="html-italic">N. benthamiana</span> cells 3 days after co-agroinfiltration with either 35S-NbMIG21:GFP or 35S-NbMIG21<sup>NoLSmut</sup>:GFP and 35S-NbFib2:RFP. Nucleus is marked with a dashed line. Bar = 10 µm.</p>
Full article ">Figure 8
<p>NbMIG21p colocalizes with fibrillarin and coilin. YFP fluorescence analyzed using fluorescent microscopy 3 days after infiltration of <span class="html-italic">N. benthamiana</span> leaves with pairs of agrobacteria containing plasmids for the expression of 35S-NbMIG21:YC and 35S-NbFib2:YN (<b>left</b>), 35S-NbMIG21:YC and 35S-NbCoil:YN (<b>right</b>). For each pair, a fluorescence image and superimposed-on-visible-light image are presented. 35S-NbRGP1:YC is used as a negative control. Nucleus is marked with a dashed line. Bar = 5 µm.</p>
Full article ">Figure 9
<p>Increased expression of <span class="html-italic">NbMIG21</span> interferes with GFP:NLS nuclear import. (<b>A</b>) Representative images of nuclear (left) and nucleocytoplasmic (right) GFP:NLS distribution. Fluorescent image (lower panel) and overlay on bright-field image (upper panel). Bar = 20 µm. (<b>B</b>) Quantification of GFP:NLS subcellular localization 48 h after infiltration with 35S-GFP:NLS and 35S-NbMIG21 or “empty” vector. (<b>C</b>) Number of GFP:NLS-containing cells per square cm.</p>
Full article ">Figure 10
<p>Increased <span class="html-italic">NbMIG21</span> expression stimulates TMV:GFP intercellular transport and reproduction. (<b>A</b>) The percentage of TMV:GFP-expressing foci of different sizes quantified at 5th day after agroinfiltration with TMV:GFP and “empty” vector or 35S-NbMIG21. (<b>B</b>) Relative amount of viral RNA as defined by qRT-PCR. *: <span class="html-italic">p</span> &lt; 0.05; **: <span class="html-italic">p</span> &lt; 0.01 (Student’s <span class="html-italic">t</span>-test).</p>
Full article ">Figure 11
<p>Gel retardation assay of PCR fragments representing promoter regions and 6xHis-NbMIG21. Two amounts of PCR fragments were used—45 and 90 ng—as indicated above each lane. NbMIG21p and control proteins were used in a concentration of 200 ng. Yellow dots indicate retarded NbMIG21p-bound PCR fragments; arrow indicates fully bound PCR fragment of 35S promoter. AtKTI: <span class="html-italic">A. thaliana</span> Kunitz trypsin inhibitor; BG: <span class="html-italic">N. benthamiana</span> beta-1,3-glucanase; and BSA: bovine serum albumin, were used as negative controls.</p>
Full article ">
24 pages, 1018 KiB  
Review
Viruses and Cajal Bodies: A Critical Cellular Target in Virus Infection?
by Lucy Lettin, Bilgi Erbay and G. Eric Blair
Viruses 2023, 15(12), 2311; https://doi.org/10.3390/v15122311 - 25 Nov 2023
Cited by 4 | Viewed by 2141
Abstract
Nuclear bodies (NBs) are dynamic structures present in eukaryotic cell nuclei. They are not bounded by membranes and are often considered biomolecular condensates, defined structurally and functionally by the localisation of core components. Nuclear architecture can be reorganised during normal cellular processes such [...] Read more.
Nuclear bodies (NBs) are dynamic structures present in eukaryotic cell nuclei. They are not bounded by membranes and are often considered biomolecular condensates, defined structurally and functionally by the localisation of core components. Nuclear architecture can be reorganised during normal cellular processes such as the cell cycle as well as in response to cellular stress. Many plant and animal viruses target their proteins to NBs, in some cases triggering their structural disruption and redistribution. Although not all such interactions have been well characterised, subversion of NBs and their functions may form a key part of the life cycle of eukaryotic viruses that require the nucleus for their replication. This review will focus on Cajal bodies (CBs) and the viruses that target them. Since CBs are dynamic structures, other NBs (principally nucleoli and promyelocytic leukaemia, PML and bodies), whose components interact with CBs, will also be considered. As well as providing important insights into key virus–host cell interactions, studies on Cajal and associated NBs may identify novel cellular targets for development of antiviral compounds. Full article
(This article belongs to the Section General Virology)
Show Figures

Figure 1

Figure 1
<p>A schematic representation of the eukaryotic nucleus. Major domains and nuclear bodies (NBs) are labelled. The NBs upon which this review is focussed are nucleoli, Cajal bodies, and PML bodies (marked in red ovals). See text for details.</p>
Full article ">Figure 2
<p>A schematic representation of the domain structure of p80-coilin. The self-interacting domain is located at the N-terminus. NLSa and NLSb refer to the two nuclear localisation signals present in coilin and NoLS refers to the nucleolar localisation signal, all of which have positive charge. The stippled boxes represent acidic serine-rich patches. The RG box domain is located at the C-terminus. The C-terminus of coilin is folded into a Tudor domain. Modified from [<a href="#B63-viruses-15-02311" class="html-bibr">63</a>].</p>
Full article ">
16 pages, 3379 KiB  
Article
The Impact of p70S6 Kinase-Dependent Phosphorylation of Gemin2 in UsnRNP Biogenesis
by Lea Marie Esser, Qiaoping Li, Maximilian Jüdt, Thilo Kähne, Björn Stork, Matthias Grimmler, Sebastian Wesselborg and Christoph Peter
Int. J. Mol. Sci. 2023, 24(21), 15552; https://doi.org/10.3390/ijms242115552 - 25 Oct 2023
Viewed by 1361
Abstract
The survival motor neuron (SMN) complex is a multi-megadalton complex involved in post-transcriptional gene expression in eukaryotes via promotion of the biogenesis of uridine-rich small nuclear ribonucleoproteins (UsnRNPs). The functional center of the complex is formed from the SMN/Gemin2 subunit. By binding the [...] Read more.
The survival motor neuron (SMN) complex is a multi-megadalton complex involved in post-transcriptional gene expression in eukaryotes via promotion of the biogenesis of uridine-rich small nuclear ribonucleoproteins (UsnRNPs). The functional center of the complex is formed from the SMN/Gemin2 subunit. By binding the pentameric ring made up of the Sm proteins SmD1/D2/E/F/G and allowing for their transfer to a uridine-rich short nuclear RNA (UsnRNA), the Gemin2 protein in particular is crucial for the selectivity of the Sm core assembly. It is well established that post-translational modifications control UsnRNP biogenesis. In our work presented here, we emphasize the crucial role of Gemin2, showing that the phospho-status of Gemin2 influences the capacity of the SMN complex to condense in Cajal bodies (CBs) in vivo. Additionally, we define Gemin2 as a novel and particular binding partner and phosphorylation substrate of the mTOR pathway kinase ribosomal protein S6 kinase beta-1 (p70S6K). Experiments using size exclusion chromatography further demonstrated that the Gemin2 protein functions as a connecting element between the 6S complex and the SMN complex. As a result, p70S6K knockdown lowered the number of CBs, which in turn inhibited in vivo UsnRNP synthesis. In summary, these findings reveal a unique regulatory mechanism of UsnRNP biogenesis. Full article
(This article belongs to the Special Issue Protein and Lipid Kinases: Structure and Function)
Show Figures

Figure 1

Figure 1
<p>p70S6 kinase influences UsnRNP biogenesis via binding and phosphorylation of Gemin2, a core component of the human SMN complex.</p>
Full article ">Figure 2
<p>The kinase p70S6 interacts with the core components of the SMN complex in cells. (<b>A</b>) Gemin2 localizes in the cytoplasm and Cajal bodies in HEK293T wt cells. All cells were fixed with 4% paraformaldehyde and permeabilized with Triton X-100 to visualize Gemin2 (magenta). DAPI (blue) was used as a DNA marker, with scale bars of 10 µm. (<b>B</b>) The staining procedure for HEK293T cells was executed as described in (<b>A</b>), to visualize p70S6K (cyan) via specific antibody. The DNA was stained with DAPI (blue), with scale bars of 10 µm. (<b>C</b>) Gemin2 (magenta) co-localizes with p70S6K (cyan) predominantly in the cytoplasm. HEK293T cells were treated as described in (<b>A</b>). DAPI (blue) was used as a DNA marker, with scale bars of 10 µm. (<b>D</b>) Immunopurification (IP) of GFP Vector control (GFP), GFP-Gemin2 wt, and GFP-SMN wt overexpressing cells. Expression of GFP proteins was induced with 0.1 µg/mL doxycycline for 24 h. After cell lysis via douncing, GFP-IP was performed and analyzed with Tris/Glycine-SDS-PAGE and Western blotting, using antibodies against p70S6K, GFP, and core components of the SMN complex. As an input, 25 µg of total protein was loaded. The p70S6K did bind more efficiently to GFP-Gemin2 wt than to GFP-SMN wt. (<b>E</b>) Endogenous immunopurification studies revealed that Gemin2 and p70S6K interact in vivo. S100 extract of HEK293 cells was used for performing an endogenous IP experiment with each 1 µg antibody and 1 mg of total protein lysate. Afterward, the IP samples and an input sample were subjected to Tris/Glycine-SDS-PAGE followed by Western blot analysis.</p>
Full article ">Figure 3
<p>The p70S6 kinase phosphorylates Gemin2 at specific residues. (<b>A</b>) The p70S6K phosphorylates components of the SMN complex. In vitro kinase assay using recombinant active His-tagged p70S6K expressed in Sf21 insect cells and GST-SMN and GST-Gemin2 purified from <span class="html-italic">E. coli</span> as substrate proteins were incubated with 10 µCi [32P]-ATP for 45 min at 30 °C. Samples were separated via Tris/Glycine-SDS-PAGE and after coomassie staining (CS) analyzed by autoradiography (AR). (<b>B</b>) The p70S6K phosphorylates Gemin2 at S81. Kinase assay using His-tagged p70S6K with the substrates GST-Gemin2 wt, GST-Gemin2 SA81, GST-Gemin2 SA166, GST-Gemin2 SD81, and GST-Gemin2 SD166 as recombinant proteins purified from <span class="html-italic">E. coli</span> was performed as described in (<b>A</b>). (<b>C</b>) Pulldown assays using recombinant GST-Gemin2 wild type (wt), GST-Gemin2 SA81, GST-Gemin2 SA166, GST-Gemin2 SD81 and GST-Gemin2 SD166 and GST purified from <span class="html-italic">E. coli</span> were executed in HEK293T cytoplasmic extracts for 2 h at 4 °C. Afterward, samples were analyzed via Tris/Glycine-SDS-PAGE and Western blotting (WB), using antibodies against p70S6 kinase, SMN and SmB/B’ (Y12) and amido black staining (ABS) of the whole membrane was used as a loading control for the recombinant proteins.</p>
Full article ">Figure 4
<p>The phosphorylation status of Gemin2 does not influence the binding of the SMN complex components. <span class="html-italic">(</span><b>A</b>) Expression of newly generated cell lines was tested using Western blot analysis of HEK293T S100 extracts and GFP-immunopurification. IP was performed with S100 extract from GFP-Vector control, GFP-SMN wt, GFP-Gemin2 wt, GFP-Gemin2 A81, GFP-Gemin2 A166, GFP-Gemin2 D81, and GFP-Gemin2 D166 overexpressing cells. Protein expression was induced with 0.1 µg/mL doxycycline for 24 h. After cell lysis, GFP-IP was performed and together with 25 µg of total protein was analyzed via Tris/Glycine-SDS-PAGE and Western blotting, using an antibody against GFP. (<b>B</b>) Immunofluorescence studies of GFP-Vector, GFP-SMN wt, GFP-Gemin2 wt, GFP-Gemin2 A81, GFP-Gemin2 A166, GFP-Gemin2 D81, and GFP-Gemin2 D166 in overexpressing cells. Protein expression was induced as described in (<b>A</b>). Cells were fixed with 4% PFA and permeabilized with Triton X-100 to visualize the GFP overexpressed proteins (magenta). The DNA was stained with DAPI (blue), with scale bars of 20 µm. (<b>C</b>) S100 extracts of inducible Flp-In T-REx 293 cells overexpressing GFP-Vector control (GFP), GFP-Gemin2 wt, GFP-Gemin2 A81, GFP-Gemin2 A166, GFP-Gemin2 D81, and GFP-Gemin2 D166, generated by douncing, were applied to a Superose 6 column. Afterward, fractions were analyzed via Tris/Glycine-SDS-PAGE and immunoblotting using antibodies against Gemin2. Each of the overexpressed GFP and the GFP-Gemin2 variants (green) as well as the endogenous Gemin2 (black) were visualized via this analysis. (<b>D</b>) IP of GFP, GFP-Gemin2 A81, GFP-Gemin2 A166, GFP-Gemin2 D81, and GFP-Gemin2 D166 overexpressing cells. Expression of GFP proteins was induced as described in (<b>A</b>). After cell lysis, GFP-IP was performed with S100 extract and analyzed via Tris/Glycine-SDS-PAGE and Western blotting, using antibodies against GFP and core components of the SMN complex.</p>
Full article ">Figure 5
<p>Gemin2 phospho-status and p70S6 kinase influence UsnRNP biogenesis in vivo. (<b>A</b>) Quantification of Cajal bodies in GFP-Vector control, GFP-SMN wt, GFP-Gemin2 wt, GFP-Gemin2 A81, GFP-Gemin2 A166, GFP-Gemin2 D81, and GFP-Gemin2 D166 overexpressing cells. Protein expression was induced with 0.1 µg/mL doxycycline for 24 h. Cells were fixed with 4% PFA and permeabilized with Triton X-100 to visualize SMN (magenta) and coilin (cyan). DAPI (blue) was used as a DNA marker, with scale bars of 10 µm. (<b>B</b>) Overexpression of GFP-SMN as well as GFP-Gemin2 A166 and GFP-Gemin2 D166 in Flp-In T-Rex cells caused an increase in the number of Cajal bodies compared to the GFP-Vector control and GFP-Gemin2 wt cells. The box in the boxplot diagram represents 5–95% of the data; outliers are shown as stacked rectangles. The <span class="html-italic">p</span>-value, calculated with Prism using an unpaired <span class="html-italic">t</span>-test, was **** <span class="html-italic">p</span> &lt; 0.0001. (<b>C</b>) HEK293T cells were treated with 50 nM p70S6K siRNA or non-targeting control for 72 h. Untreated HEK293T cells, as well as the transfected ones, were fixed with 4% PFA, and Cajal bodies were visualized with antibody staining against coilin (cyan) and SMN (magenta). The DNA was stained with DAPI (blue), with scale bars of 10 µm. (<b>D</b>) Decrease of endogenous p70S6K resulted in a dramatic reduction in the number of Cajal bodies compared to untreated cells and non-target control. Statistics and presentation of the data were performed as described in (<b>B</b>). * for <span class="html-italic">p</span> &lt; 0.05, ** for <span class="html-italic">p</span> &lt; 0.01, and ns means not significant.</p>
Full article ">
20 pages, 8185 KiB  
Article
Plant Poly(ADP-Ribose) Polymerase 1 Is a Potential Mediator of Cross-Talk between the Cajal Body Protein Coilin and Salicylic Acid-Mediated Antiviral Defence
by Nadezhda Spechenkova, Viktoriya O. Samarskaya, Natalya O. Kalinina, Sergey K. Zavriev, S. MacFarlane, Andrew J. Love and Michael Taliansky
Viruses 2023, 15(6), 1282; https://doi.org/10.3390/v15061282 - 30 May 2023
Cited by 4 | Viewed by 2573
Abstract
The nucleolus and Cajal bodies (CBs) are sub-nuclear domains with well-known roles in RNA metabolism and RNA-protein assembly. However, they also participate in other important aspects of cell functioning. This study uncovers a previously unrecognised mechanism by which these bodies and their components [...] Read more.
The nucleolus and Cajal bodies (CBs) are sub-nuclear domains with well-known roles in RNA metabolism and RNA-protein assembly. However, they also participate in other important aspects of cell functioning. This study uncovers a previously unrecognised mechanism by which these bodies and their components regulate host defences against pathogen attack. We show that the CB protein coilin interacts with poly(ADP-ribose) polymerase 1 (PARP1), redistributes it to the nucleolus and modifies its function, and that these events are accompanied by substantial increases in endogenous concentrations of salicylic acid (SA), activation of SA-responsive gene expression and callose deposition leading to the restriction of tobacco rattle virus (TRV) systemic infection. Consistent with this, we also find that treatment with SA subverts the negative effect of the pharmacological PARP inhibitor 3-aminobenzamide (3AB) on plant recovery from TRV infection. Our results suggest that PARP1 could act as a key molecular actuator in the regulatory network which integrates coilin activities as a stress sensor for virus infection and SA-mediated antivirus defence. Full article
(This article belongs to the Special Issue Plant Viruses: Pirates of Cellular Pathways)
Show Figures

Figure 1

Figure 1
<p>Effect of TRV infection on <span class="html-italic">N. benthamiana PARP1</span> (<span class="html-italic">NbPARP1</span>) gene expression at 14 dpi. Accumulation of PARP1mRNA (<b>a</b>) and TRV RNA1 used as an indicator of TRV multiplication [<a href="#B33-viruses-15-01282" class="html-bibr">33</a>] (<b>b</b>) (measured by RT-qPCR) in newly emerged leaves of WT and coilin KD <span class="html-italic">N. benthamiana</span> plants infected with or without (mock-inoculated) TRV or TRV∆16K. PARP1 mRNA and TRV RNA1 expression levels were normalized to those of internal <span class="html-italic">N. benthamiana</span> controls, <span class="html-italic">UBIQUITIN3</span> gene (<span class="html-italic">UBI3)</span> and <span class="html-italic">60S ribosomal protein 23</span> gene (<span class="html-italic">L23</span>) [<a href="#B46-viruses-15-01282" class="html-bibr">46</a>]. Statistical analysis was performed on four independent biological replicates. Data are mean ± SD. Each replicate was composed of samples from three plants pooled together (two leaves per plant). Analysis of variance and Tukey’s HSD post hoc tests were performed on the RT-qPCR data. NS, not significant; ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 2
<p>Interaction of PARP1 with coilin. (<b>a</b>) Co-immunoprecipitation of PARP1, coilin and TRV 16K protein. Protein extracts were prepared at 7 dpi from older systemically infected leaves (third and fourth leaves above the inoculated leaf) of WT or coilin KD plants infected with or without (mock-inoculated, M) TRV or TRV∆16K, as indicated. Proteins in the lysate prior to immunoprecipitation (IP) are shown on the left (input). Anti-coilin antibodies (anti-coilin) were used to co-immunoprecipitate coilin and PARP. Anti-PARP antibodies were used to co-immunoprecipitate PARP and coilin. Antibodies to 16 K (anti-16K) were used to co-precipitate 16K, coilin and PARP. Proteins were detected by western blot analysis (immunoblotting, IB) using anti-PARP and anti-coilin antibodies. Positions of molecular mass markers are on the left. Images have been cropped for presentation. Uncropped images are presented in <a href="#app1-viruses-15-01282" class="html-app">Supplementary Materials</a>. (<b>b</b>) Far-western blot analysis of the in vitro interaction between coilin and commercially-sourced PARP1. Bovine serum albumin (BSA) was used as a negative control. The left blot was stained with Ponceau red; the middle and right blots were incubated with and without the recombinant coilin, respectively (as indicated), and then probed with anti-coilin antiserum. Positions of the molecular mass markers are indicated on the left.</p>
Full article ">Figure 3
<p>The interaction between coilin and PARP1 induced by TRV results in partial nucleolar sequestration of PARP1 and over-accumulation of PARylated proteins. (<b>a</b>) Representative images of intranuclear distribution of PARP1 (immunofluorescent staining using primary rabbit anti-PARP1 antibody and secondary fluorescent anti-rabbit antibody, green ) in WT and coilin KD <span class="html-italic">N. benthamiana</span> plants infected with or without (mock-inoculated) TRV or TRV∆16K were taken at 7 dpi in older systemically infected leaves (third and fourth leaves above the inoculated leaf) or at 14 dpi in recovered newly emerging leaves of WT plants systemically infected with TRV [seventh and eighth leaves above the inoculated leaf; WT-TRV(R)] or at 3 days post-agroinfiltration (dpa) in leaves agroinfiltrated with a construct expressing the 16K protein (WT-16K or KD-16K). N, nuclei; No, nucleoli; CBs are shown by arrows. Scale bars, 5 µm. (<b>b</b>) Quantification of results presented in (<b>a</b>). The ratio of nucleolar fluorescence to nucleoplasmic fluorescence (Fno/Fnu) was averaged for at least 100 cells in three independent experiments. Data are mean ± SD. Analysis of variance and Tukey’s HSD post hoc tests were performed on the data obtained. ** <span class="html-italic">p</span> &lt; 0.01. (<b>c</b>) Accumulation of PARylated proteins measured by ELISA using rabbit anti-PAR polyclonal antibody, in plants described in (<b>a</b>). Data are mean ± SD, n = 6 from three independent experiments. Analysis of variance and Tukey’s HSD post hoc tests were performed on the data obtained. ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 4
<p>Effect of 3-aminobenzamide (3AB) on the development of TRV infection and accumulation of PARylated proteins in WT <span class="html-italic">N. benthamiana</span> plants. (<b>a</b>) Symptoms induced in TRV-infected and mock-inoculated plants treated with or without 3AB. (<b>b</b>) Northern blot analysis of TRV RNA1 in inoculated (inoc, 7 dpi) and newly emerging systemically infected leaves (seventh and eighth leaves above the inoculated leaf; sys-N, 14 dpi). The positions of RNA size markers are indicated on the left. Ethidium bromide (EtBr)-stained rRNA (bottom panel) is shown as a loading control. Images have been cropped for presentation. Uncropped images are presented in <a href="#app1-viruses-15-01282" class="html-app">Supplementary Materials</a>. (<b>c</b>) Accumulation of TRV RNA1 (measured using RT-qPCR) in inoculated (inoc; 2 and 7 dpi), older systemically infected leaves (third and fourth leaves above the inoculated leaf; sys-O, 7 and 14 dpi) and newly emerging systemically infected leaves (seventh and eighth leaves above the inoculated leaf; sys-N, 14 and 21 dpi). TRV RNA1 expression levels were normalized to those of the internal controls, <span class="html-italic">UBIQUITIN3</span> gene <span class="html-italic">(UBI3)</span> and <span class="html-italic">60S ribosomal protein 23</span> gene <span class="html-italic">(L23)</span>. (<b>d</b>) Accumulation of PARylated proteins was measured by ELISA using rabbit anti-PAR polyclonal antibody, in TRV- systemically infected or mock-inoculated plants treated with or without 3AB (older third and fourth leaves above the inoculated leaf). Data are mean ± SD, <span class="html-italic">n</span> = 6 from three independent experiments. Analysis of variance and Tukey’s HSD post hoc tests were performed on the data obtained. ** <span class="html-italic">p</span> &lt; 0.01; NS, non-significant (<b>c</b>,<b>d</b>). These data could suggest that a change in PARP1 shuttling activity leading to over-accumulation of PAR (associated with PARP target proteins) in TRV-infected leaves subsequently activates host defence and results in plant recovery. However, given that pharmacological PARP inhibitors including 3AB may not only affect the activity of canonical PARPs, but also have off-target effects [<a href="#B1-viruses-15-01282" class="html-bibr">1</a>,<a href="#B2-viruses-15-01282" class="html-bibr">2</a>,<a href="#B48-viruses-15-01282" class="html-bibr">48</a>], results of pharmacological experiments to infer PARP function in plants should be verified by the genetic inhibition of PARP activity.</p>
Full article ">Figure 5
<p>Effect of virus-induced silencing of <span class="html-italic">PARP1</span> (<span class="html-italic">NbPARP1</span>) expression on TRV infection. Two separate PVX-PARP1 VIGS constructs made in this work (see Materials and Methods) exhibited similar effects on <span class="html-italic">NbPARP1</span> gene expression (<b>a</b>), accumulation of PARylated proteins (<b>b</b>) and TRV infection (<b>c</b>,<b>d</b>), which are exemplified in this figure by the data obtained in the experiments using fragment 1. (<b>a</b>) Virus-induced silencing of the <span class="html-italic">PARP1</span> gene in <span class="html-italic">N. benthamiana</span> mediated by a PVX vector which contains fragment 1 of the <span class="html-italic">NbPARP1</span> gene (PVX-PARP), compared with an empty PVX vector control (PVX-C). Accumulation of PARP1 mRNA was measured using RT-qPCR in inoculated (inoc) and newly emerging systemically infected (sys-N) leaves at 10 dpi. Results from three independent experiments (I, II, III) are shown. (<b>b</b>) Effect of PVX-induced PARP1 silencing on the accumulation of PARylated proteins measured by ELISA using a rabbit anti-PAR polyclonal antibody, in the plant leaves shown in (<b>a</b>). Results from three independent experiments (I, II, III) are shown. (<b>c</b>) Accumulation of TRV RNA1 (measured using RT-qPCR) in inoculated (inoc) and newly emerging systemically infected (sys-N) leaves of PARP-silenced plants at 10 dpi; the same leaves as in (<b>a</b>) and (<b>b</b>) were analysed. Results from three independent experiments (I, II, III) are shown (<b>a</b>–<b>c</b>). Data are mean ± SD, <span class="html-italic">n</span> = 4 from three independent biological replicates. Analysis of variance and Tukey’s HSD post hoc tests were performed on the data obtained. ** <span class="html-italic">p</span> &lt; 0.01; NS, non-significant (<b>c</b>,<b>d</b>). (<b>d</b>) Symptoms induced in plants infected with TRV in PARP1 silenced (PVX-PARP) and non-silenced [PVX-C and (-)PVX] plants.</p>
Full article ">Figure 6
<p>Concentrations of free (<b>a</b>) and conjugated (SA-b-glucoside) salicylic acid (SA) (<b>b</b>) in inoculated (inoc; 7 dpi), older systemically infected leaves (sys-O; 14 dpi) and newly emerging sys-N; 14 dpi) leaves of <span class="html-italic">N. benthamiana</span> plants infected or uninfected (mock-inoculated) with TRV and treated with or without 3AB. Statistical analysis was performed on four independent biological replicates. Data are mean ± SD. Each replicate was composed of samples from three plants pooled together (two leaves per plant). Analysis of variance and Tukey’s HSD post hoc tests were performed on the data obtained. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01. NS, non-significant.</p>
Full article ">Figure 7
<p>Effect of foliar treatment with salicylic acid (SA) on the transcript level of the PR-1a protein gene (measured using RT-qPCR) (<b>a</b>), callose deposition (<b>b</b>) and accumulation of TRV (<b>c</b>) in inoculated (inoc; 7 dpi), older systemically infected leaves (sys-O; 14 dpi) and newly emerging sys-N; 14 dpi) leaves of <span class="html-italic">N. benthamiana</span> plants infected or uninfected (mock-inoculated) with TRV and treated with or without 3AB. Statistical analysis was performed on four independent biological replicates. Data are mean ± SD. Each replicate was composed of samples from three plants pooled together (two leaves per plant). Analysis of variance and Tukey’s HSD post hoc tests were performed on the data obtained. *<span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01. (<b>d</b>) Symptoms induced in plants infected with TRV after treatment with 3AB and SA as indicated.</p>
Full article ">Figure 8
<p>A model of the TRV infection process and the involvements of the 16K protein, coilin and PARP1. In healthy plants (<b>a</b>), coilin is located within CBs and the nucleoplasm but is not present in the nucleolus. PARP1, a nuclear protein, modifies the function and subcellular localisation of a variety of nuclear “target” proteins (acceptors) by attaching chains of ADP ribose (PAR) to them. To re-activate these target proteins, PARP1 shuttles them from both the nucleolus (NO) and chromatin (chromatin not shown) to CBs for PAR removal and recycling. Upon TRV infection (<b>b</b>), the viral 16K protein is produced in the cytoplasm and is targeted to the nucleus. In the nucleus (CBs and nucleoplasm), the 16K protein interacts with coilin and relocalises it to the nucleolus, which in turn traps PARP1 within this sub-nuclear domain, preventing it trafficking to CBs for PAR cleavage and recycling. This leads to over-accumulation of PAR/PARylated proteins and may enhance accumulation of salicylic acid (SA) and increased elicitation of SA-mediated defence responses (represented here by increased expression of the <span class="html-italic">PR-1a</span> gene and by callose deposition). These responses restrict TRV spread into newly emerging leaves, leading to the plant’s recovery from TRV infection. Thus, PARP1 can act as a mediator in a functional link between stress-sensing activities of coilin and SA-mediated antivirus defence.</p>
Full article ">
16 pages, 3289 KiB  
Review
Diabetic Polyneuropathy: New Strategies to Target Sensory Neurons in Dorsal Root Ganglia
by Akiko Miyashita, Masaki Kobayashi, Takanori Yokota and Douglas W. Zochodne
Int. J. Mol. Sci. 2023, 24(6), 5977; https://doi.org/10.3390/ijms24065977 - 22 Mar 2023
Cited by 11 | Viewed by 3780
Abstract
Diabetic polyneuropathy (DPN) is the most common type of diabetic neuropathy, rendering a slowly progressive, symmetrical, and length-dependent dying-back axonopathy with preferential sensory involvement. Although the pathogenesis of DPN is complex, this review emphasizes the concept that hyperglycemia and metabolic stressors directly target [...] Read more.
Diabetic polyneuropathy (DPN) is the most common type of diabetic neuropathy, rendering a slowly progressive, symmetrical, and length-dependent dying-back axonopathy with preferential sensory involvement. Although the pathogenesis of DPN is complex, this review emphasizes the concept that hyperglycemia and metabolic stressors directly target sensory neurons in the dorsal root ganglia (DRG), leading to distal axonal degeneration. In this context, we discuss the role for DRG-targeting gene delivery, specifically oligonucleotide therapeutics for DPN. Molecules including insulin, GLP-1, PTEN, HSP27, RAGE, CWC22, and DUSP1 that impact neurotrophic signal transduction (for example, phosphatidylinositol-3 kinase/phosphorylated protein kinase B [PI3/pAkt] signaling) and other cellular networks may promote regeneration. Regenerative strategies may be essential in maintaining axon integrity during ongoing degeneration in diabetes mellitus (DM). We discuss specific new findings that relate to sensory neuron function in DM associated with abnormal dynamics of nuclear bodies such as Cajal bodies and nuclear speckles in which mRNA transcription and post-transcriptional processing occur. Manipulating noncoding RNAs such as microRNA and long-noncoding RNA (specifically MALAT1) that regulate gene expression through post-transcriptional modification are interesting avenues to consider in supporting neurons during DM. Finally, we present therapeutic possibilities around the use of a novel DNA/RNA heteroduplex oligonucleotide that provides more efficient gene knockdown in DRG than the single-stranded antisense oligonucleotide. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic and simplified representation of pathogenetic mechanisms and alterations of gene expression within DRG during DPN. DPN is associated with DRG sensory neuronal atrophy and a reduced intraepidermal nerve fiber density (IENFD) due to a dying-back degeneration of distal axons. Gene expression changes in sensory neurons in DPN are accompanied by structural changes in nuclear bodies, which are essential for transcriptional activity.</p>
Full article ">Figure 2
<p>The structural alternations of nuclear bodies in DPN. In non-diabetic DRG sensory neurons, a single CB (coilin) was in contact with a nucleolus (DAPI), and nuclear speckles (SRSF2) were located in the interchromatin regions of the nucleoplasm in controls (<b>A</b>). Scale bar: 10 μm. In non-diabetic DRG sensory neurons, SMN protein was distributed throughout the cytoplasm and as nuclear foci (<b>B</b>); Scale bars: 20 μm, 10 μm in insets. SMN nuclear foci that collaborate on the assembly of snRNPs were localized within CBs (coilin) in controls (middle row, white arrows) but numerous CBs lost co-localization with SMN nuclear foci (bottom row, yellow arrow) in diabetic nuclei (<b>C</b>). Scale bars: 20 μm, 10 μm in insets. Arrowheads indicate sensory neurons magnified in the insets. SRSF2 is expressed in the nuclear speckles, where MALAT1 is localized, in the DRG sensory neurons in non-diabetic control mice. Anti-neurofilament 200 (NF200) is a marker of large and small myelinated neurons (<b>D</b>). DRG neurons with SRSF2-positive nuclear speckles are moderately reduced in the diabetic mice (<b>E</b>); They are further decreased in diabetic mice with MALAT1 silencing (<b>F</b>). Scale bars = 50 μm, and 20 μm, in insets. (<b>A</b>–<b>C</b>) were adapted with permission from Ref. [<a href="#B41-ijms-24-05977" class="html-bibr">41</a>]. 2017 Zochodne, D.W. and (<b>C</b>–<b>F</b>) from Ref. [<a href="#B44-ijms-24-05977" class="html-bibr">44</a>]. 2022 Yokota, T.</p>
Full article ">Figure 3
<p>Schematic representation of oligonucleotide therapeutics and their mechanisms. HDO suppresses target RNAs. It is a double-stranded artificial functional nucleic acid consisting of a DNA strand as the main strand and an RNA complementary to the main strand, which is called a “gapmer” nucleic acid (LNA). This part is recognized by RNase H, an enzyme that degrades RNA in the cell, and the complementary strand RNA is cleaved. The resulting single main strand binds to the target RNA, and RNase H again cleaves the target RNA to exert its gene suppression effect.</p>
Full article ">
18 pages, 1614 KiB  
Review
The Role of WRAP53 in Cell Homeostasis and Carcinogenesis Onset
by Renan Brito Gadelha, Caio Bezerra Machado, Flávia Melo Cunha de Pinho Pessoa, Laudreísa da Costa Pantoja, Igor Valentim Barreto, Rodrigo Monteiro Ribeiro, Manoel Odorico de Moraes Filho, Maria Elisabete Amaral de Moraes, André Salim Khayat and Caroline Aquino Moreira-Nunes
Curr. Issues Mol. Biol. 2022, 44(11), 5498-5515; https://doi.org/10.3390/cimb44110372 - 4 Nov 2022
Cited by 2 | Viewed by 2653
Abstract
The WD repeat containing antisense to TP53 (WRAP53) gene codifies an antisense transcript for tumor protein p53 (TP53), stabilization (WRAP53α), and a functional protein (WRAP53β, WDR79, or TCAB1). The WRAP53β protein functions as a scaffolding protein that is important [...] Read more.
The WD repeat containing antisense to TP53 (WRAP53) gene codifies an antisense transcript for tumor protein p53 (TP53), stabilization (WRAP53α), and a functional protein (WRAP53β, WDR79, or TCAB1). The WRAP53β protein functions as a scaffolding protein that is important for telomerase localization, telomere assembly, Cajal body integrity, and DNA double-strand break repair. WRAP53β is one of many proteins known for containing WD40 domains, which are responsible for mediating a variety of cell interactions. Currently, WRAP53 overexpression is considered a biomarker for a diverse subset of cancer types, and in this study, we describe what is known about WRAP53β’s multiple interactions in cell protein trafficking, Cajal body formation, and DNA double-strand break repair and its current perspectives as a biomarker for cancer. Full article
(This article belongs to the Special Issue Linking Genomic Changes with Cancer in the NGS Era)
Show Figures

Figure 1

Figure 1
<p>WRAP53β roles in cellular homeostasis. (<b>A</b>) WRAP53β mediates MDC1 and RNF8 interaction at DNA double-strand breaks. Phosphorylation of histone H2AX at DNA damage sites by DNA protein kinases induces its binding to MDC1, which in turn binds to RNF8 through WRAP53β-mediated activity. RNF8 then ubiquitylates the phosphorylated histone and triggers recruitment and accumulation of DNA damage repair machinery at the break point. (<b>B</b>) WRAP53β is essential for Cajal body stability and nuclear function maintenance. WRAP53β mediates SMN1 protein localization in Cajal bodies through transport from the cytoplasm to the nucleus. WRAP53β is also responsible for interacting with and ensuring the activity of the metabolic active telomerase enzyme and for correctly localizing scaRNAs to Cajal bodies, where they will mature and suffer post-transcriptional modifications. Created with BioRender.com (accessed on 30 March 2022).</p>
Full article ">Figure 2
<p>Impacts of WRAP53β mutations in pathological states. Improper binding of WRAP53β to telomerase components leads to progressive shortening of telomeres and is highly implied as a determining factor for the severity of dyskeratosis congenita and for the onset of malignancies. Mutations impairing its ability to mediate DNA double-strand break repairs are also worrisome in the general context of genome stability and may be linked to diverse biological events observed in WRAP53β-deficient cells. Lastly, mutations in WRAP53β or in SMN1 that lead to deficient interactions between the two proteins result in mislocalization of SMN1 in the nucleus and consequent loss of Cajal bodies’ structural integrity, inducing the spinal muscular atrophy phenotype. Created with BioRender.com (accessed 30 March 2022).</p>
Full article ">Figure 3
<p>Frequency of genetic alterations present in the WRAP53 gene in cancer cell lines. A minimum change limit of 0.5% was applied in the creation of the graph (cBioPortal, accessed on 12 August 2022). The numbers in parentheses represent the number of patients analyzed.</p>
Full article ">
14 pages, 2963 KiB  
Communication
A Viral Suppressor of RNA Silencing May Be Targeting a Plant Defence Pathway Involving Fibrillarin
by Miryam Pérez-Cañamás, Michael Taliansky and Carmen Hernández
Plants 2022, 11(15), 1903; https://doi.org/10.3390/plants11151903 - 22 Jul 2022
Cited by 3 | Viewed by 2026
Abstract
To establish productive infections, viruses must be able both to subdue the host metabolism for their own benefit and to counteract host defences. This frequently results in the establishment of viral–host protein–protein interactions that may have either proviral or antiviral functions. The study [...] Read more.
To establish productive infections, viruses must be able both to subdue the host metabolism for their own benefit and to counteract host defences. This frequently results in the establishment of viral–host protein–protein interactions that may have either proviral or antiviral functions. The study of such interactions is essential for understanding the virus–host interplay. Plant viruses with RNA genomes are typically translated, replicated, and encapsidated in the cytoplasm of infected cells. Despite this, a significant array of their encoded proteins has been reported to enter the nucleus, often showing high accumulation at subnuclear structures such as the nucleolus and/or Cajal bodies. However, the biological significance of such a distribution pattern is frequently unknown. Here, we explored whether the nucleolar/Cajal body localization of protein p37 of Pelargonium line pattern virus (PLPV, genus Pelarspovirus, family Tombusviridae), might be related to potential interactions with the nucleolar/Cajal body marker proteins, fibrillarin and coilin. The results revealed that p37, which has a dual role as coat protein and as suppressor of RNA silencing, a major antiviral system in plants, is able to associate with these cellular factors. Analysis of (wildtype and/or mutant) PLPV accumulation in plants with up- or downregulated levels of fibrillarin or coilin have suggested that the former might be involved in an as yet unknown antiviral pathway, which may be targeted by p37. The results suggest that the growing number of functions uncovered for fibrillarin can be wider and may prompt future investigations to unveil the plant antiviral responses in which this key nucleolar component may take part. Full article
(This article belongs to the Special Issue Plant Defense Responses against Viruses)
Show Figures

Figure 1

Figure 1
<p>In vivo analysis of p37 interaction with fibrillarin and coilin. WT p37, fibrillarin (FIB), and coilin (COIL) were tagged at their N-terminus with YFP halves (sYFPN and sYFPC) and transiently co-expressed in <span class="html-italic">N. benthamiana</span> leaves to study the protein–protein interactions through a BiFC assay. FIB-mRFP was also co-expressed in these assays as a nucleolar/CB marker. Confocal laser scanning microscopy was used for the observation of fluorescence at 3 dpif. The micrographs (<b>A1</b>,<b>B1</b>) show a general view of sYFP-derived fluorescence in epidermal cells expressing sYFPN:FIB and sYFPN:COIL, respectively, in combination with sYFPC-tagged p37 molecules; the micrographs (<b>A2</b>,<b>B2</b>) show a close view of sYFP-derived fluorescence from the same protein-tagged combinations; the micrographs (<b>A3</b>,<b>B3</b>) correspond to the same close views but record the mRFP fluorescence derived from FIB-mRFP; the micrographs (<b>A4</b>,<b>B4</b>) show overlays of (<b>A2</b>,<b>A3</b>) and of (<b>B2</b>,<b>B3</b>), respectively. Equivalent images were obtained with reverse combinations (sYFPC:FIB or sYFPC:COIL co-expressed with sYFPN-tagged p37 molecule) (data not shown). Negative control combinations (sYFPN-sYFPC:FIB; sYFPN-sYFPC:COIL) are displayed in micrographs (<b>A5</b>,<b>B5</b>), respectively. Additional negative controls were included and shown in micrographs (<b>C1</b>–<b>C4</b>), (<b>D1</b>–<b>D4</b>), and (<b>E1</b>–<b>E4</b>). These controls corresponded to two transcription factors, FUL and SOC, that interact between them in the nucleus and cytoplasm (<b>E1</b>–<b>E4</b>) but that did not show interaction with p37 (<b>C1</b>–<b>C4</b>,<b>D1</b>–<b>D4</b>). Nu: Nucleolus. CB: Cajal body. The inset scale bar corresponds to 20 µm in the first column panels and to 10 µm in all remaining panels.</p>
Full article ">Figure 2
<p>Accumulation of wt PLPV in lines of <span class="html-italic">N. benthamiana</span> with downregulated fibrillarin or coilin content. The wt PLPV was agroinoculated in wt and in FIBi and COILi <span class="html-italic">N. benthamiana</span> lines. Northern blot (left) and RT-qPCR (right) analyses of local (<b>A)</b> and systemic (<b>B</b>) leaves collected at 7 dpi and 30 dpi, respectively. For Northern blot, a PLPV- specific riboprobe was employed. Positions of the PLPV genomic (g) and subgenomic (sg) RNAs are indicated at the margin of the blots, and ethidium bromide staining of rRNAs is included below the blots as loading controls. Two distinct samples are shown for each virus–plant line combination. A mock-inoculated plant was included as negative control. In the graphs to the right, corresponding to the RT-qPCR measurement of relative PLPV accumulation, the bars depict the standard deviation from three independent biological replicates. The statistical significance was evaluated using a paired t-test (no significant differences were detected).</p>
Full article ">Figure 3
<p>In vivo analysis of the potential interaction of p37WA mutant with fibrillarin and coilin. Protein p37WA, fibrillarin (FIB), and Coilin (COIL) were tagged at their N-terminus with YFP halves (sYFPN and sYFPC) and transiently co-expressed in <span class="html-italic">N. benthamiana</span> leaves to study the protein–protein interactions through a BiFC assay. FIB-mRFP was also co-expressed in these assays as a nucleolar/CB marker. Confocal laser scanning microscopy was used for the observation of fluorescence at 3 dpi. The micrographs (<b>A1</b>,<b>B1</b>) show a general view of YFP-derived fluorescence in epidermal cells expressing sYFPN:FIB and sYFPN:COIL, respectively, in combination with sYFPC-tagged p37WA molecules; the micrographs (<b>A2</b>,<b>B2</b>) show a close view of YFP-derived fluorescence from the same protein-tagged combinations; the micrographs (<b>A3</b>,<b>B3</b>) correspond to the same close views but record the RFP fluorescence derived from FIB:mRFP; the micrographs (<b>A4</b>,<b>B4</b>) show overlays of (<b>A2,A3,B2,B3</b>), respectively. Equivalent images were obtained with reverse combinations (sYFPC:FIB or sYFPC:COIL co-expressed with sYFPN-tagged p37WA molecule) (data not shown). Negative control combinations (sYFPN-sYFPC:FIB; sYFPN-sYFPC:COIL) are displayed in micrographs (<b>A5</b>,<b>B5</b>), respectively. Nu: nucleolus. CB: Cajal body. The inset scale bar corresponds to 20 µm in the first column panels and to 10 µm in all remaining panels.</p>
Full article ">Figure 4
<p>Accumulation of PLPV-mutp37GA in lines of <span class="html-italic">N. benthamiana</span> with downregulated fibrillarin or coilin content. The PLPV-mutp37WA was inoculated in wt, FIBi, and COILi <span class="html-italic">N. benthamiana</span> lines. Northern blot (<b>A</b>) and RT-qPCR (<b>B</b>) analyses of local leaves collected at 7 dpi. For Northern blot (<b>A</b>), a PLPV- specific riboprobe was employed. Positions of the PLPV genomic (g) and subgenomic (sg) RNAs are indicated at the left of the blot, and ethidium bromide staining of rRNAs is included below the blot as loading control. Two distinct samples are shown for each virus–plant line combination. A mock-inoculated plant was included as negative control. In (<b>B</b>), the bars depict the standard deviation from three independent biological replicates. The statistical significance was checked using a paired <span class="html-italic">t</span>-test (** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 5
<p>Analysis of fibrillarin interaction with wt p37 and mutant p37WA through a co-immunoprecipitation assay. FIB:mRFP was co-expressed with HA-tagged wt p37 or p37WA in <span class="html-italic">N. benthamiana</span> leaves. <span class="html-italic">N. bentamiana</span> leaves expressing only HA-tagged p37WA were included as control. Input protein extracts or immunoprecipitates (IP) obtained with the employment of ChromoTek RFP-Trap agarose were subjected to Western blot (WB) analysis using either an anti-RFP antibody (for detection of fibrillarin; upper blot) or anti-HA antibody (for detection of p37 molecule; lower blot).</p>
Full article ">Figure 6
<p>Analysis of PLPV accumulation in <span class="html-italic">N. benthamiana</span> leaves overexpressing fibrillarin. <span class="html-italic">N. benthamiana</span> leaves were agroinfiltrated with a pROK2-based plasmid directing expression of FIB:mRFP or with a plasmid without insert (mock control), and at 3 dpif, the leaves were agroinoculated with wt PLPV. Total RNA preparations were obtained at 3 dpi and subjected to RT-qPCR to estimate relative viral accumulation. The bars depict the standard deviation from three independent biological replicates. The statistical significance was evaluated using a paired <span class="html-italic">t</span>-test (* <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 7
<p>Assessment of mRNA fibrillarin levels. Total RNA preparations were obtained from mock and PLPV-infected wt <span class="html-italic">N. benthamiana</span> plants and subjected to RT-qPCR to estimate relative levels of fibrillarin transcripts. The bars depict the standard deviation from three independent biological replicates. The statistical significance was evaluated using a paired <span class="html-italic">t</span>-test (no significant differences were detected).</p>
Full article ">Figure 8
<p>Proposed model reflecting the involvement of fibrillarin in an antiviral pathway targeted by PLPV p37. Protein p37 enters the nucleus assisted by an importin (α, β)-mediated pathway (Pérez-Cañamás and Hernández, 2018) and reaches high accumulation at nucleolus and Cajal bodies (CB). In these subnuclear structures, p37 interacts with the signature proteins, coilin and fibrillarin. The latter interaction interferes with a fibrillarin-linked antiviral pathway, thus favouring PLPV infection.</p>
Full article ">
18 pages, 3940 KiB  
Article
Function of Cajal Bodies in Nuclear RNA Retention in A. thaliana Leaves Subjected to Hypoxia
by Sylwia Górka, Dawid Kubiak, Małgorzata Ciesińska, Katarzyna Niedojadło, Jarosław Tyburski and Janusz Niedojadło
Int. J. Mol. Sci. 2022, 23(14), 7568; https://doi.org/10.3390/ijms23147568 - 8 Jul 2022
Cited by 1 | Viewed by 2161
Abstract
Retention of RNA in the nucleus precisely regulates the time and rate of translation and controls transcriptional bursts that can generate profound variability in mRNA levels among identical cells in tissues. In this study, we investigated the function of Cajal bodies (CBs) in [...] Read more.
Retention of RNA in the nucleus precisely regulates the time and rate of translation and controls transcriptional bursts that can generate profound variability in mRNA levels among identical cells in tissues. In this study, we investigated the function of Cajal bodies (CBs) in RNA retention in A. thaliana leaf nuclei during hypoxia stress was investigated. It was observed that in ncb-1 mutants with a complete absence of CBs, the accumulation of poly(A+) RNA in the leaf nuclei was lower than that in wt under stress. Moreover, unlike in root cells, CBs store less RNA, and RNA retention in the nuclei is much less intense. Our results reveal that the function of CBs in the accumulation of RNA in nuclei under stress depends on the plant organ. Additionally, in ncb-1, retention of introns of mRNA RPB1 (largest subunit of RNA polymerase II) mRNA was observed. However, this isoform is highly accumulated in the nucleus. It thus follows that intron retention in transcripts is more important than CBs for the accumulation of RNA in nuclei. Accumulated mRNAs with introns in the nucleus could escape transcript degradation by NMD (nonsense-mediated mRNA decay). From non-fully spliced mRNAs in ncb-1 nuclei, whose levels increase during hypoxia, introns are removed during reoxygenation. Then, the mRNA is transferred to the cytoplasm, and the RPB1 protein is translated. Despite the accumulation of isoforms in nuclei with retention of introns in reoxygenation, ncb-1 coped much worse with long hypoxia, and manifested faster yellowing and shrinkage of leaves. Full article
(This article belongs to the Special Issue Response to Environmental Stress in Plants)
Show Figures

Figure 1

Figure 1
<p>The relative expression levels of mRNA ADH1 (alcohol dehydrogenase 1). X-axis control (Control), time of hypoxia (1, 3, 6H—hours and 1, 5D—days of hypoxia, 5D + 1D—5 days hypoxia and 1-day reoxygenation). Variants labeled with the same letters are not significantly different (<span class="html-italic">p</span> ≤ 0.05).</p>
Full article ">Figure 2
<p>Localization of poly(A<sup>+</sup>) RNA by whole-mount FISH in leaves in wt U2B”:GFP (<b>A</b>–<b>E</b>) and <span class="html-italic">ncb-1</span> U2B”:GFP (<b>F</b>–<b>J</b>) cultivated to control (<b>A</b>,<b>F</b>) hypoxia (<b>B</b>–<b>D</b>,<b>G</b>–<b>I</b>) and reoxygenation (<b>E</b>,<b>J</b>) condition. Identification nuclei and CBs by U2B”-GFP (<b>A’</b>–<b>J’</b>). 1, 3, 5H—hours and 1, 5D—days of hypoxia, 5D + 1D—5 days of hypoxia and 1-day reoxygenation (indicated in the top right corner of the image). The inserts show enlarge, nuclei, and CBs in wt (<b>A’1</b>–<b>A’3</b>,<b>D’1</b>–<b>D’3</b>) and <span class="html-italic">ncb-1</span> (<b>F’1</b>–<b>F’3</b>,<b>I’1</b>–<b>I’3</b>). Arrowhead—nuclei, arrow—Cajal bodies; C—cytoplasm, Nu—nucleolus; in inserts localization of poly(A<sup>+</sup>) RNA w nuclei (<b>A’1</b>,<b>B’1</b>,<b>D’1</b>,<b>I’1</b>), identification of nuclei and CBs (<b>A’2</b>,<b>B’2</b>,<b>D’2</b>,<b>I’2</b>) with U2B”-GFP and merge (<b>A’3</b>,<b>B’3</b>,<b>D’3</b>,<b>I’3</b>). Bar 10 µm.</p>
Full article ">Figure 3
<p>The relative fluorescence intensity of poly(A<sup>+</sup>) RNA in cytoplasm (<b>A</b>), nucleus (<b>B</b>), CBs (<b>C</b>). X-axis—1, 3, 5D—days of hypoxia, 5D + 1D—5 days hypoxia and 1-day reoxygenation. Variants labeled with the same letters are not significantly different (<span class="html-italic">p</span> ≤ 0.05).</p>
Full article ">Figure 4
<p>Localization and quantitative analysis of active RNA Pol II in the roots cells <span class="html-italic">Arabidopsis thaliana.</span> Distribution of ser2 of CTD RNA Pol II RNA in wt (<b>A</b>–<b>F</b>) and <span class="html-italic">ncb-1</span> (<b>A’</b>–<b>F’</b>) in control (Control), hypoxia (6H, 1D, 3D, 5D) and reoxygenation condition (5D + 1D) (indicated in the top right corner of the image). In B’ dashes line indicate nucleus. Histogram shows the quantity of ser2 of CTD RNA Pol II RNA in nuclei in different of stages (<b>G</b>). X-axis—6H- hours and 1, 3, 5D—days of hypoxia, 5D + 1D—5 days of hypoxia and 1-day reoxygenation. Variants labeled with the same letters are not significantly different (<span class="html-italic">p</span> ≤ 0.05). Bar 5 µm.</p>
Full article ">Figure 5
<p>Structure of the largest RPB1 subunit for RNA polymerase II in <span class="html-italic">Arabidopsis thaliana</span> (the thicker dark green bars represent exon sequences, and the thinner bars indicate intron sequences) [<a href="#B55-ijms-23-07568" class="html-bibr">55</a>] (<b>A</b>), Analysis of quantity and ratio of isoforms mRNA RPB1 with intron retention and without introns. The relative expression of transcripts of <span class="html-italic">RPB1</span> with exon 8 and 9 (<b>B</b>), Gel separation of PCR products with specific primers to exon 3 and 4 mRNA RPB1. line 1—Control <span class="html-italic">ncb-1</span>-; line 2—Control wt; line 3—hypoxia <span class="html-italic">ncb-1</span>; 4—hypoxia wt; 5—five-day hypoxia and one day reoxidation <span class="html-italic">ncb-1</span>; 6—five-day hypoxia and one-day reoxidation wt (<b>C</b>), The relative expression of transcripts of <span class="html-italic">RPB1</span> with intron 3 in wt and <span class="html-italic">ncb-1</span> (<b>D</b>). Variants labeled with the same letters are not significantly different (<span class="html-italic">p</span> ≤ 0.05). 1D—one-day hypoxia, 5D—five-day hypoxia, 5D + 1D five-day hypoxia and one-day reoxygenation.</p>
Full article ">Figure 6
<p>Localization of mRNA RPB1 with intron 3 in the nuclei (<b>A</b>,<b>B</b>,<b>E</b>,<b>F</b>) and leaf sections (<b>A’</b>,<b>B’</b>,<b>E’</b>,<b>F’</b>) and exon 8 in the nuclei (<b>C</b>,<b>D</b>,<b>G</b>,<b>H</b>) and leaves sections (<b>C’</b>,<b>D’</b>,<b>G’</b>,<b>H’</b>). In A and B dashes line indicate nucleus. Control—control condition, 5D—five day of hypoxia (indicated in the top right corner of the image). Bar: 5 µm (<b>A</b>–<b>H</b>) and 100 µm (<b>A’</b>–<b>H’</b>), V—vacuole. On sections nuclei are stained with DAPI.</p>
Full article ">Figure 7
<p>Measurements of the amount of RPB1 mRNA with exon 8 and intron 3 in nuclei wt and <span class="html-italic">ncb-1</span> in normoxia (Control), hypoxia (1D, 5D) and reoxygenation (5D + 1D). Variants labeled with the same letters are not significantly different (<span class="html-italic">p</span> ≤ 0.05).</p>
Full article ">
20 pages, 2696 KiB  
Review
Regulation of Cellular Ribonucleoprotein Granules: From Assembly to Degradation via Post-translational Modification
by Pureum Jeon, Hyun-Ji Ham, Semin Park and Jin-A Lee
Cells 2022, 11(13), 2063; https://doi.org/10.3390/cells11132063 - 29 Jun 2022
Cited by 13 | Viewed by 4779
Abstract
Cells possess membraneless ribonucleoprotein (RNP) granules, including stress granules, processing bodies, Cajal bodies, or paraspeckles, that play physiological or pathological roles. RNP granules contain RNA and numerous RNA-binding proteins, transiently formed through the liquid–liquid phase separation. The assembly or disassembly of numerous RNP [...] Read more.
Cells possess membraneless ribonucleoprotein (RNP) granules, including stress granules, processing bodies, Cajal bodies, or paraspeckles, that play physiological or pathological roles. RNP granules contain RNA and numerous RNA-binding proteins, transiently formed through the liquid–liquid phase separation. The assembly or disassembly of numerous RNP granules is strongly controlled to maintain their homeostasis and perform their cellular functions properly. Normal RNA granules are reversibly assembled, whereas abnormal RNP granules accumulate and associate with various neurodegenerative diseases. This review summarizes current studies on the physiological or pathological roles of post-translational modifications of various cellular RNP granules and discusses the therapeutic methods in curing diseases related to abnormal RNP granules by autophagy. Full article
(This article belongs to the Special Issue The Autophagic Process in Human Physiology and Pathogenesis)
Show Figures

Figure 1

Figure 1
<p>Cellular ribonucleoprotein (RNP) granules. (<b>a</b>,<b>b</b>) Schematic structural representation of cytoplasmic RNP granules. (<b>a</b>) Stress granules (SGs) containing untranslated mRNA, ribosomes, translational initiation factors, and RBPs. (<b>b</b>) P-bodies containing untranslated mRNAs and RBPs. (c and d) Schematic structure of nuclear RNP granules. (<b>c</b>) Paraspeckles containing lncRNA NEAT1 and nuclear-localized RBPs. (<b>d</b>) Cajal bodies containing snRNPs, snoRNPs, and nuclear RNPs.</p>
Full article ">Figure 2
<p>Physiological/pathological RNP granules regulated by PTMs. (<b>a</b>) Physiological RNP granules composed of proteins modified by PTMs, such as arginine methylation, phosphorylation, acetylation, ubiquitination, and glycosylation. PTMs can regulate RNP granule dynamics by affecting the interaction strength between proteins and nucleic acids. (<b>b</b>) RNP-binding proteins linked to neurodegenerative diseases can be modified by phosphorylation, acetylation, or PARylation (<a href="#cells-11-02063-f003" class="html-fig">Figure 3</a>), altering the biophysical properties. Accumulation of aggregates associated with altered RNP granules in neurons is a hallmark of several neurodegenerative diseases. (<b>c</b>) Abnormal RNP granules and aggregates can be degraded by granulophagy.</p>
Full article ">Figure 3
<p>PTMs on pathological RNP granules. (<b>a</b>) Increased aggregation property by PTMs on RBPs: phosphorylation of tau [<a href="#B114-cells-11-02063" class="html-bibr">114</a>] or TDP-43 [<a href="#B121-cells-11-02063" class="html-bibr">121</a>], acetylation of TDP-43 [<a href="#B122-cells-11-02063" class="html-bibr">122</a>], PARylation of TDP-43, or hnRNPA1 [<a href="#B103-cells-11-02063" class="html-bibr">103</a>] accelerate the solid transition. (<b>b</b>) Decreased aggregation property by PTMs on RBPs: methylation [<a href="#B85-cells-11-02063" class="html-bibr">85</a>,<a href="#B116-cells-11-02063" class="html-bibr">116</a>], phosphorylation [<a href="#B119-cells-11-02063" class="html-bibr">119</a>], or acetylation [<a href="#B120-cells-11-02063" class="html-bibr">120</a>] of FUS; phosphorylation [<a href="#B117-cells-11-02063" class="html-bibr">117</a>,<a href="#B118-cells-11-02063" class="html-bibr">118</a>,<a href="#B122-cells-11-02063" class="html-bibr">122</a>], or acetylation of tau [<a href="#B115-cells-11-02063" class="html-bibr">115</a>] prevent the phase to solid transition.</p>
Full article ">
Back to TopTop