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

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

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (2,532)

Search Parameters:
Keywords = DNA polymerase

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
15 pages, 807 KiB  
Review
ATP-Dependent Chromatin Remodeler CSB Couples DNA Repair Pathways to Transcription with Implications for Cockayne Syndrome and Cancer Therapy
by Rabeya Bilkis, Robert J. Lake and Hua-Ying Fan
Cells 2025, 14(4), 239; https://doi.org/10.3390/cells14040239 (registering DOI) - 7 Feb 2025
Viewed by 81
Abstract
Efficient DNA lesion repair is crucial for cell survival, especially within actively transcribed DNA regions that contain essential genetic information. Additionally, DNA breaks in regions of active transcription are prone to generating insertions and deletions, which are hallmark features of cancer genomes. Cockayne [...] Read more.
Efficient DNA lesion repair is crucial for cell survival, especially within actively transcribed DNA regions that contain essential genetic information. Additionally, DNA breaks in regions of active transcription are prone to generating insertions and deletions, which are hallmark features of cancer genomes. Cockayne syndrome protein B (CSB) is the sole ATP-dependent chromatin remodeler that is essential for coupling DNA repair pathways with transcription, leading to more efficient DNA repair in regions of active transcription. CSB is best known for its essential function in transcription-coupled nucleotide excision repair (TC-NER), a process that rapidly removes helix-distorting DNA lesions that stall RNA polymerase II, such as those created by chemotherapeutic platinum compounds and UV irradiation. In addition to NER, CSB has also been reported to couple homologous recombination to transcription. Most recently, CSB has also been shown to couple single-strand DNA break repair to transcription. In this review, we will discuss the overlapping and distinct mechanisms by which CSB couples these different DNA repair pathways to transcription. We will also discuss how these CSB functions may account for Cockayne syndrome and the emerging roles of CSB as an innovative target for cancer therapy. Full article
Show Figures

Figure 1

Figure 1
<p>Coupling of NER and SSBR to transcription by CSB. (<b>A</b>) Schematic of CSB’s functional domains. PBM: PARylation binding module. AD: acidic domain. N1: coupling chromatin remodeling to ATPase activity. (<b>B</b>) A minimalized TC-NER model. CSB locates bulky DNA lesion-stalled transcription in an ATP-dependent manner. ATP hydrolysis by CSB is predicted to induce a conformational change that permits stable chromatin association. CSB replaces DSIF, transitioning polymerase II from a transcription elongation complex to a DNA repair complex. CSB is essential for the rapid recruitment of the NER protein as compared to GG-NER. ATP-dependent chromatin remodeling by CSB may enhance the DNA repair efficiency by altering the chromatin environment and stimulating the recruitment of NER proteins. (<b>C</b>) A proposed TA-SSBR model. Single-stranded DNA breaks are detected and bound by PARP1 and PARP2, which activates their enzymatic activity. CSB is then recruited through direct interaction with PARP1 or PARP2, promoting the recruitment of HPF1, which alters the substrate specificity of PARP1 and PARP2, most notably facilitating histone PARylation, which likely promotes chromatin relaxation to facilitate DNA repair. XRCC1-containing SSBR complexes are then rapidly recruited by CSB and PARylated PARP1, PARP2, and histones. CSB may likely facilitate the dissociation of PARP1/2 from DNA breaks to promote the rapid progression of DNA repair in transcribed DNA regions.</p>
Full article ">Figure 2
<p>Functional domain comparison between PARP1 and PARP2. PARP1 binds to SSBs with a variety of end chemistries through its zinc fingers and Trp-Gly-Arg (WGR) domain. The zinc fingers are absent in PARP2, which binds primarily to the 5′ phosphate ends of SSBs through its WGR domain and N-terminal region (NTR). The WGR domain, present in both PARP1 and PARP2, is critical in inducing conformational changes and the allosteric regulation of enzymatic activity.</p>
Full article ">
25 pages, 719 KiB  
Review
Diagnostic Assays for Avian Influenza Virus Surveillance and Monitoring in Poultry
by Shahan Azeem and Kyoung-Jin Yoon
Viruses 2025, 17(2), 228; https://doi.org/10.3390/v17020228 - 6 Feb 2025
Viewed by 299
Abstract
Diagnostic testing plays a key role in a surveillance program as diagnostic testing aims to accurately determine the infection or disease status of an individual animal. Diagnostic assays for AIV can be categorized into four broad types: tests for detecting the virus, its [...] Read more.
Diagnostic testing plays a key role in a surveillance program as diagnostic testing aims to accurately determine the infection or disease status of an individual animal. Diagnostic assays for AIV can be categorized into four broad types: tests for detecting the virus, its antigen, its genomic material, and antibodies to the virus. Virus characterization almost always follows virus detection. The present article surveys the current literature on the goals, principles, test performance, advantages, and disadvantages of these diagnostic assays. Virus isolation can be achieved using embryonating eggs or cell cultures in a lab setting. Virus antigens can be detected by antigen-capturing immunoassays or tissue immunoassays. Viral RNA can be detected by PCR-based assays (gel-based reverse transcription–polymerase chain reaction (RT-PCR), or probe or SYBR® Green-based real-time RT-PCR), loop-mediated isothermal amplification, in situ hybridization, and nucleic acid sequence-based amplification. Antibodies to AIV can be detected by ELISA, agar gel immunodiffusion, hemagglutination inhibition, and microneutralization. Avian influenza virus can be characterized by hemagglutination inhibition, neuraminidase inhibition, sequencing (dideoxynucleotide chain-termination sequencing, next-generation sequencing), genetic sequence-based pathotype prediction, and pathogenicity testing. Novel and variant AIVs can be recognized by DNA microarrays, electron microscopy, mass spectroscopy, and Biological Microelectromechanical Systems. A variety of diagnostic tests are employed in AIV surveillance and monitoring. The choice of their use depends on the goal of testing (fit for purpose), the time of testing during the disease, the assay target, the sample matrix, assay performance, and the advantages and disadvantages of the assay. The article concludes with authors’ perspective of the use of diagnostic assays in the surveillance and monitoring of AIV in poultry. Full article
Show Figures

Figure 1

Figure 1
<p>The Nucleic Acid Sequence-Based Amplification process (adopted from Rodríguez-Lázaro et al. 2006 with permission) [<a href="#B64-viruses-17-00228" class="html-bibr">64</a>]. The wavy lines show RNA, and the straight lines show DNA.</p>
Full article ">
15 pages, 1055 KiB  
Article
Metagenomic Insights into Microbial Signatures in Thrombi from Acute Ischemic Stroke Patients Undergoing Endovascular Treatment
by Kasthuri Thirupathi, Sherief Ghozy, Abdullah Reda, Wasantha K. Ranatunga, Mars A. Ruben, Zarrintan Armin, Oana M. Mereuta, Sekhon Prabhjot, Daying Dai, Waleed Brinjikji, David F. Kallmes and Ramanathan Kadirvel
Brain Sci. 2025, 15(2), 157; https://doi.org/10.3390/brainsci15020157 - 6 Feb 2025
Viewed by 345
Abstract
Background: Variability in recanalization success during endovascular treatment for acute ischemic stroke (AIS) has led to increased interests in thrombus composition and associated cellular materials. While evidence suggests that bacteria may influence thrombus characteristics, limited data exist on microbiological profiles of thrombi in [...] Read more.
Background: Variability in recanalization success during endovascular treatment for acute ischemic stroke (AIS) has led to increased interests in thrombus composition and associated cellular materials. While evidence suggests that bacteria may influence thrombus characteristics, limited data exist on microbiological profiles of thrombi in stroke patients. Objectives: Characterization of bacterial communities present in thrombi of AIS patients undergoing mechanical thrombectomy, providing insights into microbial contributions to stroke pathogenesis and treatment outcomes. Methods: Thrombi were collected from 20 AIS patients. After extracting metagenome, 16S rDNA sequencing was performed. Bioinformatic analysis included taxonomy and diversity assessments. The presence of bacterial DNA and viable bacteria in thrombi was validated using polymerase chain reaction (PCR) and bacterial culturing followed by matrix-assisted laser desorption ionization–time of flight (MALDI-TOF) analysis, respectively. Results: 16S rDNA was amplified in 19/20 thrombi (95%). Analysis identified a diverse microbial community, with Corynebacterium spp. as the most prevalent genus, followed by Staphylococcus spp., Bifidobacterium spp., Methylobacterium spp., and Anaerococcus spp. Alpha diversity analyses (Shannon index: 4.0–6.0 and Simpson index: 0.8–1.0) revealed moderate to high microbial diversity across samples; beta diversity demonstrated distinct clustering, indicating inter-patient variability in microbial profiles. PCR confirmed the presence of DNA specific to dominant bacterial taxa identified through sequencing. Culturing showed the presence of Staphylococcus epidermidis and Enterococcus faecalis in some clots as identified through MALDI analysis. Conclusions: This study shows bacterial communities present in AIS patients’ thrombi, suggesting a potential link between microbial signatures and thrombus characteristics. Full article
(This article belongs to the Section Neurorehabilitation)
Show Figures

Figure 1

Figure 1
<p>Relative abundance of the 10 most prevalent bacterial taxa in thrombotic blood clots.</p>
Full article ">Figure 2
<p>Alpha diversity indices for each sample. (<b>A</b>) Shannon diversity index; (<b>B</b>) Simpson diversity index; (<b>C</b>) Rarefaction curves showing taxonomic richness across all samples.</p>
Full article ">Figure 3
<p>Principal Coordinates Analysis (PCoA) plot based on Bray–Curtis dissimilarity distances.</p>
Full article ">
23 pages, 4635 KiB  
Article
Spontaneous Necrosis of a High-Risk Bladder Tumor Under Immunotherapy for Concurrent Malignant Melanoma: Role of BRAF Mutations and PD-L1 Expression
by Cristian Condoiu, Mihael Musta, Alin Adrian Cumpanas, Razvan Bardan, Vlad Dema, Flavia Zara, Cristian Silviu Suciu, Cristina-Stefania Dumitru, Andreea Ciucurita, Raluca Dumache, Hossam Ismail and Dorin Novacescu
Biomedicines 2025, 13(2), 377; https://doi.org/10.3390/biomedicines13020377 - 5 Feb 2025
Viewed by 447
Abstract
Background: Bladder cancer (BC) is a heterogeneous malignancy, and predicting response to immune checkpoint inhibitors (ICIs) remains a challenge. Herein, we investigate a high-risk bladder tumor, which developed during anti-BRAF/MEK therapy for a concurrent advanced BRAF-V600E-positive malignant melanoma (MM) and subsequently underwent [...] Read more.
Background: Bladder cancer (BC) is a heterogeneous malignancy, and predicting response to immune checkpoint inhibitors (ICIs) remains a challenge. Herein, we investigate a high-risk bladder tumor, which developed during anti-BRAF/MEK therapy for a concurrent advanced BRAF-V600E-positive malignant melanoma (MM) and subsequently underwent complete spontaneous necrosis following Nivolumab immunotherapy, only to recur thereafter while still under the same treatment. This unique scenario provided an opportunity to investigate the roles of BRAF gene mutations in BC pathogenesis, respectively, of PD-L1 expression in immunotherapy response prediction. Methods: We retrospectively analyzed BC specimens obtained via transurethral resection at two critical time-points: prior to the complete spontaneous necrosis under Nivolumab (prenecrosis) and after tumor recurrence postnecrosis (postnecrosis). The BRAF gene mutation status was evaluated using quantitative polymerase chain reaction (qPCR). PD-L1 expression was assessed by immunohistochemistry (IHC), quantified using the combined positive score (CPS), and a cutoff of ≥10 for positivity. Results: Neither pre- nor postnecrosis BC samples harbored BRAF gene mutations. Prenecrosis PD-L1 expression (CPS = 5) indicated a minimal likelihood of response to immunotherapy. However, complete spontaneous necrosis occurred under Nivolumab, followed by recurrence with further reduced PD-L1 expression (CPS = 1). Conclusions: The complete BC regression challenges the conventional role of PD-L1 as a sole predictive biomarker for immunotherapy. This study also highlights the potential role of BRAF/MEK inhibitors in BC oncogenesis and underscores the need for alternative biomarkers, such as tumor mutation burden (TMB) and circulating tumor DNA (ctDNA), to guide treatment selection in BC better. Full article
Show Figures

Figure 1

Figure 1
<p>Imaging at initial presentation, i.e., details from contrast-enhanced computer tomography of thorax and abdomen, arterial phase, demonstrating malignant melanoma clinical stage at diagnosis (pT4b cN2 M1lym): (<b>A</b>) coronal view, primary lesion (red circle) in the right subscapular area; (<b>B</b>) axial view, right axillary adenopathic block (yellow circle); (<b>C</b>) axial view, large upper mediastinal adenopathy (green circle); (<b>D</b>) coronal view, right hilar and peribronchial adenopathic block (blue square).</p>
Full article ">Figure 2
<p>Contrast-enhanced computer tomography scan of the pelvis, axial view, and excretory phase (15 min), showing a voluminous recurrence of the bladder tumor during combined targeted chemotherapy (anti-BRAF/MEK) for malignant melanoma.</p>
Full article ">Figure 3
<p>Microscopic findings in prespontaneous bladder tumor necrosis tissue fragments obtained through transurethral resection (biopsy): (<b>A</b>) 200×, HE, nests of neoplastic urothelial cells in the deep portion of the proliferation, showing focal invasion of the submucosa; (<b>B</b>) 200×, HE, high-grade traits, G2 cellularity, in the superficial portion of the proliferation, with some mitotic activity; (<b>C</b>) 200×, HE, intratumoral vascular elements in the superficial portion of the proliferation; (<b>D</b>) 200×, IHC with anti-PD-L1 (Dako clone 22C3), moderate to intense staining reaction in tumor cells, and isolated infiltrating immune cells (CPS = 5%); (<b>E</b>) 200×, HE, high-grade area with abundant vascularity and tumor-associated inflammatory cells; (<b>F</b>) 200×, IHC with anti-PD-L1 (Dako clone 22C3), moderate to intense staining reaction in tumor cells and tumor-associated inflammatory cells (CPS = 5%).</p>
Full article ">Figure 4
<p>Microscopic findings in postspontaneous bladder tumor necrosis relapse tissue fragments obtained through transurethral resection: (<b>A</b>) 200×, HE, urothelial carcinoma proliferation, invading the basal membrane and superficial submucosa focally, while associating a significant submucosal desmoplastic reaction; (<b>B</b>) 200×, IHC with anti-PD-L1 (Dako clone 22C3), a moderate color reaction in tumor-associated inflammatory cells (CPS = 1); (<b>C</b>) 200×, HE, urothelial carcinoma proliferation, with abundant vascularization and a microcalcification focus; (<b>D</b>) 200×, IHC with anti-PD-L1 (Dako clone 22C3), a very weak color reaction in rare tumor-associated inflammatory cells (CPS = 1); (<b>E</b>) 200×, HE, urothelial carcinoma with a classic papillary-type growth pattern; (<b>F</b>) 200×, IHC with anti-PD-L1 (Dako clone 22C3), a weak color reaction in scarce tumor cells and a few tumor-associated inflammatory cells (CPS = 1).</p>
Full article ">
Article
The POLG Variant c.678G>C; p.(Gln226His) Is Associated with Mitochondrial Abnormalities in Fibroblasts Derived from a Patient Compared to a First-Degree Relative
by Imra Mantey, Felix Langerscheidt, Çağla Çakmak-Durmaz, Naomi Baba, Katharina Burghardt, Mert Karakaya and Hans Zempel
Genes 2025, 16(2), 198; https://doi.org/10.3390/genes16020198 - 5 Feb 2025
Viewed by 276
Abstract
Background: The nuclear-encoded enzyme polymerase gamma (Pol-γ) is crucial in the replication of the mitochondrial genome (mtDNA), which in turn is vital for mitochondria and hence numerous metabolic processes and energy production in eukaryotic cells. Variants in the POLG gene, which encodes [...] Read more.
Background: The nuclear-encoded enzyme polymerase gamma (Pol-γ) is crucial in the replication of the mitochondrial genome (mtDNA), which in turn is vital for mitochondria and hence numerous metabolic processes and energy production in eukaryotic cells. Variants in the POLG gene, which encodes the catalytic subunit of Pol-γ, can significantly impair Pol-γ enzyme function. Pol-γ-associated disorders are referred to as POLG-spectrum disorders (POLG-SDs) and are mainly autosomal-recessively inherited. Clinical manifestations include muscle weakness and fatigue, and severe forms of the disease can lead to premature death in infancy, childhood, and early adulthood, often associated with seizures, liver failure, or intractable epilepsy. Here, we analyzed fibroblasts from a compound heterozygous patient with the established pathogenic variant c.2419C>T; p.(Arg807Cys) and a previously undescribed variant c.678G>C; p.(Gln226His) with a clinical manifestation compatible with POLG-SDs, sensory ataxic neuropathy, and infantile muscular atrophy. We conducted a battery of functional studies for Pol-γ and mitochondrial dysfunction on the patient’s fibroblasts, to test whether the novel variant c.678G>C; p.(Gln226His) may be causative in human disease. Aims/Methods: We analyzed skin-derived fibroblasts in comparison to a first-degree relative (the mother of the patient), an asymptomatic carrier harboring only the established c.2419C>T; p.(Arg807Cys) mutation. Assessments of mitochondrial function included measurements of mtDNA content, mRNA levels of mitochondrial genes, mitochondrial mass, and mitochondrial morphology. Case Presentation and Results: A 13-year-old male presented with symptoms starting at three years of age, including muscle weakness and atrophy in the lower extremities and facial muscles, which later extended to the upper limbs, voice, and back muscles, without further progression. The patient also reported fatigue and muscle pain after physical activity, with no sensory deficits. Extensive diagnostic tests such as electromyography, nerve conduction studies, muscle biopsy, and MRI were unremarkable. Exome sequencing revealed that he carried the compound heterozygous variants in POLG c.678G>C; p.(Gln226His) and c.2419C>T; p.(Arg807Cys), but no other potential genetic pathogenic causes. In comparison to a first-degree relative (his mother) who only carried the c.2419C>T; p.(Arg807Cys) pathogenic mutation, in vitro analyses revealed a significant reduction in mtDNA content (~50%) and mRNA levels of mtDNA-encoded proteins. Mitochondrial mass was reduced by approximately 20%, and mitochondrial interconnectivity within cells was impaired, as determined by fluorescence microscopy and mitochondrial staining. Conclusions: Our findings suggest that the c.678G>C; p.(Gln226His) variant, in conjunction with the c.2419C>T; p.(Arg807Cys) mutation, may compromise mtDNA replication and mitochondrial function and could result in clinically significant mitochondriopathy. As this study is based on one patient compared to a first-degree relative (but with an identical mitochondrial genome), the pathogenicity of c.678G>C; p.(Gln226His) of POLG should be confirmed in future studies, in particular, in conjunction with other POLG-variants. Full article
24 pages, 3436 KiB  
Article
Transcription Factor Inhibition as a Potential Additional Mechanism of Action of Pyrrolobenzodiazepine (PBD) Dimers
by Julia Mantaj, Paul J. M. Jackson, Richard B. Parsons, Tam T. T. Bui, David E. Thurston and Khondaker Miraz Rahman
DNA 2025, 5(1), 8; https://doi.org/10.3390/dna5010008 - 5 Feb 2025
Viewed by 325
Abstract
Background: The pyrrolobenzodiazepine (PBD) dimer SJG-136 reached Phase II clinical trials in ovarian cancer and leukaemia in the UK and USA in the 2000s. Several structural analogues of SJG-136 are currently in clinical development as payloads for Antibody-Drug Conjugates (ADCs). There is growing [...] Read more.
Background: The pyrrolobenzodiazepine (PBD) dimer SJG-136 reached Phase II clinical trials in ovarian cancer and leukaemia in the UK and USA in the 2000s. Several structural analogues of SJG-136 are currently in clinical development as payloads for Antibody-Drug Conjugates (ADCs). There is growing evidence that PBDs exert their pharmacological effects through inhibition of transcription factors (TFs) in addition to arrest at the replication fork, DNA strand breakage, and inhibition of enzymes including endonucleases and RNA polymerases. Hence, PBDs can be used to target specific DNA sequences to inhibit TFs as a novel anticancer therapy. Objective: To explore the ability of SJG-136 to bind to the cognate sequences of transcription factors using a previously described HPLC/MS method, to obtain preliminary mechanistic evidence of its ability to inhibit transcription factors (TF), and to determine its effect on TF-dependent gene expression. Methods: An HPLC/MS method was used to assess the kinetics and thermodynamics of adduct formation between the PBD dimer SJG-136 and the cognate recognition sequence of the TFs NF-κB, EGR-1, AP-1, and STAT3. CD spectroscopy, molecular dynamics simulations, and gene expression analyses were used to rationalize the findings of the HPLC/MS study. Results: Notable differences in the rate and extent of adduct formation were observed with different DNA sequences, which might explain the variations in cytotoxicity of SJG-136 observed across different tumour cell lines. The differences in adduct formation result in variable downregulation of several STAT3-dependent genes in the human colon carcinoma cell line HT-29 and the human breast cancer cell line MDA-MB-231. Conclusions: SJG-136 can disrupt transcription factor-mediated gene expression, which contributes to its exceptional cytotoxicity in addition to the DNA-strand cleavage initiated by its ability to crosslink DNA. Full article
Show Figures

Figure 1

Figure 1
<p>Structures of the naturally occurring anthramycin (<b>A</b>), the C8-<span class="html-italic">bis</span>-pyrrole PBD Conjugate GWL-78 (<b>B</b>), the PBD 4-(1-methyl-1<span class="html-italic">H</span>-pyrrol-3-yl)benzenamine (MPB) conjugate KMR-28-39 (<b>C</b>), the C8/C8’-linked PBD dimer SJG-136 (<b>D</b>), the structurally related PBD dimer ADC payloads, Tesirine (<b>E</b>), and Talirine (<b>F</b>).</p>
Full article ">Figure 2
<p>(<b>A</b>) Schematic diagram of the mechanism of covalent binding of a PBD molecule to a guanine base; (<b>B</b>) Low-energy snapshot of a molecular model of the PBD dimer SJG-136 (green) covalently bound to G5 and G14 (purple/magenta) of the consensus sequence of the transcription factor EGR-1. DNA bases involved in the non-covalent interactions are shown as cyan sticks. Blue colour represents nitrogen atom. Dash lines represent hydrogen bonds.</p>
Full article ">Figure 3
<p>Structure of the hairpin oligonucleotides used in this study that contain the cognate sequences of the transcription factors NF-κB (two possible sequences NF-κB-1 and NF-κB-2), EGR-1, AP-1, and STAT3.</p>
Full article ">Figure 4
<p>The interaction of SJG–136 with NF–κB–1. (<b>A</b>) HPLC chromatogram of NF–κB–1 alone; (<b>B</b>) HPLC chromatogram of the SJG–136/NF–κB–1 adduct; (<b>C</b>) MALDI–TOF spectrum of the NF–κB–1 sequence alone; (<b>D</b>) MALDI–TOF spectrum of the SJG–136/NF–κB–1 adduct.</p>
Full article ">Figure 5
<p>Interaction of SJG–136 with NF–κB–2. (<b>A</b>) HPLC chromatogram of NF–κB–2 alone; (<b>B</b>) HPLC chromatogram of the SJG–136/NF–κB–2 adduct; (<b>C</b>) MALDI-TOF spectrum of NF–κB–2 alone; (<b>D</b>) MALDI–TOF spectrum of the SJG–136/NF–κB–2 adduct.</p>
Full article ">Figure 6
<p>Interaction of SJG–136 with EGR–1. (<b>A</b>) HPLC chromatogram of EGR–1 alone; (<b>B</b>) HPLC chromatogram of SJG–136/EGR–1 adduct; (<b>C</b>) MALDI–TOF spectrum of EGR–1 alone; (<b>D</b>) MALDI–TOF spectrum of the SJG–136/EGR–1 adduct.</p>
Full article ">Figure 7
<p>Interaction of SJG–136 with AP–1. (<b>A</b>) HPLC chromatogram of AP–1 alone; (<b>B</b>) HPLC chromatogram of the SJG–136/AP–1 adduct; (<b>C</b>) MALDI-TOF spectrum of the AP–1 sequence alone; (<b>D</b>) MALDI–TOF spectrum of the SJG–136/AP–1 adduct.</p>
Full article ">Figure 8
<p>Interaction of SJG–136 with the STAT3 sequence. (<b>A</b>) HPLC chromatogram of the STAT3 hairpin alone; (<b>B</b>) HPLC chromatogram of the SJG–136/STAT-3 adducts; (<b>C</b>) MALDI–TOF spectrum of the STAT3 hairpin alone; (<b>D</b>) MALDI–TOF spectrum of the SJG-136/STAT–3 adduct. The same mass was observed for all three adducts formed.</p>
Full article ">Figure 9
<p>Interaction of SJG–136 with the AP–1 hairpin sequence. (<b>A</b>) CD spectrum of the AP–1 sequence alone (black) and the AP–1/SJG–136 complex at t = 0 h; (<b>B</b>) CD spectrum of the AP–1 sequence alone (black) and the AP–1/SJG–136 complex at t = 24 h.</p>
Full article ">Figure 10
<p>Low-energy snapshots of molecular models of the interaction of SJG-136 with the NF-κB-1 hairpin. (<b>A</b>) Mono-Alkylated Adduct: SJG-136 (blue) covalently bound to G3 (purple/magenta) of the NF-κB-1 hairpin. The central methylene linker of SJG-136 forms extensive van der Waals interactions with the A4:T20 base pair (yellow), and the unreacted PBD forms non-covalent interactions with the A6:T18 base pair (cyan), allowing the molecule to fit isosterically in the DNA minor groove; (<b>B</b>) Interstrand cross-linked Adduct: SJG-136 (blue) covalently bound to both G2 and G19 (magenta) of the NF-κB-1 hairpin. The central methylene linker of SJG-136 forms extensive van der Waals interactions with the A4:T20 base pair (yellow) with stabilising hydrogen bonds between the N10-proton of one PBD moiety and the ring nitrogen (N3) of the adjacent G3, and between the N10-proton of the other PBD moiety and the O4 atom of the neighbouring T20 base.</p>
Full article ">Figure 11
<p>The effect of SJG-136 on the expression of (<b>A</b>) STAT3-dependent genes in MDA-MB-231 cells and (<b>B</b>) AP-1-dependent genes in HT-29 cells expressed as fold-decrease. Experiments were performed in triplicates (<span class="html-italic">n</span> = 3). All data are mean ± SD. * = <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, NS = not significant.</p>
Full article ">
13 pages, 1724 KiB  
Article
A Novel Genotype of Orientia tsutsugamushi in Human Cases of Scrub Typhus from Southeastern India
by Krishnamoorthy Nallan, Bhuvaneshwari Chinnathambi Kalidoss, Eunice Swarna Jacob, Samyuktha Krishnasamy Mahadevan, Steny Joseph, Ramkumar Ramalingam, Govindarajan Renu, Balaji Thirupathi, Balajinathan Ramasamy, Bhavna Gupta, Manju Rahi and Paramasivan Rajaiah
Microorganisms 2025, 13(2), 333; https://doi.org/10.3390/microorganisms13020333 - 4 Feb 2025
Viewed by 388
Abstract
Scrub typhus is a mite-borne, re-emerging public health problem in India, particularly in Tamil Nadu, South India. More than 40 serotypes of Orientia tsutsugamushi have been documented worldwide. However, the information on the circulation of its molecular sub-types in India is scanty. A [...] Read more.
Scrub typhus is a mite-borne, re-emerging public health problem in India, particularly in Tamil Nadu, South India. More than 40 serotypes of Orientia tsutsugamushi have been documented worldwide. However, the information on the circulation of its molecular sub-types in India is scanty. A retrospective study was conducted among serologically confirmed cases of scrub typhus. DNA isolated from blood was screened by a nested polymerase chain reaction (nPCR) targeting the GroEL and the 56 kDa type-specific antigen (TSA) genes. Out of 59 samples, 14 partial fragments of GroEL and the twelve 56 kDa genes were PCR-amplified and DNA-sequenced. The neighbor-joining (NJ) analysis indicated three distinct phylogenetic clades, including a novel genotype designated as Ot-Thanjavur-Tamil Nadu (Ot-TJTN, 9 nos. 64.3%); Karp-like (4 nos. 28.6%); and Kuroki-Gilliam type (1 no. 7.1%). Also, phylogenetic analysis of twelve 56 kDa variable domains (VDΙ-ΙΙΙ) of TSA gene sequences revealed a distinctive new genotypic cluster of eight samples (66.6%), and the remaining four (33.4%) were Karp-like genotypes. The Simplot analysis for the similarity and event of recombination testing elucidated the existence of the new genotype of the Ot-TJTN cluster, which was undescribed so far, in the Kato and TA716 lineages. The significant findings recommend further studies to understand the ongoing transmission dynamics of different O. tsutsugamushi strains in vector mites, rodent hosts, and humans in this region. Full article
(This article belongs to the Special Issue The Molecular Epidemiology of Infectious Diseases)
Show Figures

Figure 1

Figure 1
<p>Neighbor-joining (NJ) tree of the GroEL (<b>a</b>) and 56 kDa TSA gene (<b>b</b>) with 1000 bootstrap replicates, K2P substitution model. The 56 kDa gene (<b>b</b>) with reference sequences showing two clusters, a novel TJTN and Karp-like. Sequences obtained from this study were labeled with a blue circle; the out-group sequence is marked with a red triangle for GroEL (<b>a</b>).</p>
Full article ">Figure 2
<p>A radiation distance phylogenetic tree of GroEL gene sequences showing the three different clusters, and the designated new <span class="html-italic">Ot</span>-TJTN is clustered (red) separately, deviating from the Karp, Kato, and TA716 (blue) and Gilliam clade (grey).</p>
Full article ">Figure 3
<p>A radiation distance phylogenetic tree of the 56 kDa TSA sequences showing the two different clusters of designated new <span class="html-italic">Ot</span>-TJTN and Karp-like genotypes. Kato and TA716 were ancestral lineage to the novel <span class="html-italic">Ot</span>-TJTN genotype (blue) and Karp-like sequences (red), and the Gilliam and Kuroki genotypes formed a separate cluster. (Square in red: sequences from this study and circle in green: reference sequences).</p>
Full article ">Figure 4
<p>Multiple sequence alignment of the Karp-like and <span class="html-italic">Ot</span>-TJTN 56 kDa TSA amino acid sequences generated in this study with the closely related reference Kato and TA716 genotypes. The alignment shows the non-synonymous change of amino acids in 56 kDa protein VD-ΙI-III due to the nucleotide substitutions.</p>
Full article ">
14 pages, 1523 KiB  
Review
The p12 Subunit Choreographs the Regulation and Functions of Two Forms of DNA Polymerase δ in Mammalian Cells
by Dazhong Xu, Selvaraj Ayyamperumal, Sufang Zhang, Jinjin Chen, Ernest Y. C. Lee and Marietta Y. W. T. Lee
Genes 2025, 16(2), 188; https://doi.org/10.3390/genes16020188 - 3 Feb 2025
Viewed by 451
Abstract
There are two forms of DNA polymerase δ in human cells, Pol δ4 and Pol δ3, which differ based on their possession of the p12 subunit. The degradation of p12 has emerged as an important regulatory mechanism that controls the generation of Pol [...] Read more.
There are two forms of DNA polymerase δ in human cells, Pol δ4 and Pol δ3, which differ based on their possession of the p12 subunit. The degradation of p12 has emerged as an important regulatory mechanism that controls the generation of Pol δ3. The underlying importance of this system lies in the altered enzymatic properties of the two forms of Pol δ engendered by the influence of p12. We briefly review how the balance of these two forms is regulated through the degradation of p12. We focus on the roles of Pol δ4, whose cellular functions are less well known. This is significant because recent studies show that this is the form engaged in the homology-dependent repair of double-strand breaks. We consider new horizons for future research into this system and their potential involvement in tumorigenesis. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p><b>The degradation of p12 controls the interplay between Pol δ4 and Pol δ3.</b> The left arm of the diagram gives a concise summary of our previous studies [<a href="#B1-genes-16-00188" class="html-bibr">1</a>,<a href="#B2-genes-16-00188" class="html-bibr">2</a>,<a href="#B3-genes-16-00188" class="html-bibr">3</a>,<a href="#B4-genes-16-00188" class="html-bibr">4</a>,<a href="#B58-genes-16-00188" class="html-bibr">58</a>,<a href="#B59-genes-16-00188" class="html-bibr">59</a>]. The regulated degradation of p12 is the key mechanism for the generation of Pol δ3. These studies also support an essential role for Pol δ3 in lagging-strand DNA synthesis and provide evidence regarding the differential roles of Pol δ4 in HDR. The delineation of the functions of Pol δ4 is still incomplete, and it may serve specific roles in BIR (break-induced replication) and MiDAS (mitotic DNA replication). The question marks are present to indicate that these are important areas for further study, especially regarding their comparative contributions to these processes.</p>
Full article ">Figure 2
<p><b>Pol δ4 but not Pol δ3 exhibits stand displacement activity.</b> (<b>A</b>–<b>D</b>). Diagrammatic view of the ability of Pol δ4 and Pol δ3 to perform strand displacement on model oligonucleotide substrates and the processing of Okazaki fragments in concert with Fen 1 and DNA ligase using model oligonucleotide substrates [<a href="#B4-genes-16-00188" class="html-bibr">4</a>,<a href="#B69-genes-16-00188" class="html-bibr">69</a>]. (<b>E</b>). Displacement synthesis in a model D-loop (extension of invading strand). Pol δ4, but not Pol δ3, is predicted to be able to perform D-loop extension [<a href="#B58-genes-16-00188" class="html-bibr">58</a>].</p>
Full article ">Figure 3
<p><b>The degradation of p12 in mammalian cells containing Pol δ4 to cells containing Pol δ3 choreographs their participation in DNA repair and replication processes.</b> (<b>A</b>) p12 degradation acts as a mechanism that dictates the two central functions in DNA replication and DNA repair. This diagram provides an overview of the cellular regulation of the degradation of p12, which converts the Pol δ4 tetramer to the Pol δ3 trimer. The removal of p12 leaves behind the Pol δ3 enzyme in vivo via the regulated destruction of the p12 subunit. Thus, cells containing Pol δ4 are converted to cells containing Pol δ3, a phenomenon which occurs under two scenarios. The first is the presence of genotoxic or replication stress (left-hand column). This occurs in an ATR-dependent manner by proteasomal degradation mediated by E3 ubiquitin ligases. Multiple E3 ubiquitin ligases may be involved, including RNF8 and CRL4<sup>Cdt2</sup>. This mechanism ensures that Pol δ3 is the operative form of Pol δ activity for gap-filling in excision repair (NER, MMR). The second trigger for p12 degradation (right-hand column) is integrally embedded in the cell cycle regulation of the initiation of DNA synthesis. Here, p12 degradation is ubiquitinylated and targeted for proteasomal degradation by CRL4<sup>Cdt2</sup>, which plays a key role in the destruction of licensing factors Cdt1, Set8, and p21 to prevent the re-licensing of the origins. Thus, cells in the S phase contain only Pol δ3, which is the lagging-strand polymerase in DNA replication. The system acts as a flip–flop switch between two cellular states where either Pol δ4 or Pol δ3 is the primary polymerase. This also constrains Pol δ4 from acting when Pol δ3’s function is operative. (<b>B</b>). The p12 subunit binds to the p125 subunit and thereby acts to modulate Pol δ4’s activity. This is of consequence because Pol δ4 possesses functions in DNA transactions that are not shared by Pol δ3. These functions represent “gain of function” attributes that expand the cellular repertoire of Pol δ activity. Notably, Pol δ4, but not Pol δ3, is the primary form engaged in the HDR repair of double-stranded DNA breaks. Thus, the involvement of p12/Pol δ4 in human DNA synthesis in homologous recombination is also subject to control by the p12 regulatory switch shown in (<b>A</b>).</p>
Full article ">
27 pages, 1435 KiB  
Opinion
60 Years of Studies into the Initiation of Chromosome Replication in Bacteria
by John Herrick, Vic Norris and Masamichi Kohiyama
Biomolecules 2025, 15(2), 203; https://doi.org/10.3390/biom15020203 - 1 Feb 2025
Viewed by 503
Abstract
The Replicon Theory has guided the way experiments into DNA replication have been designed and interpreted for 60 years. As part of the related, explanatory package guiding experiments, it is thought that the timing of the cell cycle depends in some way on [...] Read more.
The Replicon Theory has guided the way experiments into DNA replication have been designed and interpreted for 60 years. As part of the related, explanatory package guiding experiments, it is thought that the timing of the cell cycle depends in some way on a critical mass for initiation, Mi, as licensed by a variety of macromolecules and molecules reflecting the state of the cell. To help in the re-interpretation of this data, we focus mainly on the roles of DnaA, RNA polymerase, SeqA, and ribonucleotide reductase in the context of the “nucleotypic effect”. Full article
(This article belongs to the Collection Molecular Biology: Feature Papers)
Show Figures

Figure 1

Figure 1
<p>Loop-back model (adapted from [<a href="#B52-biomolecules-15-00203" class="html-bibr">52</a>]): <span class="html-italic">oriC</span> containing DUE (duplex unwinding element), DnaA boxes (R1, R5M, R2, and R4), IHF (and the IHF-binding site), and Fis (and the Fis-binding site) is folded by IHF facilitating interactions between DnaA and DUE.</p>
Full article ">Figure 2
<p>Mother cell licensing model.</p>
Full article ">Figure 3
<p>Model of the nucleotypic effect (adapted from [<a href="#B168-biomolecules-15-00203" class="html-bibr">168</a>]).</p>
Full article ">Figure 4
<p>Model of DnaA:RNR homeostatic pair (adapted from [<a href="#B112-biomolecules-15-00203" class="html-bibr">112</a>]). Expression of the <span class="html-italic">dnaA</span> gene (black) and the expression of the <span class="html-italic">nrdA</span> gene (red) oscillate out of phase in the first half of the cell cycle (see Extended Data Figure 6 in [<a href="#B112-biomolecules-15-00203" class="html-bibr">112</a>]): <span class="html-italic">dnaA</span> expression increases after cell division (div) when <span class="html-italic">nrdA</span> expression is low. Low levels of newly formed DnaA-ATP inversely correlate with high levels of <span class="html-italic">dnaA</span> expression (due to autoregulation) and stimulate <span class="html-italic">nrdA</span> expression. As DnaA levels increase, DnaA inhibits <span class="html-italic">nrdA</span> expression [<a href="#B110-biomolecules-15-00203" class="html-bibr">110</a>]. Consequently, DnaA protein will repress its own expression at or after initiation and will concomitantly limit RNR production by repressing <span class="html-italic">nrdA</span> expression during the S phase/C period, presumably to a level of expression that is in equilibrium with the number of active replication forks. DnaA and RNR thus act as a homeostatic pair to regulate the number of forks according to growth rate. The cytoplasm in <span class="html-italic">C. crescentus</span> is reduced in G1. At the G1/S transition (initiation of DNA replication), the cytoplasm switches to a highly oxidized state that becomes increasingly reduced during the S phase until G2 when the cytoplasm switches back to a reduced state [<a href="#B149-biomolecules-15-00203" class="html-bibr">149</a>,<a href="#B153-biomolecules-15-00203" class="html-bibr">153</a>,<a href="#B156-biomolecules-15-00203" class="html-bibr">156</a>]. This gated redox cycle also occurs in yeast and metazoan cells. It has been proposed to be a universal feature of cell cycle control [<a href="#B149-biomolecules-15-00203" class="html-bibr">149</a>,<a href="#B156-biomolecules-15-00203" class="html-bibr">156</a>] and, therefore, might explain the coordination between cell growth/mass and the major cell cycle events, including the initiation of DNA replication, which coincides with a relatively constant <span class="html-italic">Mi</span>. Steps in the hypothetical relationship between licensing in the mother cell and initiation of replication in daughter cells. (1) DnaA binds <span class="html-italic">oriC</span> immediately after sequestration leading to its full licensing in late C- or D-periods; (2) licensing in the mother cell titrates DnaA prior to cell division; (3) low levels of DnaA at mother cell division and daughter cell birth induce <span class="html-italic">dnaA</span> expression in the B-period of the daughter cell cycle; (4) low levels of DnaA in the B-period of the daughter cell stimulate expression of <span class="html-italic">nrdAB</span> prior to initiation (late B- and early C-period); (5) high levels of DnaA in the early C-period repress <span class="html-italic">nrdAB</span> and limit RNR levels post-initiation (<span class="html-italic">nrdAB</span> expression is not necessary for elongation, but its expression is absolutely necessary for initiation, underlining the important role dNTP synthesis plays in the cell cycle regulation of replication initiation) [<a href="#B159-biomolecules-15-00203" class="html-bibr">159</a>,<a href="#B169-biomolecules-15-00203" class="html-bibr">169</a>]. An equilibrium (homeostasis) is established between rates of dNTP synthesis and the number of active forks, an equilibrium that serves to protect the cell against the mutagenic effects of abnormally high levels of RNR and/or rates of dNTP production (in conjunction with dATP allosteric regulation of RNR and NrdR repression of <span class="html-italic">nrdAB</span> [<a href="#B175-biomolecules-15-00203" class="html-bibr">175</a>]). It should be noted that during fast growth, there is no B-period. DnaA, therefore, licenses <span class="html-italic">oriC</span> in the mother cell (immediately after sequestration), and DnaA:RNR activates initiation prior to cell division. Black curve: <span class="html-italic">dnaA</span> gene expression level; red curve: <span class="html-italic">nrdAB</span> expression level.</p>
Full article ">
11 pages, 5901 KiB  
Article
Quantitative Real-Time Polymerase Chain Reaction (PCR) Assay for Rapid Monitoring of the Harmful Algal Bloom Species Cochlodinium polykrikoides
by Min-Jeong Kim, Hyun-Jung Kim, Joon Sang Park, Donhyug Kang, Sungho Cho, Hansoo Kim, Seung Ho Baek, Jordan Jun Chul Park, Jeonghoon Han, Kang Eun Kim and Seung Won Jung
J. Mar. Sci. Eng. 2025, 13(2), 277; https://doi.org/10.3390/jmse13020277 (registering DOI) - 31 Jan 2025
Viewed by 336
Abstract
Harmful blooms of the dinoflagellate Cochlodinium polykrikoides (Margalefidinium polykrikoides) had detrimental aquacultural and economic effects globally, and to reduce the damage caused by these blooms, early biomonitoring and quantitative analysis of C. polykrikoides are of the utmost importance. Here, for the [...] Read more.
Harmful blooms of the dinoflagellate Cochlodinium polykrikoides (Margalefidinium polykrikoides) had detrimental aquacultural and economic effects globally, and to reduce the damage caused by these blooms, early biomonitoring and quantitative analysis of C. polykrikoides are of the utmost importance. Here, for the detection of C. polykrikoides using quantitative real-time polymerase chain reactions, we developed specific primers targeting the large subunit ribosomal DNA (LSU rDNA) and evaluated their applicability in the field during the occurrence of a C. polykrikoides bloom. The specific primers not only accurately detected C. polykirkoides but also had a detection performance comparable with that obtained using microscopic observations. Accordingly, we developed a system that can be used in the field and applied when red tides occur, with accurate results being obtained more than five times more rapidly than those obtained based on microscopic analysis. Collectively, our findings indicate that the C. polykrikoides bloom detection system developed in this study can be applied to rapidly detect and accurately quantify C. polykrikoides in environmental samples. Data obtained using this system could be used as a basis for developing prompt monitoring and warning systems for the early detection of C. polykrikoides blooms in the field. Full article
Show Figures

Figure 1

Figure 1
<p><span class="html-italic">Cochlodinium polykrikoides</span> morphology, specific primer design, and standard curve construction. (<b>a</b>) Light microscopic and (<b>b</b>) scanning electron microscopic micrographs of <span class="html-italic">C. polykrikoides</span>, (<b>c</b>) <span class="html-italic">C. polykrikoides</span>-specific primer target sites, (<b>d</b>) a cell number-based standard curve, and (<b>e</b>) a copy number-based standard curve.</p>
Full article ">Figure 2
<p>Schematic diagram of the <span class="html-italic">Cochlodinium polykrikoides</span> field detection system.</p>
Full article ">Figure 3
<p>Field application of a qPCR monitoring system for <span class="html-italic">Cochlodinium polykrikoides</span> red tides. (<b>a</b>,<b>b</b>) Photographs of <span class="html-italic">C. polykrikoides</span> mitigation using natural yellow clay. Contour charts of <span class="html-italic">C. polykrikoides</span> red tide occurrence based on microscopic (<b>c</b>) and qPCR (<b>d</b>) analyses. (<b>e</b>) Comparison of <span class="html-italic">C. polykrikoides</span> red tide cell counts in the field using microscopic analysis and qPCR methods.</p>
Full article ">
16 pages, 2795 KiB  
Article
Mitochondria-Derived Vesicles and Inflammatory Profiles of Adults with Long COVID Supplemented with Red Beetroot Juice: Secondary Analysis of a Randomized Controlled Trial
by Emanuele Marzetti, Hélio José Coelho-Júnior, Riccardo Calvani, Giulia Girolimetti, Riccardo Di Corato, Francesca Ciciarello, Vincenzo Galluzzo, Clara Di Mario, Barbara Tolusso, Luca Santoro, Ottavia Giampaoli, Alberta Tomassini, Walter Aureli, Matteo Tosato, Francesco Landi, Cecilia Bucci, Flora Guerra and Anna Picca
Int. J. Mol. Sci. 2025, 26(3), 1224; https://doi.org/10.3390/ijms26031224 - 30 Jan 2025
Viewed by 443
Abstract
In a recent clinical trial, beetroot juice supplementation for 14 days yielded positive effects on systemic inflammation in adults with long COVID. Here, we explored the relationship between circulating markers of mitochondrial quality and inflammation in adults with long COVID as well as [...] Read more.
In a recent clinical trial, beetroot juice supplementation for 14 days yielded positive effects on systemic inflammation in adults with long COVID. Here, we explored the relationship between circulating markers of mitochondrial quality and inflammation in adults with long COVID as well as the impact of beetroot administration on those markers. We conducted secondary analyses of a placebo-controlled randomized clinical trial testing beetroot juice supplementation as a remedy against long COVID. Analyses were conducted in 25 participants, 10 assigned to placebo (mean age: 40.2 ± 11.5 years, 60% women) and 15 allocated to beetroot juice (mean age: 38.3 ± 7.7 years, 53.3% women). Extracellular vesicles were purified from serum by ultracentrifugation and assayed for components of the electron transport chain and mitochondrial DNA (mtDNA) by Western blot and droplet digital polymerase chain reaction (ddPCR), respectively. Inflammatory markers and circulating cell-free mtDNA were quantified in serum through a multiplex immunoassay and ddPCR, respectively. Beetroot juice administration for 14 days decreased serum levels of interleukin (IL)-1β, IL-8, and tumor necrosis factor alpha, with no effects on circulating markers of mitochondrial quality control. Significant negative associations were observed between vesicular markers of mitochondrial quality control and the performance on the 6 min walk test and flow-mediated dilation irrespective of group allocation. These findings suggest that an amelioration of mitochondrial quality, possibly mediated by mitochondria-derived vesicle recycling, may be among the mechanisms supporting improvements in physical performance and endothelial function during the resolution of long COVID. Full article
(This article belongs to the Special Issue Advances in Anti-Aging Treatment Development, 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Changes from baseline to day 14 in serum concentrations of inflammatory markers in the placebo (<span class="html-italic">n</span> = 10) and beetroot juice (<span class="html-italic">n</span> = 15) groups. Abbreviations: IFN-γ, interferon gamma; IL, interleukin; IL-1ra, interleukin 1 receptor antagonist; and TNF-α, tumor necrosis factor alpha. * <span class="html-italic">p</span> &lt; 0.05 vs. baseline.</p>
Full article ">Figure 2
<p>Changes from baseline to day 14 in vesicular/mitochondrial markers in the placebo (<span class="html-italic">n</span> = 10) and beetroot juice (<span class="html-italic">n</span> = 15) groups. The lower panels show representative Western blot bands and transmission electron microscope images of extracellular vesicles. Abbreviations: ATP5A, ATP synthase F1 subunit alpha; a.u., arbitrary unit; B, beetroot juice; ccf-mtDNA, circulating cell-free mitochondrial DNA; CD, cluster of differentiation; Ctrl, control; EVs, extracellular vesicles; MTCOI, mitochondrial cytochrome C oxidase I; OD, optical density; NDUFB8, NADH:ubiquinone oxidoreductase subunit B8; RPS6, ribosomal protein s6; and SDHB, succinate dehydrogenase [ubiquinone] iron–sulfur subunit.</p>
Full article ">Figure 3
<p>Changes from baseline to day 14 in inflammatory and vesicular/mitochondrial markers adjusted for pre-intervention values in (<b>A</b>) placebo (<span class="html-italic">n</span> = 10) and (<b>B</b>) beetroot juice (<span class="html-italic">n</span> = 15) groups. Abbreviations: ATP5A, ATP synthase F1 subunit alpha; CD, cluster of differentiation; ccf-mtDNA, circulating cell-free mitochondrial DNA; EVs, extracellular vesicles; IL, interleukin; IL-1ra, interleukin 1 receptor antagonist; MTCOI, mitochondrial cytochrome C oxidase I; NDUFB8, NADH:ubiquinone oxidoreductase subunit B8; SDHB, succinate dehydrogenase [ubiquinone] iron–sulfur subunit; and vs-mtDNA, vesicular mitochondrial DNA. * <span class="html-italic">p</span> &lt; 0.05 for comparisons between pre- and post-intervention adjusted for baseline values according to analysis of covariance (ANCOVA).</p>
Full article ">Figure 4
<p>Correlation analyses between inflammatory and vesicular/mitochondrial markers and measures of physical and endothelial function at 14 days in the whole study sample (<span class="html-italic">n</span> = 25). Abbreviations: ATP5A, ATP synthase F1 subunit alpha; CD, cluster of differentiation; ccf-mtDNA, circulating cell-free mitochondrial DNA; EVs, extracellular vesicles; FMD, flow-mediated dilation; IFN-γ, interferon gamma; IL, interleukin; IL-1ra, interleukin 1 receptor antagonist; MTCOI, mitochondrial cytochrome C oxidase I; NDUFB8, NADH:ubiquinone oxidoreductase subunit B8; SDHB, succinate dehydrogenase [ubiquinone] iron–sulfur subunit; and vs-mtDNA, vesicular mitochondrial DNA. * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 5
<p>Correlation analyses between pre/post-intervention differences in inflammatory and vesicular/mitochondrial markers and measures of physical and endothelial function at 14 days in (<b>A</b>) the placebo group (<span class="html-italic">n</span> = 10) and (<b>B</b>) participants assigned to beetroot juice (<span class="html-italic">n</span> = 15). Abbreviations: ATP5A, ATP synthase F1 subunit alpha; CD, cluster of differentiation; ccf-mtDNA, circulating cell-free mitochondrial DNA; EVs, extracellular vesicles; IFN-γ, interferon gamma; IL, interleukin; IL-1ra, interleukin 1 receptor antagonist; MTCOI, mitochondrial cytochrome C oxidase I; NDUFB8, NADH:ubiquinone oxidoreductase subunit B8; and SDHB, succinate dehydrogenase [ubiquinone] iron–sulfur subunit; vs-mtDNA, vesicular mitochondrial DNA. * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 6
<p>Correlation analyses between pre/post-intervention differences in inflammatory and vesicular/mitochondrial markers and measures of physical and endothelial function at 14 days independent of intervention group (<span class="html-italic">n</span> = 25). Abbreviations: ATP5A, ATP synthase F1 subunit alpha; CD, cluster of differentiation; ccf-mtDNA, circulating cell-free mitochondrial DNA; EVs, extracellular vesicles; IFN-γ, interferon gamma; IL, interleukin; IL-1ra, interleukin 1 receptor antagonist; MTCOI, mitochondrial cytochrome C oxidase I; NDUFB8, NADH:ubiquinone oxidoreductase subunit B8; SDHB, succinate dehydrogenase [ubiquinone] iron–sulfur subunit; vs-mtDNA, vesicular mitochondrial DNA * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">
45 pages, 6031 KiB  
Review
The Theoretical Basis of qPCR and ddPCR Copy Number Estimates: A Critical Review and Exposition
by James N. McNair, Daniel Frobish, Richard R. Rediske, John J. Hart, Megan N. Jamison and David C. Szlag
Water 2025, 17(3), 381; https://doi.org/10.3390/w17030381 - 30 Jan 2025
Viewed by 451
Abstract
The polymerase chain reaction (PCR) is a molecular biology tool with diverse applications in the aquatic sciences. Classical PCR is a nonquantitative method that can be used to detect target DNA sequences that are characteristic of particular microbial taxa but cannot determine their [...] Read more.
The polymerase chain reaction (PCR) is a molecular biology tool with diverse applications in the aquatic sciences. Classical PCR is a nonquantitative method that can be used to detect target DNA sequences that are characteristic of particular microbial taxa but cannot determine their concentrations in water samples. Various quantitative forms of PCR have been developed to remove this limitation. Of these, the two that currently are used most widely are real-time quantitative PCR (qPCR) and droplet digital PCR (ddPCR). Several outlines of the mathematical and statistical basis of these methods for estimating target sequence concentrations are available in the literature, but we are aware of no thorough and rigorous derivation of the theoretical underpinnings of either. The purpose of this review is to provide such derivations, and to identify and compare the main strengths and weaknesses of the two methods. We find that both estimation methods are sound, provided careful attention is paid to specific details that differ between the two. With qPCR, it is especially important to reduce any significant PCR inhibition by sample constituents and to properly fit the standard curve to heteroskedastic calibration data. With ddPCR, it is important to ensure that the value of the mean droplet volume used in calculating concentrations is correct for the particular combination of droplet generator and master mix used. The advantages of qPCR include lower instrument and per-sample costs, a shorter turnaround time for obtaining results, a higher upper limit of quantification, and a wider dynamic range. The advantages of ddPCR include freedom from dependence on a standard curve, an inherently lower sensitivity to PCR inhibitors, a lower limit of quantification, a simpler theoretical basis, and simpler data analysis. We suggest qPCR often will be preferable in laboratory studies where investigators have significant control over the range of target sequence concentrations in samples, concentrations are sufficiently high so proper calibration does not require standards with concentrations low enough to exhibit exaggerated variability in the threshold cycle, and no significant inhibition is present, or more generally, in studies where funding levels do not permit the higher cost of instrumentation and supplies required by ddPCR or where the shorter turnaround time for qPCR is essential. If sufficient funds are available, ddPCR often will be preferable when the ability to quantify low concentrations is important, especially if inhibitors are likely to be present at concentrations that are problematic for qPCR. Full article
23 pages, 8693 KiB  
Article
DNA-Binding Activities of KSHV DNA Polymerase Processivity Factor (PF-8) Complexes
by Jennifer Kneas Travis, Megan Martin and Lindsey M. Costantini
Viruses 2025, 17(2), 190; https://doi.org/10.3390/v17020190 - 29 Jan 2025
Viewed by 375
Abstract
Kaposi’s Sarcoma Herpesvirus (KSHV) is the causative agent of several human diseases. There are few effective treatments available to treat infection and KSHV oncogenesis. Disrupting the KSHV infectious cycle would diminish the viral spread. The KSHV lytic phase and production of new virions [...] Read more.
Kaposi’s Sarcoma Herpesvirus (KSHV) is the causative agent of several human diseases. There are few effective treatments available to treat infection and KSHV oncogenesis. Disrupting the KSHV infectious cycle would diminish the viral spread. The KSHV lytic phase and production of new virions require efficient copying and packaging of the KSHV genome. KSHV encodes its own lytic DNA replication machinery, including the processivity factor (PF-8), which presents itself as an attractive target for antiviral development. We characterized PF-8 at the single molecule level using transmission electron microscopy to identify key molecular interactions that mediate viral DNA replication initiation. Our results indicate that PF-8 forms oligomeric ring structures (tetramer, hexamer, and/or dodecamer) similar to the related Epstein–Barr virus processivity factor (BMRF1). Our DNA positional mapping revealed high-frequency binding locations of PF-8 within the lytic origin of replication (OriLyt). A multi-variable analysis of PF-8 DNA-binding activity with three mutant OriLyts provides new insights into the mechanisms that PF-8 associates with viral DNA and complexes to form multi-ring-like structures. Collectively, these data enhance the mechanistic understanding of the molecular interactions (protein–protein and protein-DNA) of an essential KSHV DNA replication protein. Full article
(This article belongs to the Special Issue Molecular and Cellular Biology of Human Oncogenic Viruses)
Show Figures

Graphical abstract

Graphical abstract
Full article ">
12 pages, 4045 KiB  
Article
Analysis of Short Tandem Repeat Expansions in a Cohort of 12,496 Exomes from Patients with Neurological Diseases Reveals Variable Genotyping Rate Dependent on Exome Capture Kits
by Clarissa Rocca, David Murphy, Chris Clarkson, Matteo Zanovello, Delia Gagliardi, Queen Square Genomics, Rauan Kaiyrzhanov, Javeria Alvi, Reza Maroofian, Stephanie Efthymiou, Tipu Sultan, Jana Vandrovcova, James Polke, Robyn Labrum, Henry Houlden and Arianna Tucci
Genes 2025, 16(2), 169; https://doi.org/10.3390/genes16020169 - 28 Jan 2025
Viewed by 597
Abstract
Background/Objectives: Short tandem repeat expansions are the most common cause of inherited neurological diseases. These disorders are clinically and genetically heterogeneous, such as in myotonic dystrophy and spinocerebellar ataxia, and they are caused by different repeat motifs in different genomic locations. Major advances [...] Read more.
Background/Objectives: Short tandem repeat expansions are the most common cause of inherited neurological diseases. These disorders are clinically and genetically heterogeneous, such as in myotonic dystrophy and spinocerebellar ataxia, and they are caused by different repeat motifs in different genomic locations. Major advances in bioinformatic tools used to detect repeat expansions from short read sequencing data in the last few years have led to the implementation of these workflows into next generation sequencing pipelines in healthcare. Here, we aimed to evaluate the clinical utility of analysing repeat expansions through exome sequencing in a large cohort of genetically undiagnosed patients with neurological disorders. Methods: We here analyse 27 disease-causing DNA repeats found in the coding, intronic and untranslated regions in 12,496 exomes in patients with a range of neurogenetic conditions. Results: We identified—and validated by polymerase chain reaction—29 repeat expansions across a range of loci, 48% (n = 14) of which were diagnostic. We then analysed the genotyping performance across all repeat loci and found that, despite high coverage in most repeats in coding regions, some loci had low genotyping rates, such as those that cause spinocerebellar ataxia 2 (ATXN2, 0.1–8.4%) and Huntington disease (HTT, 0.2–58.2%), depending on the capture kit. Conversely, while most intronic repeats were not genotyped, we found a high genotyping rate in the intronic locus that causes spinocerebellar ataxia 36 (NOP56, 30.1–98.3%) and in the one that causes myotonic dystrophy type 1 (DMPK, myotonic dystrophy type 1). Conclusions: We show that the key factors that influence the genotyping rate of repeat expansion loci analysis are the sequencing read length and exome capture kit. These results provide important information about the performance of exome sequencing as a genetic test for repeat expansion disorders. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic overview of the study workflow.</p>
Full article ">Figure 2
<p>Cohort overview and study design. The map illustrates the global distribution of 12,496 cases included in the cohort, with participant numbers represented by coloured circles: Europe (N = 8649, blue), East Asia (N = 1602, yellow), Africa (N = 404, red), America (N = 334, dark red), and South Asia (N = 68, green). The right panel provides the demographic information and diagnostic categories included in the analysis. The study design is summarised in the blue boxes at the bottom.</p>
Full article ">Figure 3
<p>Total number of repeat expansions identified by EH, visual inspection and PCR validation. (<b>A</b>) 365 repeat expansions identified by EH with the visual inspection outcome. Loci are divided into three groups: coding, intron and UTR. Green bars represent calls that passed visual inspection, yellow bars are for calls that were categorised in the “borderline” group and red bars indicate samples that failed visual inspection. Loci that do not have a bar next to them did not have any expanded calls predicted by EH. (<b>B</b>) The outcome of PCR-tested samples. The light blue bars indicate samples that tested positive for PCR, while the pink bars represent samples that tested negative. Stripes indicate cases that were in the visual inspection “Pass” category, whereas dots represent cases that were “borderline” after visual inspection.</p>
Full article ">Figure 4
<p>Pedigree of SCA3 family and MRI scan of proband. The red arrow shows the proband. (<b>A</b>) Square = male; circle = female; black filled symbol = affected individual; white symbols = unaffected individuals; diagonal line = deceased individual. Double lines indicate consanguinity. (<b>B</b>) MRI scan of patient IV.8. The red arrow indicates cerebellar atrophy.</p>
Full article ">Figure 5
<p>Targeted loci and coverage according to the four most used exome sequencing kits in this cohort. (<b>A</b>) The RED loci are categorised based on their genomic location: coding, intron and UTR. Target (purple): the specific region of the gene is targeted by the exome kit. Not target (yellow): the region of interest is not covered by the exome kit. The percentage indicates how much of the region is not covered. For example, in <span class="html-italic">ATN1</span>, 60% of the region of interest is not covered by the SureSelect V4 kit. When not specified, the percentage of target or not target is 0%. The exome sequencing kits are represented by different bars: SureSelect V6, SureSelect V4, Nextera and TruSeq. The dashed lines under each group indicate the total number of RED loci analysed in each category: 12 coding, 7 intronic and 8 UTRs. (<b>B</b>) Heatmap showing the coverage of the analysed RED loci across different genomic regions. Coverage is represented by the number of sequencing reads mapping to each locus, as indicated by the colour scale. (<b>C</b>) 3D plots of the genotyping rate for EH-generated calls by read length and sequencing kit. The three plots show EH calls in coding, intron and UTR loci. In each plot, calls are divided by locus and read length. The four different colours represent the different exome capture kits used.</p>
Full article ">
11 pages, 817 KiB  
Article
Sevoflurane Preconditioning Rescues PKMζ Gene Expression from Broad Hypoxia-Induced mRNA Downregulation Correlating with Improved Neuronal Recovery
by Joan Y. Hou, Kim D. Allen, A. Iván Hernandez, James E. Cottrell and Ira S. Kass
NeuroSci 2025, 6(1), 9; https://doi.org/10.3390/neurosci6010009 - 28 Jan 2025
Viewed by 569
Abstract
Hypoxia due to stroke is a major cause of neuronal damage, leading to loss of cognition and other brain functions. Sevoflurane preconditioning improves recovery after hypoxia. Hypoxia interferes with protein expression at the translational level; however, its effect on mRNA levels for neuronal [...] Read more.
Hypoxia due to stroke is a major cause of neuronal damage, leading to loss of cognition and other brain functions. Sevoflurane preconditioning improves recovery after hypoxia. Hypoxia interferes with protein expression at the translational level; however, its effect on mRNA levels for neuronal protein kinase and anti-apoptotic genes is unclear. To investigate the link between sevoflurane preconditioning and gene expression, hippocampal slices were treated with 4% sevoflurane for 15 min, a 5 min washout, 10 min of hypoxia, and 60 min of recovery. We used quantitative PCR to measure mRNA levels in the CA1 region of rat hippocampi. The mRNA levels for specific critical proteins were examined, as follows: Protein kinases, PKCγ (0.22), PKCε (0.38), and PKMζ (0.55) mRNAs, and anti-apoptotic, bcl-2 (0.44) and bcl-xl (0.41), were reduced 60 min after hypoxia relative to their expression in tissue not subjected to hypoxia (set to 1.0). Sevoflurane preconditioning prevented the reduction in PKMζ (0.88 vs. 1.0) mRNA levels after hypoxia. Pro-apoptotic BAD mRNA was not significantly changed after hypoxia, even with sevoflurane preconditioning (hypoxia 0.81, sevo hypoxia 0.84 vs. normoxia 1.0). However, BAD mRNA was increased by sevoflurane in non-hypoxic conditions (1.48 vs. 1.0), which may partially explain the deleterious effects of volatile anesthetics under certain conditions. The DNA repair enzyme poly ADP-ribose polymerase 1 (PARP-1) was increased by sevoflurane in tissue not subjected to hypoxia (1.23). PARP-1 mRNA was reduced in untreated tissue after hypoxia (0.21 vs. 1.0); sevoflurane did not improve PARP-1 after hypoxia (0.27). Interestingly, the mRNA level of the cognitive kinase PKMζ, a kinase essential for learning and memory, was the only one protected against hypoxic downregulation by sevoflurane preconditioning. These findings correlate with previous studies that found that sevoflurane-induced improvement of neuronal survival after hypoxia was dependent on PKMζ. Maintaining mRNA levels for critical proteins may provide an important mechanism for preserving neuronal function after stroke. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The effect of sevoflurane-induced preconditioning on mRNA levels of Protein Kinase C family genes after hypoxia. (<b>A</b>,<b>B</b>) There was a significant reduction of PKCγ and PKCε mRNA after hypoxia with and without sevoflurane when compared to normal oxygen. (<b>C</b>) There was a significant reduction of PKMζ mRNA after hypoxia without sevoflurane when compared to normal oxygen; sevoflurane significantly increased PKMζ mRNA after hypoxia when compared to hypoxia without sevoflurane. Hippocampal slices were perfused with normoxic aCSF (95% O<sub>2</sub>/5%CO<sub>2</sub>) and then subjected to either 0% sevoflurane (Nor) or 4% sevoflurane (Sev) for 15 min, washed for 5 min and then subjected to 10 min of hypoxia (95% N<sub>2</sub>/5% CO<sub>2</sub>) (Hyp or Hyp Sev) or normal oxygen (Nor or Nor Sev) followed by 60 of recovery in normoxic aCSF. The CA1 region of the hippocampus was dissected and analyzed using quantitative RT PCR. Data were analyzed with an ANOVA followed by the Student–Neuman–Keuls multiple comparison test; * <span class="html-italic">p</span> &lt; 0.05 vs. normoxia group; # <span class="html-italic">p</span> &lt; 0.01 vs. hypoxia group. Values are the mean ± the standard error of the mean (normoxia <span class="html-italic">n</span> = 4; sevo normoxia <span class="html-italic">n</span> = 5; hypoxia <span class="html-italic">n</span> = 7; sevo hypoxia <span class="html-italic">n</span> = 7).</p>
Full article ">Figure 2
<p>The effect of sevoflurane-induced preconditioning on mRNA levels of bcl-2 family pro- and anti-apoptotic genes. (<b>A</b>,<b>B</b>) There was a significant reduction in anti-apoptotic bcl-2 and bcl-xl mRNA after hypoxia with and without sevoflurane when compared to normal oxygen. (<b>C</b>) There was a significant increase in pro-apoptotic BAD mRNA in normal oxygen with sevoflurane when compared to normal oxygen without sevoflurane. Hippocampal slices were perfused with normoxic aCSF (95% O<sub>2</sub>/5%CO<sub>2</sub>) and then subjected to either 0% sevoflurane (Nor) or 4% sevoflurane (Sev) for 15 min, washed for 5 min and then subjected to 10 min of hypoxia (95% N<sub>2</sub>/5% CO<sub>2</sub>) (Hyp or Hyp Sev) or normal oxygen (Nor or Nor Sev), followed by 60 of recovery in normoxic aCSF. The CA1 region of the hippocampus was dissected and analyzed using quantitative RT PCR. Data were analyzed with an ANOVA followed by the Student–Neuman–Keuls multiple comparison test; * <span class="html-italic">p</span> &lt; 0.01 vs. normoxia group. Values are the mean ± the standard error of the mean (normoxia <span class="html-italic">n</span> = 4; sevo normoxia <span class="html-italic">n</span> = 5; hypoxia <span class="html-italic">n</span> = 7; sevo hypoxia <span class="html-italic">n</span> = 7).</p>
Full article ">Figure 3
<p>The effect of sevoflurane-induced preconditioning on mRNA levels of poly ADP-ribose polymerase 1 (PARP1). There was a significant increase in PARP-1 mRNA with sevoflurane in normal oxygen compared to normal oxygen without sevoflurane; there was a significant decrease in PARP-1 mRNA after hypoxia with or without sevoflurane. Hippocampal slices were perfused with normoxic aCSF (95% O<sub>2</sub>/5%CO<sub>2</sub>) and then subjected to either 0% sevoflurane (Nor) or 4% sevoflurane (Sev) for 15 min, washed for 5 min and then subjected to 10 min of hypoxia (95% N<sub>2</sub>/5% CO<sub>2</sub>) (Hyp or Hyp Sev) or normal oxygen (Nor or Nor Sev) followed by 60 of recovery in normoxic aCSF. The CA1 region of the hippocampus was dissected and analyzed using quantitative RT PCR. Data were analyzed with an ANOVA followed by the Student–Neuman–Keuls multiple comparison test; * <span class="html-italic">p</span> &lt; 0.01 vs. normoxia group. Values are the mean ± the standard error of the mean (normoxia <span class="html-italic">n</span> = 4; sevo normoxia <span class="html-italic">n</span> = 5; hypoxia <span class="html-italic">n</span> = 7; sevo hypoxia <span class="html-italic">n</span> = 7).</p>
Full article ">
Back to TopTop