[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

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

Search Results (853)

Search Parameters:
Keywords = Runx2

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
19 pages, 4508 KiB  
Article
Limited Adipogenic Differentiation Potential of Human Dental Pulp Stem Cells Compared to Human Bone Marrow Stem Cells
by Isaac Maximiliano Bugueno, Giuseppe Alastra, Anamaria Balic, Bernd Stadlinger and Thimios A. Mitsiadis
Int. J. Mol. Sci. 2024, 25(20), 11105; https://doi.org/10.3390/ijms252011105 - 16 Oct 2024
Viewed by 203
Abstract
Bone marrow and teeth contain mesenchymal stem cells (MSCs) that could be used for cell-based regenerative therapies. MSCs from these two tissues represent heterogeneous cell populations with varying degrees of lineage commitment. Although human bone marrow stem cells (hBMSCs) and human dental pulp [...] Read more.
Bone marrow and teeth contain mesenchymal stem cells (MSCs) that could be used for cell-based regenerative therapies. MSCs from these two tissues represent heterogeneous cell populations with varying degrees of lineage commitment. Although human bone marrow stem cells (hBMSCs) and human dental pulp stem cells (hDPSCs) have been extensively studied, it is not yet fully defined if their adipogenic potential differs. Therefore, in this study, we compared the in vitro adipogenic differentiation potential of hDPSCs and hBMSCs. Both cell populations were cultured in adipogenic differentiation media, followed by specific lipid droplet staining to visualise cytodifferentiation. The in vitro differentiation assays were complemented with the expression of specific genes for adipogenesis and osteogenesis–dentinogenesis, as well as for genes involved in the Wnt and Notch signalling pathways. Our findings showed that hBMSCs formed adipocytes containing numerous and large lipid vesicles. In contrast to hBMSCs, hDPSCs did not acquire the typical adipocyte morphology and formed fewer lipid droplets of small size. Regarding the gene expression, cultured hBMSCs upregulated the expression of adipogenic-specific genes (e.g., PPARγ2, LPL, ADIPONECTIN). Furthermore, in these cells most Wnt pathway genes were downregulated, while the expression of NOTCH pathway genes (e.g., NOTCH1, NOTCH3, JAGGED1, HES5, HEY2) was upregulated. hDPSCs retained their osteogenic/dentinogenic molecular profile (e.g., RUNX2, ALP, COLIA1) and upregulated the WNT-specific genes but not the NOTCH pathway genes. Taken together, our in vitro findings demonstrate that hDPSCs are not entirely committed to the adipogenic fate, in contrast to the hBMSCs, which are more effective to fully differentiate into adipocytes. Full article
Show Figures

Figure 1

Figure 1
<p><b>Experimental set-up of adipogenic induction in hBMSCs and hDPSCs.</b> hDPSCs and hBMSCs were cultured in 2D (6-well plates, 24-well plates, and μ-slide 4-well uncoated plates), and three time points were chosen for gene expression by RT qPCR (0, 7, and 21 days of induction) and six time points for lipid vesicle analysis (0, 7, 10, 14, and 21 days of induction). Bright-field and confocal scanning microscopy analyses before and after specific staining of lipid droplets (Nile red and LipidSpot<sup>TM</sup>) were performed up to 21 days of adipogenic induction.</p>
Full article ">Figure 2
<p><b>Adipogenic differentiation of hBMSCs and hDPSCs.</b> (<b>A</b>) Inverted bright-field microscopy on cells at 0 (a), 7 (b), 10 (c), 14 (d), and 21 (e) days of adipogenic induction in hBMSCs. (<b>B</b>) Inverted bright-field microscopy on cells at 0 (a), 7 (b), 10 (c), 14 (d), and 21 (e) days of adipogenic induction in hDPSCs. (<b>C</b>) Wide-field inverted fluorescence microscopy of hBMSCs at 0 (a,f), 7 (b,g), 10 (c,h), 14 (d,i), and 21 (e,j) days of differentiation (in yellow and red: Nile red staining). All scale bars represent 10 μm; n = 6. Below each photo, the isolated yellow channel allows better visualisation of the lipid droplets, indicated by the white arrows. (<b>D</b>) Wide-field inverted fluorescence microscopy of hDPSCs at 0 (a,f), 7 (b,g), 10 (c,h), 14 (d,i), and 21 (e,j) days of differentiation (in yellow and red: Nile red staining). All scale bars represent 10 μm; n = 6. Below each photo, the yellow channel allows better visualisation of the lipid droplets, indicated by the white arrows. (<b>E</b>,<b>F</b>) Confocal high-speed multispectral spinning-disk microscopy of lipid droplet staining of hBMSCs and hDPSCs at 0 (a), 7 (b), 10 (c), 14 (d), and 21 (e) days of differentiation (green colour: LipidSpot<sup>TM</sup> 488 staining; blue colour: DAPI staining). Scale bars: 10 μm.</p>
Full article ">Figure 3
<p><b>Comparison of the lipid vesicle formation between the hBMSC- and hDPSC-originated adipocytes</b>. (<b>A</b>) Quantification of fluorescence emitted by lipid droplet staining in fixed cells normalised with the cell number in three separate cultures of hBMSCs at 0, 7, 10, 14, and 21 days of differentiation; the curve shows the ratio of the green staining quantification versus the total cell number. (<b>B</b>) Histogram of fluorescence emitted by lipid droplet staining in live cells at 0, 7, 10, 14, and 21 days of adipogenic differentiation of hBMSCs and graphical representation of these values. (<b>C</b>) Quantification of fluorescence emitted by lipid droplet staining in fixed cells normalised with the cell number in three separate cultures of hDPSCs at 0, 7, 10, 14, and 21 days of differentiation; the curve shows the ratio of the green staining quantification versus the total cell number. (<b>D</b>) Histogram of fluorescence emitted by lipid droplet staining in live cells at 0, 7, 10, 14, and 21 days of adipogenic differentiation of hDPSCs and graphical representation of these values. (<b>E</b>) Lipid vesicles’ number quantification normalised by the cell number from live cell lipid droplets’ staining. (<b>F</b>) Lipid vesicles’ size quantification from live cell lipid droplets’ staining. One-way ANOVA followed by Dunnett’s post hoc test was used to compare time points for each cell type. Asterisks represent statistically significant differences between different time points and T0 control (* <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01, n = 6).</p>
Full article ">Figure 4
<p><b>Expression and comparison of adipogenic genes in cultured hBMSCs and hDPSCs under adipogenic conditions.</b> (<b>A</b>) Relative mRNA expression of the early adipogenic genes <span class="html-italic">PPARγ2</span> and <span class="html-italic">CEBPα</span> in cultured hBMSCs and hDPSCs for 0, 7, and 21 days under adipogenic conditions. (<b>B</b>) Expression of the characteristic late adipogenic genes <span class="html-italic">ADIPOQ</span>, <span class="html-italic">FABP4</span>, and <span class="html-italic">LPL</span> in cultured hBMSCs and hDPSCs at 0, 7, and 21 days of adipogenic induction was analysed by RT qPCR. The value of relative expression on comparison graphs is normalised to hBMSCs at T0 for each gene in statistical analysis. Data are presented as average values ± SD. One-way ANOVA followed by Dunnett’s post hoc test was used to compare time points for each cell type. Two-way ANOVA followed by Šídák post hoc test was used to compare hBMSCs and hDPSCs. Asterisks represent statistically significant differences between different time points and T0 control (* <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001; **** <span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 5
<p><b>Expression and comparison of stem cell and osteogenic genes in cultured hBMSCs and hDPSCs under adipogenic conditions</b>. (<b>A</b>) Relative mRNA expression of the stem cell gene <span class="html-italic">CD90</span> in cultured hBMSCs and hDPSCs at 0, 7, and 21 days under adipogenic conditions was analysed by RT qPCR. (<b>B</b>) Relative mRNA expression of the osteogenic genes <span class="html-italic">CD90</span>, <span class="html-italic">COLIII</span>, <span class="html-italic">ALPL</span>, <span class="html-italic">COLIA1</span>, and <span class="html-italic">RUNX2</span> in cultured hBMSCs and hDPSCs at 0, 7, and 21 days upon adipogenic induction was analysed by RT qPCR. The value of relative expression on comparison graphs is normalised to hBMSCs at T0 for the statistical analysis of each gene. Data are presented as average values ± SD. One-way ANOVA followed by Dunnett’s post hoc test was used to compare time points for each cell type. Two-way ANOVA followed by Šídák post hoc test was used to compare hBMSCs and hDPSCs. Asterisks represent statistically significant differences between different time points and T0 control (* <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001; **** <span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 6
<p><b>Expression and comparison of WNT and SWAT cell genes in hBMSCs and hDPSCs cultured under adipogenic conditions.</b> (<b>A</b>) Relative mRNA expression of the <span class="html-italic">WNT2</span> and <span class="html-italic">WNT10A</span> genes in cultured hBMSCs and hDPSCs at 0, 7, and 21 days under adipogenic conditions was analysed by RT qPCR. (<b>B</b>) Relative mRNA expression of the <span class="html-italic">PLIN1</span>, <span class="html-italic">DCN</span>, and <span class="html-italic">MFAP4</span> genes in cultured hBMSCs and hDPSCs at 0, 7, and 21 days of adipogenic induction was analysed by RT qPCR. The value of relative expression on comparison graphs is normalised to hBMSCs at T0 for each gene. Statistical analysis data are presented as average values ± SD. One-way ANOVA followed by Dunnett’s post hoc test was used to compare time points for each cell type. Two-way ANOVA followed by Šídák post hoc test was used for comparison between hBMSCs and hDPSCs. Asterisks represent statistically significant differences between different time points and T0 control (<b>*</b> <span class="html-italic">p</span> &lt; 0.05; <b>**</b> <span class="html-italic">p</span> &lt; 0.01; <b>***</b> <span class="html-italic">p</span> &lt; 0.001; <b>****</b> <span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 7
<p><b>Expression and comparison of NOTCH pathway genes in hBMSCs and hDPSCs cultured in adipogenic conditions</b>. (<b>A</b>) Relative mRNA expression of <span class="html-italic">NOTCH1</span> and <span class="html-italic">NOTCH3</span> in cultured hBMSCs and hDPSCs at 0, 7, and 21 days of adipogenic induction. (<b>B</b>) Relative mRNA expression of <span class="html-italic">JAGGED1</span> and <span class="html-italic">DELTA-LIKE4</span> (<span class="html-italic">DLL4</span>) in cultured hBMSCs and hDPSCs at 0, 7, and 21 days of adipogenic induction. (<b>C</b>) Relative mRNA expression of the Notch pathway transcription factors <span class="html-italic">HES5</span> and <span class="html-italic">HEY2</span> in cultured hBMSCs and hDPSCs at 0, 7, and 21 days of adipogenic induction. The value of relative expression on comparison graphs is normalised to hBMSCs at T0 for each gene. Statistical analysis data are presented as average values ± SD. One-way ANOVA followed by Dunnett’s post hoc test was used to compare time points for each cell type. Two-way ANOVA followed by Šídák post hoc test was used to compare hBMSCs and hDPSCs. Asterisks represent statistically significant differences between different time points and T0 control (* <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001; **** <span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 8
<p><b>Schematic representation of the differentiation of cultured hBMSCs and hDPSCs under adipogenic conditions and the expression of various genes during the differentiation process.</b> Cultured hBMSCs form numerous large lipid droplets (yellow colour). Upregulation (green arrows) of adipogenic genes and concomitant downregulation (red arrows) of WNT, stem cell, and osteogenic genes indicate their unreserved commitment towards the adipogenic fate. By contrast, hDPSCs form tiny and less numerous lipid vesicles when compared to hBMSCs and demonstrate a less evident adipocyte morphology. Furthermore, hDPSCs keep high expression of most WNT, stem cell, and osteogenic genes, which indicates their partial commitment towards the adipogenic fate.</p>
Full article ">
11 pages, 1307 KiB  
Case Report
Blast Transformation of Chronic Myeloid Leukemia Driven by Acquisition of t(8;21)(q22;q22)/RUNX1::RUNX1T1: Selecting Optimal Treatment Based on Clinical and Molecular Findings
by Adolfo Fernández-Sánchez, Alberto Hernández-Sánchez, Cristina De Ramón, María-Carmen Chillón, María Belén Vidriales, Mónica Baile-González, Cristina-Teresa Fuentes-Morales, Magdalena Sierra-Pacho, Lucía López-Corral and Fermín Sánchez-Guijo
Biomedicines 2024, 12(10), 2339; https://doi.org/10.3390/biomedicines12102339 (registering DOI) - 15 Oct 2024
Viewed by 371
Abstract
The advent of tyrosine kinase inhibitors (TKIs) has changed the natural history of chronic myeloid leukemia (CML), and the transformation from the chronic phase to the blast phase (BP) is currently an uncommon situation. However, it is one of the major remaining challenges [...] Read more.
The advent of tyrosine kinase inhibitors (TKIs) has changed the natural history of chronic myeloid leukemia (CML), and the transformation from the chronic phase to the blast phase (BP) is currently an uncommon situation. However, it is one of the major remaining challenges in the management of this disease, as it is associated with dismal outcomes. We report the case of a 63-year-old woman with a history of CML with poor response to imatinib who progressed to myeloid BP-CML, driven by the acquisition of t(8;21)(q22;q22)/RUNX1::RUNX1T1. The patient received intensive chemotherapy and dasatinib, followed by allogeneic hematopoietic stem cell transplantation (allo-HSCT). However, she suffered an early relapse after allo-HSCT with the acquisition of the T315I mutation in ABL1. Ponatinib and azacitidine were started as salvage treatment, allowing for the achievement of complete remission with deep molecular response after five cycles. Advances in the knowledge of disease biology and clonal evolution are crucial for optimal treatment selection, which ultimately translates into better patient outcomes. Full article
(This article belongs to the Special Issue Advances in the Pathogenesis and Treatment of Acute Myeloid Leukemia)
Show Figures

Figure 1

Figure 1
<p>Cytogenetic studies of the patient at myeloid BP-CML, showing the acquisition of t(8;21)(q22;q22)/<span class="html-italic">RUNX1::RUNX1T1</span> as the driver event of BP. (<b>a</b>): <span class="html-italic">BCR::ABL1</span> fusion probe (Dual Color, Dual Fusion, Vysis LSI) in the FISH study, showing two fusions in most of the cells, representative of the presence of t(9;22)(q34.1;q11.2)/<span class="html-italic">BCR::ABL1</span> (<b>b</b>): <span class="html-italic">RUNX1::RUNX1T1</span> fusion probe (Dual Color, Dual Fusion, Vysis LSI) in the FISH study, showing two fusions in approximately half of the cells, representative of the presence of t(8;21)(q22;q22)/<span class="html-italic">RUNX1::RUNX1T1</span> (<b>c</b>): A Circos plot of optical genome mapping (Bionano) showing concomitant t(8;21)(q22;q22)/<span class="html-italic">RUNX1::RUNX1T1</span> and t(9;22)(q34.1;q11.2)/<span class="html-italic">BCR::ABL1</span> in the patient sample.</p>
Full article ">Figure 2
<p>Polymerase chain reaction and Sanger sequencing analysis of T315I mutation in <span class="html-italic">BCR::ABL1</span> fusion gene (c.944C &gt; T, p.Thr315Ile).</p>
Full article ">Figure 3
<p>A summary of the clinical case report. The temporal evolution of the <span class="html-italic">BCR::ABL1/ABL1</span> ratio according to the International Scale (represented in blue) and the number of <span class="html-italic">RUNX1::RUNX1T1</span> transcripts (represented in red) in peripheral blood samples are shown, together with the main events that the patient presented during the disease evolution. CML: chronic myeloid leukemia; BP: blast phase; FLAG-IDA: fludarabine, cytarabine, idarubicin and granulocyte colony-stimulating factor; allo-HCST: allogenic hematopoietic stem cell transplantation; MR: molecular response.</p>
Full article ">
20 pages, 2775 KiB  
Systematic Review
Genetics of Calcific Aortic Stenosis: A Systematic Review
by Vassilios S. Vassiliou, Nicholas Johnson, Kenneth Langlands and Vasiliki Tsampasian
Genes 2024, 15(10), 1309; https://doi.org/10.3390/genes15101309 - 10 Oct 2024
Viewed by 523
Abstract
Background: Calcific aortic stenosis is the most prevalent valvular abnormality in the Western world. Factors commonly associated with calcific aortic stenosis include advanced age, male sex, hypertension, diabetes and impaired renal function. This review synthesises the existing literature on genetic associations with calcific [...] Read more.
Background: Calcific aortic stenosis is the most prevalent valvular abnormality in the Western world. Factors commonly associated with calcific aortic stenosis include advanced age, male sex, hypertension, diabetes and impaired renal function. This review synthesises the existing literature on genetic associations with calcific aortic stenosis. Methods: A systematic search was conducted in the PubMed, Ovid and Cochrane libraries from inception to 21 July 2024 to identify human studies investigating the genetic factors involved in calcific aortic stenosis. From an initial pool of 1392 articles, 78 were selected for full-text review and 31 were included in the final qualitative synthesis. The risk of bias in these studies was assessed using the Newcastle Ottawa Scale. Results: Multiple genes have been associated with calcific aortic stenosis. These genes are involved in different biological pathways, including the lipid metabolism pathway (PLA, LDL, APO, PCSK9, Lp-PLA2, PONS1), the inflammatory pathway (IL-6, IL-10), the calcification pathway (PALMD, TEX41) and the endocrine pathway (PTH, VIT D, RUNX2, CACNA1C, ALPL). Additional genes such as NOTCH1, NAV1 and FADS1/2 influence different pathways. Mechanistically, these genes may promote a pro-inflammatory and pro-calcific environment in the aortic valve itself, leading to increased osteoblastic activity and subsequent calcific degeneration of the valve. Conclusions: Numerous genetic associations contribute to calcific aortic stenosis. Recognition of these associations can enhance risk stratification for individuals and their first-degree relatives, facilitate family screening, and importantly, pave the way for targeted therapeutic interventions focusing on the identified genetic factors. Understanding these genetic factors can also lead to gene therapy to prevent calcific aortic stenosis in the future. Full article
(This article belongs to the Special Issue Genomics and Genetics of Cardiovascular Diseases)
Show Figures

Figure 1

Figure 1
<p>A simplified illustration of the human aortic valve is shown. Panel (<b>A</b>) depicts the anatomical structure of the aortic valve in relation to aorta and ventricle. Reproduced with permission from Henderson et al. [<a href="#B2-genes-15-01309" class="html-bibr">2</a>] under a creative commons attribution 4.0 international license. Panel (<b>B</b>) focuses on the aortic valve leaflet. The left side presents a schematic cross-section of the non-coronary leaflet of the aortic valve. The enlarged section on the right highlights the three-layered structure of the extracellular matrix, indicating the locations of the aortic valve endothelial cells (VECs) and valvular interstitial cells (VICs). Reproduced with permission from Rutkovskiy et al. [<a href="#B1-genes-15-01309" class="html-bibr">1</a>] under a creative commons attribution 4.0 international license.</p>
Full article ">Figure 2
<p>PRISMA flow diagram summarising study selection.</p>
Full article ">Figure 3
<p>Schematic representation of Lp(a). Lp(a) consists of apolipoprotein B100 covalently linked to apo(a), and high levels enable a pro-inflammatory and pro-calcific environment. Variants with smaller kringle IV type 2 repeats are associated with higher blood Lp(a) levels; hence, the plasma concentration of Lp(a) is genetically determined. Reproduced with permission from Telyuk et al. [<a href="#B46-genes-15-01309" class="html-bibr">46</a>] under a creative commons attribution 4.0 international license.</p>
Full article ">Figure 4
<p>Mechanism of action of PCSK9 inhibitors. PCSK9 identifies and marks the LDL receptor for phagocytosis (number 1), thereby allowing more LDL to circulate in the bloodstream. PCSK9 inhibitors bind to LDL receptors (2), and thus allow the LDL receptors to clear more LDL particles (shown in 3). Reproduced from Beltran et al. [<a href="#B68-genes-15-01309" class="html-bibr">68</a>] under a creative commons attribution 4.0 international license.</p>
Full article ">
17 pages, 3752 KiB  
Article
Extracorporeal Magnetotransduction Therapy as a New Form of Electromagnetic Wave Therapy: From Gene Upregulation to Accelerated Matrix Mineralization in Bone Healing
by Lennart Gerdesmeyer, Jutta Tübel, Andreas Obermeier, Norbert Harrasser, Claudio Glowalla, Rüdiger von Eisenhart-Rothe and Rainer Burgkart
Biomedicines 2024, 12(10), 2269; https://doi.org/10.3390/biomedicines12102269 - 7 Oct 2024
Viewed by 779
Abstract
Background: Electromagnetic field therapy is gaining attention for its potential in treating bone disorders, with Extracorporeal Magnetotransduction Therapy (EMTT) emerging as an innovative approach. EMTT offers a higher oscillation frequency and magnetic field strength compared to traditional Pulsed Electromagnetic Field (PEMF) therapy, showing [...] Read more.
Background: Electromagnetic field therapy is gaining attention for its potential in treating bone disorders, with Extracorporeal Magnetotransduction Therapy (EMTT) emerging as an innovative approach. EMTT offers a higher oscillation frequency and magnetic field strength compared to traditional Pulsed Electromagnetic Field (PEMF) therapy, showing promise in enhancing fracture healing and non-union recovery. However, the mechanisms underlying these effects remain unclear. Results: This study demonstrates that EMTT significantly enhances osteoblast bone formation at multiple levels, from gene expression to extracellular matrix mineralization. Key osteoblastogenesis regulators, including SP7 and RUNX2, and bone-related genes such as COL1A1, ALPL, and BGLAP, were upregulated, with expression levels surpassing those of the control group by over sevenfold (p < 0.001). Enhanced collagen synthesis and mineralization were confirmed by von Kossa and Alizarin Red staining, indicating increased calcium and phosphate deposition. Additionally, calcium imaging revealed heightened calcium influx, suggesting a cellular mechanism for EMTT’s osteogenic effects. Importantly, EMTT did not compromise cell viability, as confirmed by live/dead staining and WST-1 assays. Conclusion: This study is the first to show that EMTT can enhance all phases of osteoblastogenesis and improve the production of critical mineralization components, offering potential clinical applications in accelerating fracture healing, treating osteonecrosis, and enhancing implant osseointegration. Full article
(This article belongs to the Section Biomedical Engineering and Materials)
Show Figures

Figure 1

Figure 1
<p>Experimental setup and protocol. (<b>a</b>) Experimental setup used in our study. The control and stimulated groups were placed identically in the incubator with the EMTT device turned on for the stimulated group and kept off for the control group. (<b>b</b>) Stimulation protocol used in our study. The day of stimulation and medium change are illustrated.</p>
Full article ">Figure 2
<p>Illustration of the physical parameters of EMTT. (<b>a</b>) Variation in the measured magnetic field strength relative to the lateral and axial axis. (<b>b</b>) Oscillation frequency reaching up to 300 kHz.</p>
Full article ">Figure 3
<p>EMTT stimulation has no impact on the cell viability. (<b>a</b>–<b>d</b>) The hOBs were stained using the live/dead viability/cytotoxicity kit on days 7 (<b>a</b>,<b>b</b>) and 14 (<b>c</b>,<b>d</b>). Four representative images are illustrated. Stained with calcein AM (viable cells stain green) and ethidium homodimer 1 (non-viable cells stain red; see red arrow). (<b>e</b>) Quantification of living and dead cells was conducted based on counting cells of three wells. Each well is represented by five image excerpts, totaling 15 representative sections (5 per well). (<b>f</b>) The measured absorbance at 450 nm (corrected with 620 nm) represents the proliferation rate of hOBs using the WST-1 reagent. Cells were stimulated with EMTT according to our stimulation protocol (<a href="#biomedicines-12-02269-f001" class="html-fig">Figure 1</a>b). Data are expressed as the average ± SD of three independent experiments (<span class="html-italic">n</span> = 3). ns: non-significant. two-tailed Student’s <span class="html-italic">t</span>-test.</p>
Full article ">Figure 4
<p>EMTT increases the intracellular calcium concentration in hOBs. The hOBs were incubated at 37 °C for 45 min with Calbryte 520 AM containing buffer in the presence of 2.5 mM probenecid. (<b>a</b>,<b>b</b>) Cells were stimulated with EMTT for 10 min, and the images were taken before (<b>a</b>) and after (<b>b</b>) stimulation with a fluorescence microscope using the FITC channel. Two representative images are illustrated. Red arrows show calcium increase in the cells. (<b>c</b>) The cellular fluorescence images were quantitated using ImageJ software. (<b>d</b>) Fluorescence was measured using a fluorescence plate reader (Ex: 485/Em: 538). Data are expressed as the average ± SD of five independent experiments (<span class="html-italic">n</span> = 5). ns: non-significant. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001, paired Student’s <span class="html-italic">t</span>-test and two-tailed Student’s <span class="html-italic">t</span>-test.</p>
Full article ">Figure 5
<p>EMTT stimulation enhanced OCN and ALP synthesis and induced an upregulation in osteogenic genes in hOBs. (<b>a</b>–<b>e</b>) Data analysis was performed by using the 2<sup>−ΔΔCT</sup> method. Gene expression was normalized to GAPDH and compared by setting control cultures to 1 as a reference value. Cells were stimulated with EMTT according to our stimulation protocol (<a href="#biomedicines-12-02269-f001" class="html-fig">Figure 1</a>b), followed by RNA extraction and PCR. (<b>f</b>) OCN Protein levels were determined using ELISA according to the manufacturer’s instructions. Supernatants were collected at different time points: day 0 (pre-first EMTT stimulation), 4, 7, 11, 14, and 16. (<b>g</b>) ALP staining. Two representative wells are illustrated. (<b>h</b>) The images were quantitated using ImageJ software. (<b>i</b>) ALP activity was measured through a colorimetric kinetic assay and compared by setting control cultures to 1 as a reference value. Cells were stimulated once until day 2, twice until day 4, and thrice until day 7. Data are expressed as the average ± SD of three to six independent experiments (<span class="html-italic">n</span> = 3–6). ns: non-significant. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001, two-tailed Student’s <span class="html-italic">t</span>-test.</p>
Full article ">Figure 6
<p>EMTT stimulation enhanced collagen synthesis in human osteoblasts (hOBs). (<b>a</b>) Sirius Red staining. Two representative wells are illustrated. (<b>b</b>) The images were quantitated using ImageJ software. (<b>c</b>) The Sirius Red dye was eluted using 0.1 N Sodium Hydroxide (NaOH), and the absorbance was measured at 570 nm. Results are presented relative to the control (normalized to 1). (<b>d</b>) Data analysis was performed by using the 2<sup>−ΔΔCT</sup> method. Gene expression was normalized to GAPDH and compared by setting control cultures to 1 as a reference value. Cells were stimulated with EMTT according to our stimulation protocol (<a href="#biomedicines-12-02269-f001" class="html-fig">Figure 1</a>b), followed by RNA extraction and PCR. (<b>e</b>) Pro-Collagen-1-α1 protein levels were determined using ELISA according to the manufacturer’s instructions. Supernatants were collected at different time points: day 0 (pre-first EMTT stimulation), 4, 7, 11, 14, and 16. Results are presented relative to the control (normalized to 1). Data are expressed as the average ± SD of three to six independent experiments (<span class="html-italic">n</span> = 3–6). ns: non-significant. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001, two-tailed Student’s <span class="html-italic">t</span>-test.</p>
Full article ">Figure 7
<p>EMTT stimulation enhanced mineralization and upregulated mineralization-related genes in human osteoblasts (hOBs). (<b>a</b>) Alizarin Red S staining. (<b>b</b>) Von Kossa staining. Two representative wells are illustrated. (<b>c</b>,<b>d</b>) The images were quantitated using ImageJ software. (<b>e</b>) The Alizarin Red dye was eluted using cetylpyridinium chloride, and the absorbance was measured at 405 nm. Results are presented relative to the control (normalized to 1). (<b>f</b>–<b>h</b>) Data analysis was performed by using the 2<sup>−ΔΔCT</sup> method. Gene expression was normalized to GAPDH and compared by setting control cultures to 1 as a reference value. (<b>f</b>–<b>h</b>) Cells were stimulated with EMTT according to our stimulation protocol (<a href="#biomedicines-12-02269-f001" class="html-fig">Figure 1</a>b), followed by RNA extraction and PCR. Data are expressed as the average ± SD of three to six independent experiments (<span class="html-italic">n</span> = 3–6). ns: non-significant. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001, two-tailed Student’s <span class="html-italic">t</span>-test.</p>
Full article ">
18 pages, 6667 KiB  
Article
Impact of Polydeoxyribonucleotides on the Morphology, Viability, and Osteogenic Differentiation of Gingiva-Derived Stem Cell Spheroids
by Heera Lee, Somyeong Hwa, Sunga Cho, Ju-Hwan Kim, Hye-Jung Song, Youngkyung Ko and Jun-Beom Park
Medicina 2024, 60(10), 1610; https://doi.org/10.3390/medicina60101610 - 1 Oct 2024
Viewed by 591
Abstract
Background and Objectives: Polydeoxyribonucleotides (PDRN), composed of DNA fragments derived from salmon DNA, is widely recognized for its regenerative properties. It has been extensively used in medical applications, such as dermatology and wound healing, due to its ability to enhance cellular metabolic [...] Read more.
Background and Objectives: Polydeoxyribonucleotides (PDRN), composed of DNA fragments derived from salmon DNA, is widely recognized for its regenerative properties. It has been extensively used in medical applications, such as dermatology and wound healing, due to its ability to enhance cellular metabolic activity, stimulate angiogenesis, and promote tissue regeneration. In the field of dentistry, PDRN has shown potential in promoting periodontal healing and bone regeneration. This study aims to investigate the effects of PDRN on the morphology, survival, and osteogenic differentiation of gingiva-derived stem cell spheroids, with a focus on its potential applications in tissue engineering and regenerative dentistry. Materials and Methods: Gingiva-derived mesenchymal stem cells were cultured and formed into spheroids using microwells. The cells were treated with varying concentrations of PDRN (0, 25, 50, 75, and 100 μg/mL) and cultivated in osteogenic media. Cell morphology was observed over seven days using an inverted microscope, and viability was assessed with Live/Dead Kit assays and Cell Counting Kit-8. Osteogenic differentiation was evaluated by measuring alkaline phosphatase activity and calcium deposition. The expression levels of osteogenic markers RUNX2 and COL1A1 were quantified using real-time polymerase chain reaction. RNA sequencing was performed to assess the gene expression profiles related to osteogenesis. Results: The results demonstrated that PDRN treatment had no significant effect on spheroid diameter or cellular viability during the observation period. However, a PDRN concentration of 75 μg/mL significantly enhanced calcium deposition by Day 14, suggesting increased mineralization. RUNX2 and COL1A1 mRNA expression levels varied with PDRN concentration, with the highest RUNX2 expression observed at 25 μg/mL and the highest COL1A1 expression at 75 μg/mL. RNA sequencing further confirmed the upregulation of genes involved in osteogenic differentiation, with enhanced expression of RUNX2 and COL1A1 in PDRN-treated gingiva-derived stem cell spheroids. Conclusions: In summary, PDRN did not significantly affect the viability or morphology of gingiva-derived stem cell spheroids but influenced their osteogenic differentiation and mineralization in a concentration-dependent manner. These findings suggest that PDRN may play a role in promoting osteogenic processes in tissue engineering and regenerative dentistry applications, with specific effects observed at different concentrations. Full article
(This article belongs to the Section Dentistry and Oral Health)
Show Figures

Figure 1

Figure 1
<p>Morphological evaluation of stem cell spheroids. (<b>A</b>) Representative images of spheroids treated with PDRN at concentrations of 0, 25, 50, 75, and 100 μg/mL on Days 0, 1, 3, 5, and 7. Scale bar = 200 μm (original magnification ×200). (<b>B</b>) Changes in spheroid diameter over time, measured on Days 0, 1, 3, 5, and 7.</p>
Full article ">Figure 1 Cont.
<p>Morphological evaluation of stem cell spheroids. (<b>A</b>) Representative images of spheroids treated with PDRN at concentrations of 0, 25, 50, 75, and 100 μg/mL on Days 0, 1, 3, 5, and 7. Scale bar = 200 μm (original magnification ×200). (<b>B</b>) Changes in spheroid diameter over time, measured on Days 0, 1, 3, 5, and 7.</p>
Full article ">Figure 2
<p>Cellular viability assessment. (<b>A</b>) Optical, live, dead, and merged images of stem cell spheroids on Day 1. Scale bar = 200 μm (original magnification ×200). (<b>B</b>) Optical, live, dead, and merged images of stem cell spheroids on Day 7. Scale bar = 200 μm (original magnification ×200). (<b>C</b>) Quantitative analysis of cell viability using the Cell Counting Kit-8 on Days 1, 3, 5, and 7. Varying concentrations of PDRN did not have a significant effect on cell viability over the seven-day period (<span class="html-italic">p</span> &gt; 0.05).</p>
Full article ">Figure 2 Cont.
<p>Cellular viability assessment. (<b>A</b>) Optical, live, dead, and merged images of stem cell spheroids on Day 1. Scale bar = 200 μm (original magnification ×200). (<b>B</b>) Optical, live, dead, and merged images of stem cell spheroids on Day 7. Scale bar = 200 μm (original magnification ×200). (<b>C</b>) Quantitative analysis of cell viability using the Cell Counting Kit-8 on Days 1, 3, 5, and 7. Varying concentrations of PDRN did not have a significant effect on cell viability over the seven-day period (<span class="html-italic">p</span> &gt; 0.05).</p>
Full article ">Figure 3
<p>Osteogenic differentiation in PDRN-treated spheroids. (<b>A</b>) Alkaline phosphatase activity in PDRN-treated spheroids measured on Days 7 and 14. (<b>B</b>) Evaluation of calcium deposition in PDRN-treated spheroids on Days 7 and 14. (<b>C</b>) Quantitative analysis of calcium deposition in spheroids treated with different PDRN concentrations. * <span class="html-italic">p</span> &lt; 0.05 on day 14 compared to the time-matched unloaded group.</p>
Full article ">Figure 3 Cont.
<p>Osteogenic differentiation in PDRN-treated spheroids. (<b>A</b>) Alkaline phosphatase activity in PDRN-treated spheroids measured on Days 7 and 14. (<b>B</b>) Evaluation of calcium deposition in PDRN-treated spheroids on Days 7 and 14. (<b>C</b>) Quantitative analysis of calcium deposition in spheroids treated with different PDRN concentrations. * <span class="html-italic">p</span> &lt; 0.05 on day 14 compared to the time-matched unloaded group.</p>
Full article ">Figure 4
<p>mRNA expression analysis. (<b>A</b>) Quantification of RUNX2 mRNA expression using qPCR on Day 7. * <span class="html-italic">p</span> &lt; 0.05 on day 7 compared to the unloaded group. ** <span class="html-italic">p</span> &lt; 0.05 compared to the unloaded group. (<b>B</b>) Quantification of COL1A1 mRNA expression using qPCR on Day 7. * <span class="html-italic">p</span> &lt; 0.05 compared to the unloaded group.</p>
Full article ">Figure 5
<p>RNA sequencing analysis. (<b>A</b>) Cluster heatmap with hierarchical clustering showing osteogenic differentiation in PDRN-treated gingiva-derived stem cell spheroids. (<b>B</b>) Functional enrichment analysis. (<b>C</b>) Pathway relationship network colored by cluster identity. (<b>D</b>) Pathway relationship network colored by p-value significance. (<b>E</b>) Gene expression profiling for regulation of osteogenic differentiation in PDRN-treated stem cell spheroids.</p>
Full article ">Figure 5 Cont.
<p>RNA sequencing analysis. (<b>A</b>) Cluster heatmap with hierarchical clustering showing osteogenic differentiation in PDRN-treated gingiva-derived stem cell spheroids. (<b>B</b>) Functional enrichment analysis. (<b>C</b>) Pathway relationship network colored by cluster identity. (<b>D</b>) Pathway relationship network colored by p-value significance. (<b>E</b>) Gene expression profiling for regulation of osteogenic differentiation in PDRN-treated stem cell spheroids.</p>
Full article ">Figure 5 Cont.
<p>RNA sequencing analysis. (<b>A</b>) Cluster heatmap with hierarchical clustering showing osteogenic differentiation in PDRN-treated gingiva-derived stem cell spheroids. (<b>B</b>) Functional enrichment analysis. (<b>C</b>) Pathway relationship network colored by cluster identity. (<b>D</b>) Pathway relationship network colored by p-value significance. (<b>E</b>) Gene expression profiling for regulation of osteogenic differentiation in PDRN-treated stem cell spheroids.</p>
Full article ">
24 pages, 13262 KiB  
Article
Placental Tissue Calcification and Its Molecular Pathways in Female Patients with Late-Onset Preeclampsia
by Miguel A. Ortega, Tatiana Pekarek, Diego De Leon-Oliva, Diego Liviu Boaru, Oscar Fraile-Martinez, Cielo García-Montero, Julia Bujan, Leonel Pekarek, Silvestra Barrena-Blázquez, Raquel Gragera, Patrocinio Rodríguez-Benitez, Mauricio Hernández-Fernández, Laura López-González, Raul Díaz-Pedrero, Ángel Asúnsolo, Melchor Álvarez-Mon, Natalio García-Honduvilla, Miguel A. Saez, Juan A. De León-Luis and Coral Bravo
Biomolecules 2024, 14(10), 1237; https://doi.org/10.3390/biom14101237 - 30 Sep 2024
Viewed by 377
Abstract
Preeclampsia (PE) is a complex multisystem disease characterized by hypertension of sudden onset (>20 weeks’ gestation) coupled with the presence of at least one additional complication, such as proteinuria, maternal organ dysfunction, or uteroplacental dysfunction. Hypertensive states during pregnancy carry life-threatening risks for [...] Read more.
Preeclampsia (PE) is a complex multisystem disease characterized by hypertension of sudden onset (>20 weeks’ gestation) coupled with the presence of at least one additional complication, such as proteinuria, maternal organ dysfunction, or uteroplacental dysfunction. Hypertensive states during pregnancy carry life-threatening risks for both mother and baby. The pathogenesis of PE develops due to a dysfunctional placenta with aberrant architecture that releases factors contributing to endothelial dysfunction, an antiangiogenic state, increased oxidative stress, and maternal inflammatory responses. Previous studies have shown a correlation between grade 3 placental calcifications and an elevated risk of developing PE at term. However, little is known about the molecular pathways leading to placental calcification. In this work, we studied the gene and protein expression of c-Jun N-terminal kinase (JNK), Runt-related transcription factor 2 (RUNX2), osteocalcin (OSC), osteopontin (OSP), pigment epithelium-derived factor (PEDF), MSX-2/HOX8, SOX-9, WNT-1, and β-catenin in placental tissue from women with late-onset PE (LO-PE). In addition, we employed von Kossa staining to detect mineral deposits in placental tissues. Our results show a significant increase of all these components in placentas from women with LO-PE. Therefore, our study suggests that LO-PE may be associated with the activation of molecular pathways of placental calcification. These results could be the starting point for future research to describe the molecular mechanisms that promote placental calcification in PE and the development of therapeutic strategies directed against it. Full article
(This article belongs to the Special Issue Tissue Calcification in Normal and Pathological Environments)
Show Figures

Figure 1

Figure 1
<p>Increased gene and protein expression of JNK in placental villi of patients with late-onset preeclampsia (LO-PE). (<b>A</b>) Relative amount of mRNA coding for JNK in LO-PE patients and healthy controls (HC). (<b>B</b>) Percentage of immunoreactivity for JNK expression in placental villi of LO-PE and HC group. (<b>C</b>,<b>D</b>) Photomicrographs showing JNK immunostaining in LO-PE and HC placental villi. <span class="html-italic">n</span> (HC) = 43; <span class="html-italic">n</span> (LO-PE) = 68. <span class="html-italic">p</span> &lt; 0.001 (***).</p>
Full article ">Figure 2
<p>Increased gene and protein expression of RUNX2 in placental villi of patients with late-onset preeclampsia (LO-PE). (<b>A</b>) Relative amount of mRNA coding for RUNX2 in LO-PE patients and healthy controls (HC). (<b>B</b>) Percentage of immunoreactivity for RUNX2 expression in placental villi of LO-PE and HC group. (<b>C</b>,<b>D</b>) Photomicrographs showing RUNX2 immunostaining in the placental villi of LO-PE and HC. <span class="html-italic">n</span> (HC) = 43; <span class="html-italic">n</span> (LO-PE) = 68. <span class="html-italic">p</span> &lt; 0.001 (***).</p>
Full article ">Figure 3
<p>Increased gene and protein expression of OSC in placental villi of patients with late-onset preeclampsia (LO-PE). (<b>A</b>) Relative amount of mRNA coding for OSC in patients with LO-PE and healthy controls (HC). (<b>B</b>) Percentage of immunoreactivity for OSC expression in placental villi of LO-PE and HC group. (<b>C</b>,<b>D</b>) Photomicrographs showing OSC immunostaining in LO-PE and HC placental villi. <span class="html-italic">n</span> (HC) = 43; <span class="html-italic">n</span> (LO-PE) = 68. <span class="html-italic">p</span> &lt; 0.01 (**); <span class="html-italic">p</span> &lt; 0.001 (***).</p>
Full article ">Figure 4
<p>Increased gene and protein expression of OSP in placental villi of patients with late-onset preeclampsia (LO-PE). (<b>A</b>) Relative amount of mRNA coding for OSP in patients with LO-PE and healthy controls (HC). (<b>B</b>) Percentage of immunoreactivity for OSP expression in placental villi of LO-PE and HC group. (<b>C</b>,<b>D</b>) Photomicrographs showing OSP immunostaining in LO-PE and HC placental villi. <span class="html-italic">n</span> (HC) = 43; <span class="html-italic">n</span> (LO-PE) = 68. <span class="html-italic">p</span> &lt; 0.05 (*); <span class="html-italic">p</span> &lt; 0.001 (***).</p>
Full article ">Figure 5
<p>Increased gene and protein expression of PEDF in placental villi of patients with late-onset preeclampsia (LO-PE). (<b>A</b>) Relative amount of mRNA coding for PEDF in patients with LO-PE and healthy controls (HC). (<b>B</b>) Percentage of immunoreactivity for PEDF expression in placental villi of LO-PE and HC group. (<b>C</b>,<b>D</b>) Photomicrographs showing PEDF immunostaining in LO-PE and HC placental villi. <span class="html-italic">n</span> (HC) = 43; <span class="html-italic">n</span> (LO-PE) = 68. <span class="html-italic">p</span> &lt; 0.001 (***).</p>
Full article ">Figure 6
<p>Increased gene and protein expression of MSX2/HOX8 in placental villi of patients with late-onset preeclampsia (LO-PE). (<b>A</b>) Relative amount of mRNA coding for MSX2/HOX8 in patients with LO-PE and healthy controls (HC). (<b>B</b>) Percentage of immunoreactivity for MSX2/HOX8 expression in placental villi of LO-PE and HC group. (<b>C</b>,<b>D</b>) Photomicrographs showing MSX2/HOX8 immunostaining in LO-PE and HC placental villi. <span class="html-italic">n</span> (HC) = 43; <span class="html-italic">n</span> (LO-PE) = 68. <span class="html-italic">p</span> &lt; 0.001 (***).</p>
Full article ">Figure 7
<p>Increased gene and protein expression of SOX9 in placental villi of patients with late-onset preeclampsia (LO-PE). (<b>A</b>) Relative amount of mRNA coding for SOX9 in LO-PE patients and healthy controls (HC). (<b>B</b>) Percentage of immunoreactivity for SOX) expression in placental villi of LO-PE and HC group. (<b>C</b>,<b>D</b>) Photomicrographs showing SOX9 immunostaining in LO-PE and HC placental villi. <span class="html-italic">n</span> (HC) = 43; <span class="html-italic">n</span> (LO-PE) = 68. <span class="html-italic">p</span> &lt; 0.001 (***).</p>
Full article ">Figure 8
<p>Increased gene and protein expression of WNT-1 in placental villi of patients with late-onset preeclampsia (LO-PE). (<b>A</b>) Relative amount of mRNA coding for WNT-1 in LO-PE patients and healthy controls (HC). (<b>B</b>) Percentage of immunoreactivity for WNT-1 expression in placental villi of LO-PE and HC group. (<b>C</b>,<b>D</b>) Photomicrographs showing WNT-1 immunostaining in LO-PE and HC placental villi. <span class="html-italic">n</span> (HC) = 43; <span class="html-italic">n</span> (LO-PE) = 68. <span class="html-italic">p</span> &lt; 0.05 (*); <span class="html-italic">p</span> &lt; 0.01 (**).</p>
Full article ">Figure 9
<p>Increased gene and protein expression of β-catenin in placental villi of patients with late-onset preeclampsia (LO-PE). (<b>A</b>) Relative amount of mRNA coding for β-catenin in patients with LO-PE and healthy controls (HC). (<b>B</b>) Percentage of immunoreactivity for β-catenin expression in placental villi of LO-PE and HC group. (<b>C</b>,<b>D</b>) Photomicrographs showing β-catenin immunostaining in LO-PE and HC placental villi. <span class="html-italic">n</span> (HC) = 43; <span class="html-italic">n</span> (LO-PE) = 68. <span class="html-italic">p</span> &lt; 0.05 (*); <span class="html-italic">p</span> &lt; 0.01 (**).</p>
Full article ">Figure 10
<p>Von Kossa staining of placental villi from patients with late-onset preeclampsia (LO-PE) and healthy controls (HC). (<b>A</b>,<b>C</b>). Photomicrographs showing dystrophic calcifications in patients with LO-PE and HC. (<b>B</b>,<b>D</b>). Photomicrographs showing metastatic calcifications in patients with LO-PE and HC.</p>
Full article ">Figure 11
<p>(<b>A</b>) Descriptive figure of the localization and function of the different molecules analyzed in this study. (<b>B</b>) Molecular signaling model of placental calcification under LO-PE conditions based on our results and studies in other cell lines. IRI: ischemia/reperfusion injury; OS: oxidative stress.</p>
Full article ">
17 pages, 4336 KiB  
Article
The Regulatory Role of miRNAs in Zebrafish Fin Regeneration
by Jiaqi Fan, Xinya Liu, Ziheng Duan, Hanya Zhao, Zhongjie Chang and Li Li
Int. J. Mol. Sci. 2024, 25(19), 10542; https://doi.org/10.3390/ijms251910542 - 30 Sep 2024
Viewed by 309
Abstract
Since Teleostei fins have a strong regenerative capacity, further research was conducted on the regulation of gene expression during fin regeneration. This research focuses on miRNA, which is a key post-transcriptional regulatory molecule. In this study, a miRNA library for the fin regeneration [...] Read more.
Since Teleostei fins have a strong regenerative capacity, further research was conducted on the regulation of gene expression during fin regeneration. This research focuses on miRNA, which is a key post-transcriptional regulatory molecule. In this study, a miRNA library for the fin regeneration of zebrafish was constructed to reveal the differential expression of miRNA during fin regeneration and to explore the regulatory pathway for fin regeneration. Following the injection of miRNA agomir into zebrafish, the proliferation of blastema cells and the overall fin regeneration area were significantly reduced. It was observed that the miRNAs impaired blastocyte formation by affecting fin regeneration through the inhibition of the expressions of genes and proteins associated with blastocyte formation (including yap1 and Smad1/5/9), which is an effect associated with the Hippo pathway. Furthermore, it has been demonstrated that miRNAs can impair the patterns and mineralization of newly formed fin rays. The miRNAs influenced fin regeneration by inhibiting the expression of a range of bone-related genes and proteins in osteoblast lineages, including sp7, runx2a, and runx2b. This study provides a valuable reference for the further exploration of morphological bone reconstruction in aquatic vertebrates. Full article
(This article belongs to the Section Molecular Biology)
Show Figures

Figure 1

Figure 1
<p>Differential expression analysis of miRNA in caudal fin regeneration. (<b>A</b>) Differential expression of miRNAs in the five fin regeneration stages. (<b>B</b>) Distribution of the differently expressed miRNAs. (<b>C</b>) The top 12 differential expression miRNAs.</p>
Full article ">Figure 1 Cont.
<p>Differential expression analysis of miRNA in caudal fin regeneration. (<b>A</b>) Differential expression of miRNAs in the five fin regeneration stages. (<b>B</b>) Distribution of the differently expressed miRNAs. (<b>C</b>) The top 12 differential expression miRNAs.</p>
Full article ">Figure 2
<p>In vitro validation of the targeting relationship between miRNA and predicted target genes. (<b>A</b>) Differential expressions of miR-338, miR-375, miR-218a, and miR-145-5p in the four stages. (<b>B</b>) Luciferase assays of miR-mimics and miR-NC co-transfected with WT and MT plasmids in HEK293T cells. Significant differences (<span class="html-italic">p</span> &lt; 0.05) between the treatment and control groups are indicated by asterisks above the bars.</p>
Full article ">Figure 3
<p>Analysis of phenotypic changes in caudal fin regeneration following miRNA injection. (<b>A</b>) Regenerative phenotype of the 4 dpa regeneration group and the miRNA injection group. (<b>B</b>) Percentage of the total caudal fin area of each group (<span class="html-italic">n</span> = 4 fish in each group). (<b>C</b>) CT presentation of 4 dpa regeneration group and miRNAs injection group. Significant differences (<span class="html-italic">p</span> &lt; 0.05) between treatment and control groups are indicated by asterisks above the bars.</p>
Full article ">Figure 4
<p>miRNA regulates fin regeneration by inhibiting the Hippo-Yap pathway. (<b>A</b>) Quantitative real-time PCR analysis of <span class="html-italic">yap1</span> mRNA of the caudal fin. (<b>B</b>) In situ hybridization on cryosections at 4 dpa (1 dpi) illustrating a severe reduction in <span class="html-italic">yap1</span> expression located at the distal blastema under miRNA-mimics injection in contrast to the control group. (<b>C</b>) Immunodetection of osteoblasts using the Yap1 osteoblast-specific antibody on longitudinal sections of fins regenerates at 4 dpa (1 dpi) after miR-375 injection and control. In the control rays, more basal cells are found on the surfaces of the distal lepidotrichia. (C#) Magnification of yellow dotted box in C. (<b>D</b>) Immunodetection of osteoblasts using the Smad1/5/9 osteoblast-specific antibody on longitudinal sections of regenerated fins at 4 dpa (1 dpi) after miR-375 injection and control. In the control rays, more basal cells are found on the surfaces of the distal lepidotrichia. (D#) Magnification of yellow dotted box in D. (<b>E</b>) Fluorescent detection of DNA-replicating cells in regenerated fin at 4 dpa, 12 h following 5-Ethynyl-20-deoxyuridine (EdU) injection and miR-375 injection, and the control at 1 dpa. (<b>F</b>) Statistical analysis of EdU-positive cells in sections of fin rays (<span class="html-italic">n</span> = 4 rays and 6 sections). (<b>G</b>) Analysis of the relative expression changes in upstream and downstream <span class="html-italic">yap1</span>-related genes during miR-375 injection. Dashed lines indicate the amputation plane. Plot values represent mean ± s.d. Significant differences (<span class="html-italic">p</span> &lt; 0.05) between treatment and control groups are indicated by asterisks above the bars. Scale bar: 100 μm (<b>B</b>–<b>F</b>).</p>
Full article ">Figure 4 Cont.
<p>miRNA regulates fin regeneration by inhibiting the Hippo-Yap pathway. (<b>A</b>) Quantitative real-time PCR analysis of <span class="html-italic">yap1</span> mRNA of the caudal fin. (<b>B</b>) In situ hybridization on cryosections at 4 dpa (1 dpi) illustrating a severe reduction in <span class="html-italic">yap1</span> expression located at the distal blastema under miRNA-mimics injection in contrast to the control group. (<b>C</b>) Immunodetection of osteoblasts using the Yap1 osteoblast-specific antibody on longitudinal sections of fins regenerates at 4 dpa (1 dpi) after miR-375 injection and control. In the control rays, more basal cells are found on the surfaces of the distal lepidotrichia. (C#) Magnification of yellow dotted box in C. (<b>D</b>) Immunodetection of osteoblasts using the Smad1/5/9 osteoblast-specific antibody on longitudinal sections of regenerated fins at 4 dpa (1 dpi) after miR-375 injection and control. In the control rays, more basal cells are found on the surfaces of the distal lepidotrichia. (D#) Magnification of yellow dotted box in D. (<b>E</b>) Fluorescent detection of DNA-replicating cells in regenerated fin at 4 dpa, 12 h following 5-Ethynyl-20-deoxyuridine (EdU) injection and miR-375 injection, and the control at 1 dpa. (<b>F</b>) Statistical analysis of EdU-positive cells in sections of fin rays (<span class="html-italic">n</span> = 4 rays and 6 sections). (<b>G</b>) Analysis of the relative expression changes in upstream and downstream <span class="html-italic">yap1</span>-related genes during miR-375 injection. Dashed lines indicate the amputation plane. Plot values represent mean ± s.d. Significant differences (<span class="html-italic">p</span> &lt; 0.05) between treatment and control groups are indicated by asterisks above the bars. Scale bar: 100 μm (<b>B</b>–<b>F</b>).</p>
Full article ">Figure 5
<p>miRNAs regulate fin regeneration by inhibiting osteoblast differentiation. (<b>A</b>,<b>B</b>) Morphology of fin rays following miR-mimic injection. Analysis of mineralization progression using Alizarin red staining shows a reduction in matrix mineralization in rays 3–8 of fish injected with miRNA mimics at 8 dpa compared with control rays. Dashed lines indicate the amputation plane. (<b>C</b>) Quantitative real-time PCR analysis of <span class="html-italic">sp7</span> mRNA of the caudal fin. (<b>D</b>) In situ hybridization on cryosections at 4 dpa (1 dpi) illustrating a severe reduction in <span class="html-italic">sp7</span> expression along the distal-proximal axis in miRNA-mimic-injected groups compared to control groups. (<b>E</b>) Immunodetection of osteoblasts using Sp7 osteoblast-specific antibody on longitudinal sections of regenerated fins at 4 dpa (1 dpi) after miR-338 injection of control. In the control rays, more mature bone-secreting cells are found on the surfaces of the proximal lepidotrichia. (E#) Magnification of yellow dotted box in E. (<b>F</b>) Quantitative real-time PCR analysis of <span class="html-italic">runx2a</span> and <span class="html-italic">runx2b</span> mRNAs of caudal fin. (<b>G</b>) In situ hybridization on cryosections at 4 dpa (1 dpi) illustrating a severe reduction in <span class="html-italic">runx2a</span> expression along the distal–proximal axis in the miRNA-218a injected group compared to the control group. (<b>H</b>) In situ hybridization on cryosections at 4 dpa (1 dpi) illustrating a severe reduction in <span class="html-italic">runx2b</span> expression along the distal–proximal axis in miRNA-145-5p injected group compared to the control group. (<b>I</b>) Immunodetection of osteoblasts using the Runx2 osteoblast-specific antibody on longitudinal sections of regenerated fins at 4 dpa (1 dpi) after miR mimic injection and control. In the control rays, more mature bone-secreting cells are found on the surfaces of the proximal lepidotrichia. (I#) Magnification of yellow dotted box in (<b>I</b>). Dashed lines indicate the amputation plane. Plot values represent mean ± s.d. Significant differences (<span class="html-italic">p</span> &lt; 0.05) between treatment and control groups are indicated by asterisks above the bars. Scale bar: 100 μm (<b>D</b>,<b>E</b>,<b>G</b>–<b>I</b>).</p>
Full article ">
12 pages, 1700 KiB  
Review
Increased Cadmium Load, Vitamin D Deficiency, and Elevated FGF23 Levels as Pathophysiological Factors Potentially Linked to the Onset of Acute Lymphoblastic Leukemia: A Review
by Vuk Djulejic, Ana Ivanovski, Ana Cirovic and Aleksandar Cirovic
J. Pers. Med. 2024, 14(10), 1036; https://doi.org/10.3390/jpm14101036 - 28 Sep 2024
Viewed by 517
Abstract
The preventability of acute lymphocytic leukemia during childhood is currently receiving great attention, as it is one of the most common cancers in children. Among the known risk factors so far are those affecting the development of gut microbiota, such as a short [...] Read more.
The preventability of acute lymphocytic leukemia during childhood is currently receiving great attention, as it is one of the most common cancers in children. Among the known risk factors so far are those affecting the development of gut microbiota, such as a short duration or absence of breastfeeding, cesarean section, a diet lacking in short-chain fatty acids (SCFAs), the use of antibiotics, absence of infection during infancy, and lack of pets, among other factors. Namely, it has been shown that iron deficiency anemia (IDA) and lack of vitamin D may cause intestinal dysbiosis, while at the same time, both increase the risk of hematological malignancies. The presence of IDA and vitamin D deficiency have been shown to lead to a decreased proportion of Firmicutes in stool, which could, as a consequence, lead to a deficit of butyrate. Moreover, children with IDA have increased blood concentrations of cadmium, which induces systemic inflammation and is linked to the onset of an inflammatory microenvironment in the bone marrow. Finally, IDA and Cd exposure increase fibroblast growth factor 23 (FGF23) blood levels, which in turn suppresses vitamin D synthesis. A lack of vitamin D has been associated with a higher risk of ALL onset. In brief, as presented in this review, there are three independent ways in which IDA increases the risk of acute lymphocytic leukemia (ALL) appearance. These are: intestinal dysbiosis, disruption of vitamin D synthesis, and an increased Cd load, which has been linked to systemic inflammation. All of the aforementioned factors could generate the appearance of a second mutation, such as ETV6/RUNX1 (TEL-AML), leading to mutation homozygosity and the onset of disease. ALL has been observed in both IDA and thalassemia. However, as IDA is the most common type of anemia and the majority of published data pertains to it, we will focus on IDA in this review. Full article
(This article belongs to the Section Mechanisms of Diseases)
Show Figures

Figure 1

Figure 1
<p>Mechanisms initiated by iron deficiency anemia which contribute to appearance of ALL.</p>
Full article ">Figure 2
<p>Summary of how iron deficiency anemia (IDA) can promote the onset of ALL in three different ways. The arrows used in this figure do not necessarily represent cause and effect but rather indicate associations between the connected boxes.</p>
Full article ">
12 pages, 2159 KiB  
Article
Genomic Landscape of Myelodysplastic/Myeloproliferative Neoplasms: A Multi-Central Study
by Fei Fei, Amar Jariwala, Sheeja Pullarkat, Eric Loo, Yan Liu, Parastou Tizro, Haris Ali, Salman Otoukesh, Idoroenyi Amanam, Andrew Artz, Feras Ally, Milhan Telatar, Ryotaro Nakamura, Guido Marcucci and Michelle Afkhami
Int. J. Mol. Sci. 2024, 25(18), 10214; https://doi.org/10.3390/ijms251810214 - 23 Sep 2024
Viewed by 631
Abstract
The accurate diagnosis and classification of myelodysplastic/myeloproliferative neoplasm (MDS/MPN) are challenging due to the overlapping pathological and molecular features of myelodysplastic syndrome (MDS) and myeloproliferative neoplasm (MPN). We investigated the genomic landscape in different MDS/MPN subtypes, including chronic myelomonocytic leukemia (CMML; n = [...] Read more.
The accurate diagnosis and classification of myelodysplastic/myeloproliferative neoplasm (MDS/MPN) are challenging due to the overlapping pathological and molecular features of myelodysplastic syndrome (MDS) and myeloproliferative neoplasm (MPN). We investigated the genomic landscape in different MDS/MPN subtypes, including chronic myelomonocytic leukemia (CMML; n = 97), atypical chronic myeloid leukemia (aCML; n = 8), MDS/MPN-unclassified (MDS/MPN-U; n = 44), and MDS/MPN with ring sideroblasts and thrombocytosis (MDS/MPN-RS-T; n = 12). Our study indicated that MDS/MPN is characterized by mutations commonly identified in myeloid neoplasms, with TET2 (52%) being the most frequently mutated gene, followed by ASXL1 (38.7%), SRSF2 (34.7%), and JAK2 (19.7%), among others. However, the distribution of recurrent mutations differs across the MDS/MPN subtypes. We confirmed that specific gene combinations correlate with specific MDS/MPN subtypes (e.g., TET2/SRSF2 in CMML, ASXL1/SETBP1 in aCML, and SF3B1/JAK2 in MDS/MPN-RS-T), with MDS/MPN-U being the most heterogeneous. Furthermore, we found that older age (≥65 years) and mutations in RUNX1 and TP53 were associated with poorer clinical outcomes in CMML (p < 0.05) by multivariate analysis. In MDS/MPN-U, CBL mutations (p < 0.05) were the sole negative prognostic factors identified in our study by multivariate analysis (p < 0.05). Overall, our study provides genetic insights into various MDS/MPN subtypes, which may aid in diagnosis and clinical decision-making for patients with MDS/MPN. Full article
(This article belongs to the Section Molecular Pathology, Diagnostics, and Therapeutics)
Show Figures

Figure 1

Figure 1
<p>The workflow and study design of our cohort. BM, bone marrow; PB, peripheral blood. (Abbreviations: aCML, atypical myeloid leukemia; AML, acute myeloid leukemia; CMML, chronic myelomonocytic leukemia; MDS/MPN-U, myelodysplastic/myeloproliferative neoplasm-unclassified; and MDS/MPN-RS-T, myelodysplastic/myeloproliferative neoplasm with ring sideroblasts and thrombocytosis.)</p>
Full article ">Figure 2
<p>Frequency of recurrent gene mutations in all myelodysplastic/myeloproliferative neoplasm (MDS/MPN) patients (n = 173).</p>
Full article ">Figure 3
<p>Molecular and cytogenetic characteristics among the different MDS/MPN subtypes (n = 173). An oncoplot showing the mutated genes among the different MDS/MPN subtypes. Each column represents a patient. Thirty-one genes are grouped into eight categories based on their functions: DNA methylation, chromatin modification, RNA splicing, transcription factors, receptor kinases, cohesion, RAS pathways, and others. Green depicts the different MDS/MPN subtypes: CMML, CMML-AML, aCML, MDS/MPN-U, and MDS/MPN-RS-T. Red depicts a single gene mutation; purple depicts more than one mutation in the same gene, mainly corresponding to biallelic <span class="html-italic">TET2</span> mutations. Cytogenetic findings are divided into three groups: normal karyotype, abnormal karyotype, and complex karyotype. Myelofibrosis (MF) status is divided into five groups: MF 0, MF 1, MF 2, MF 3, and N/A. The frequency of recurrent gene mutations among the different MDS/MPN subtypes. (Abbreviations: aCML, atypical myeloid leukemia; AML, acute myeloid leukemia; CMML, chronic myelomonocytic leukemia; MDS/MPN-U, myelodysplastic/myeloproliferative neoplasm-unclassified; MDS/MPN-RS-T, myelodysplastic/myeloproliferative neoplasm with ring sideroblasts and thrombocytosis; MF, myelofibrosis; and N/A, not applicable.)</p>
Full article ">Figure 4
<p>Frequency of mutations based on functional classification among different MDS/MPN subtypes. (<b>A</b>) CMML (n = 97); (<b>B</b>) CMML-AML (n = 12); (<b>C</b>) aCML (n = 8); (<b>D</b>) MDS/MPN-U (n = 44); and (<b>E</b>) MDS/MPN-RS-T (n = 12). (Abbreviations: aCML, atypical myeloid leukemia; AML, acute myeloid leukemia; CMML, chronic myelomonocytic leukemia; MDS/MPN-U, myelodysplastic/myeloproliferative neoplasm-unclassified; and MDS/MPN-RS-T, myelodysplastic/myeloproliferative neoplasm with ring sideroblasts and thrombocytosis).</p>
Full article ">
19 pages, 3310 KiB  
Review
Regulation of Skeletal Development and Maintenance by Runx2 and Sp7
by Toshihisa Komori
Int. J. Mol. Sci. 2024, 25(18), 10102; https://doi.org/10.3390/ijms251810102 - 20 Sep 2024
Viewed by 678
Abstract
Runx2 (runt related transcription factor 2) and Sp7 (Sp7 transcription factor 7) are crucial transcription factors for bone development. The cotranscription factor Cbfb (core binding factor beta), which enhances the DNA-binding capacity of Runx2 and stabilizes the Runx2 protein, is necessary for bone [...] Read more.
Runx2 (runt related transcription factor 2) and Sp7 (Sp7 transcription factor 7) are crucial transcription factors for bone development. The cotranscription factor Cbfb (core binding factor beta), which enhances the DNA-binding capacity of Runx2 and stabilizes the Runx2 protein, is necessary for bone development. Runx2 is essential for chondrocyte maturation, and Sp7 is partly involved. Runx2 induces the commitment of multipotent mesenchymal cells to osteoblast lineage cells and enhances the proliferation of osteoprogenitors. Reciprocal regulation between Runx2 and the Hedgehog, fibroblast growth factor (Fgf), Wnt, and parathyroid hormone-like hormone (Pthlh) signaling pathways and Dlx5 (distal-less homeobox 5) plays an important role in these processes. The induction of Fgfr2 (Fgf receptor 2) and Fgfr3 expression by Runx2 is important for the proliferation of osteoblast lineage cells. Runx2 induces Sp7 expression, and Runx2+ osteoprogenitors become Runx2+Sp7+ preosteoblasts. Sp7 induces the differentiation of preosteoblasts into osteoblasts without enhancing their proliferation. In osteoblasts, Runx2 is required for bone formation by inducing the expression of major bone matrix protein genes, including Col1a1 (collagen type I alpha 1), Col1a2, Spp1 (secreted phosphoprotein 1), Ibsp (integrin binding sialoprotein), and Bglap (bone gamma carboxyglutamate protein)/Bglap2. Bglap/Bglap2 (osteocalcin) regulates the alignment of apatite crystals parallel to collagen fibrils but does not function as a hormone that regulates glucose metabolism, testosterone synthesis, and muscle mass. Sp7 is also involved in Co1a1 expression and regulates osteoblast/osteocyte process formation, which is necessary for the survival of osteocytes and the prevention of cortical porosity. SP7 mutations cause osteogenesis imperfecta in rare cases. Runx2 is an important pathogenic factor, while Runx1, Runx3, and Cbfb are protective factors in osteoarthritis development. Full article
(This article belongs to the Section Molecular Biology)
Show Figures

Figure 1

Figure 1
<p>Functions of Runx2 and Sp7 in the proliferation and differentiation of osteoblast-lineage cells and chondrocytes: Runx2 induces the proliferation of osteoblast-lineage cells and chondrocytes and the commitment of mesenchymal stem cells to osteoprogenitors and their differentiation. Ihh is required for Runx2 expression in perichondrial mesenchymal cells during endochondral ossification. Runx2 induces the maturation of chondrocytes and the expression of Vegfa, Spp1, Ibsp, and Mmp13 in terminal hypertrophic chondrocytes and is required for the transdifferentiation of terminal hypertrophic chondrocytes into osteoblasts. Runx2 induces the expression of Ihh, which enhances chondrocyte proliferation, in prehypertrophic chondrocytes, and Ihh induces the expression of Pthlh, which inhibits Runx2 expression through Pthr1, forming a negative feedback loop. Sp7 induces the differentiation of preosteoblasts into osteoblasts and chondrocyte maturation.</p>
Full article ">Figure 2
<p>Functions of Runx2 and Sp7 in differentiated osteoblasts and osteocytes: Runx2 enhances the proliferation of immature osteoblasts, whereas Sp7 inhibits it, and both are required for osteoblast maturation. Runx2 induces the expression of Spp1 and Ibsp in immature osteoblasts, while Runx2 and Sp7 induce the expression of Col1a1. Runx2 also induces the expression of Bglap/Bglap2 (osteocalcin; Ocn) in mature osteoblasts, and Galnt<span class="html-italic">3</span> and Fgf23, the proteins that regulate phosphorus homeostasis, in osteocytes. Ocn is required for the alignment of apatite crystals parallel to collagen fibrils. Ocn does not regulate bone mass, glucose metabolism, testosterone synthesis, or muscle mass. Sp7 regulates the formation of osteoblast/osteocyte processes, and a reduction in the number of osteocyte processes causes osteocyte apoptosis and secondary necrosis, leading to cortical porosity.</p>
Full article ">Figure 3
<p>Pathogenic and protective features of Runx family transcription factors in OA: Chondrocyte maturation (hypertrophy) is a major cause of OA. Chondrocyte maturation leads to an increase in type X collagen and the expression of enzymes such as Mmp13 and Adamts5, which digest type II collagen and Acan, respectively, and a decrease in Prg4, type II collagen, and Acan. Runx2 induces chondrocyte maturation and the expression of Mmp13 and Adamts5. Yap1 (yes-associated protein 1) inhibits Runx2 and Fermt2 (fermitin family member 2, Kindlin-2) inhibits Stat3, which induces Runx2 expression. The activation of Mapk1/3 (mitogen-activated protein kinase 1/3) activates Runx2 and inhibits Yap1. Runx2 interacts with AP1 or Cebpb and induces Mmp13 expression. Runx1 increases Col2a1 expression through protein–protein interactions with Sox5, Sox6, and Sox9, and by activating Tgfβ signaling. Runx1 also inhibits chondrocyte maturation by inducing the expression of Nkx3-2 (NK3 homeobox 2) and increasing Yap protein levels, and it inhibits Wnt/β-catenin signaling, the excessive activation of which is associated with progressive joint damage [<a href="#B122-ijms-25-10102" class="html-bibr">122</a>]. Runx3 directly induces Prg4 and Acan expression. Cbfb stabilizes Runx1 and Runx3 proteins and protects against OA progression.</p>
Full article ">
14 pages, 2977 KiB  
Article
HIV Modulates Osteoblast Differentiation via Upregulation of RANKL and Vitronectin
by Rosa Nicole Freiberger, Cynthia Alicia Marcela López, María Belén Palma, Cintia Cevallos, Franco Agustin Sviercz, Patricio Jarmoluk, Marcela Nilda García, Jorge Quarleri and M. Victoria Delpino
Pathogens 2024, 13(9), 800; https://doi.org/10.3390/pathogens13090800 - 15 Sep 2024
Viewed by 548
Abstract
Bone loss is a prevalent characteristic among people with HIV (PWH). We focused on mesenchymal stem cells (MSCs) and osteoblasts, examining their susceptibility to different HIV strains (R5- and X4-tropic) and the subsequent effects on bone tissue homeostasis. Our findings suggest that MSCs [...] Read more.
Bone loss is a prevalent characteristic among people with HIV (PWH). We focused on mesenchymal stem cells (MSCs) and osteoblasts, examining their susceptibility to different HIV strains (R5- and X4-tropic) and the subsequent effects on bone tissue homeostasis. Our findings suggest that MSCs and osteoblasts are susceptible to R5- and X4-tropic HIV but do not support productive HIV replication. HIV exposure during the osteoblast differentiation process revealed that the virus could not alter mineral and organic matrix deposition. However, the reduction in runt-related transcription factor 2 (RUNX2) transcription, the increase in the transcription of nuclear receptor activator ligand kappa B (RANKL), and the augmentation of vitronectin deposition strongly suggested that X4- and R5-HIV could affect bone homeostasis. This study highlights the HIV ability to alter MSCs’ differentiation into osteoblasts, critical for maintaining bone and adipose tissue homeostasis and function. Full article
(This article belongs to the Special Issue Infections and Bone Damage)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Expression of CD4, CCR5, and CXCR4 on the cell surface. Representative dot plots obtained by flow cytometry indicating the surface marker expression of CD4, CCR5, and CXCR4 in mesenchymal stem cells (MSCs) (<b>A</b>), and at 7 and 14 days of the osteoblast differentiation process (<b>B</b>). The bars indicate the percentage of positive cells in <b>A</b> (<b>C</b>) and in <b>B</b> (<b>D</b>). Data are expressed as mean ± SD obtained from 3 independent experiments.</p>
Full article ">Figure 2
<p>Exposure of MSCs and osteoblasts to X4- and R5-tropic HIV. MSCs and 14 days’ differentiated osteoblasts were infected with an inoculum of 1 pg of p24/cell with CXCR4-tropic HIV (HIV (X4)) and CCR5-tropic HIV (HIV (R5)). Representative microscopy images showing the kinetics of HIV replication at 3, 5, and 7 days post-infection (dpi) by the immunostaining of HIV-p24 capsid antigen (<b>A</b>). Bars express the percentage of HIV-infected MSCs (<b>B</b>) and osteoblasts (<b>C</b>). Ten microscopic fields per condition were quantified for each experiment. Scale bar: 50 µm. Data are expressed as mean ± SD obtained from 4 independent experiments.</p>
Full article ">Figure 3
<p>HIV was unable to modulate osteoblast differentiation. Effect of CXCR4-tropic HIV (HIV (X4)) and CCR5-tropic HIV (HIV (R5)) exposure on osteoblast differentiation. Representative microscopy images reveal alkaline phosphatase (ALP) activity by deposition of BCIP-NTB substrate, calcium deposition by alizarin red S staining, and collagen deposition by Sirius red staining at 7, 14, and 21 days post-differentiation (<b>A</b>). Spectrophotometric quantification of ALP activity, calcium, and collagen deposition (<b>B</b>). NI (noninfected). d (days post-differentiation). Ten microscopic fields per condition were quantified for each experiment. Scale bar: 200 µm. Data are expressed as mean ± SD obtained from 4 independent experiments.</p>
Full article ">Figure 4
<p>HIV modulates RUNX2, vitronectin, and RANKL expression during osteoblast differentiation. RUNX2 transcription was determined by RT-qPCR at day 1, 7, 14, and 21 post-differentiation (<b>A</b>). Vitronectin expression revealed by immunofluorescence with a specific antibody at 14 and 21 days post-differentiation (<b>B</b>). Quantification of median fluorescence intensity (MFI) using ImageJ from images in A at 14 (<b>C</b>) and 21 days post-differentiation (<b>D</b>). Adhesion of CellTrace<sup>TM</sup> yellow-labeled U937 monocytes to differentiated osteoblasts previously infected or not with HIV-R5 or HIV-X4, uninfected osteoblast, and osteoblast without U937 (negative control, NC). The presence of adherent cells was determined by fluorescence microscopy (RED). DIC, differential interference contrast (<b>E</b>). Quantification of VPD-positive cells from images in <b>E</b> (<b>F</b>). *** <span class="html-italic">p</span> &lt; 0.0001; ** <span class="html-italic">p</span> &lt; 0.001; * <span class="html-italic">p</span> &lt; 0.01 vs. non-infected (NI). RANKL transcription determined by RT-qPCR (<b>G</b>). RANKL expression in culture supernatants (<b>H</b>). NI (noninfected). d (days post-differentiation). Ten microscopic fields per condition were quantified for each experiment. Scale bar: 50 µm. Data are expressed as mean ± SD obtained from 3 independent experiments. * <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 vs. NI.</p>
Full article ">
22 pages, 4974 KiB  
Article
E3 Ubiquitin Ligase CHIP Inhibits Haemocyte Proliferation and Differentiation via the Ubiquitination of Runx in the Pacific Oyster
by Miren Dong, Ying Song, Weilin Wang, Xiaorui Song, Wei Wu, Lingling Wang and Linsheng Song
Cells 2024, 13(18), 1535; https://doi.org/10.3390/cells13181535 - 13 Sep 2024
Viewed by 566
Abstract
Mollusca first evolve primitive immune cells (namely, haemocytes), which assemble a notable complex innate immune system, which are continuously produced through proliferation and differentiation and infused in the haemolymph. As a typical E3 ligase, CHIP is critical for immune cell turnover and homeostasis [...] Read more.
Mollusca first evolve primitive immune cells (namely, haemocytes), which assemble a notable complex innate immune system, which are continuously produced through proliferation and differentiation and infused in the haemolymph. As a typical E3 ligase, CHIP is critical for immune cell turnover and homeostasis in vertebrates. In this study, a CHIP homolog (CgCHIP) with a high expression in haemocytes was identified in oysters to investigate its role in the proliferation and differentiation of ancient innate immune cells. CgCHIP exhibited a widespread distribution across all haemocyte subpopulations, and the knockdown of CgCHIP altered the composition of haemocytes as examined by flow cytometry. Mechanistically screened with bioinformatics and immunoprecipitation, a key haematopoietic transcription factor CgRunx was identified as a substrate of CgCHIP. Moreover, amino acids in the interacted intervals of CgCHIP and CgRunx were determined by molecular docking. Experimental evidence from an in vitro culture model of an agranulocyte subpopulation and an in vivo oyster model revealed that the knockdown of CgCHIP and CgRunx had opposing effects on agranulocyte (precursor cells) differentiation and granulocyte (effector cells) proliferation. In summary, CgCHIP negatively regulated agranulocyte differentiation and granulocyte proliferation by mediating the ubiquitination and degradation of CgRunx in oysters. These results offer insight into the involvement of ubiquitylation in controlling haemocyte turnover in primitive invertebrates. Full article
Show Figures

Figure 1

Figure 1
<p>Evolutionary properties of ubiquitin E3 ligase CHIP from oyster <span class="html-italic">C</span>. <span class="html-italic">gigas</span>. (<b>A</b>) Domain and tertiary structure prediction of CHIP from oyster <span class="html-italic">C</span><b>.</b> <span class="html-italic">gigas</span> by SMART and SWISS-MODEL program. (<b>B</b>) Domain and tertiary structure prediction of CHIP from <span class="html-italic">Homo sapiens</span> by SMART and SWISS-MODEL program. The pink box indicates a low complexity domain. (<b>C</b>) Multisequence alignment analysis of <span class="html-italic">Cg</span>CHIP with its homologues from other vertebrate and invertebrate species. Amino acids with 100% identity are in black, and similar amino acids are in gray. (<b>D</b>) A phylogenetic tree for CHIP was constructed with the amino acid sequences from the indicated species including <span class="html-italic">H. sapiens</span>, <span class="html-italic">M. musculus</span>, <span class="html-italic">D. rerio</span>, <span class="html-italic">L. anatine</span>, <span class="html-italic">D. melanogaster</span>, <span class="html-italic">A. californica</span>, <span class="html-italic">B. glabrata</span>, <span class="html-italic">M. yessoensis</span>, <span class="html-italic">C</span>. <span class="html-italic">gigas</span>, <span class="html-italic">C. virginica</span>, and <span class="html-italic">C. elegans</span>. The trees were constructed using the neighbor-joining (NJ) algorithm in the Mega 6.0 program based on multiple sequence alignment by ClustalW. Bootstrap values of 1000 replicates (%) are indicated for the branches. CHIP from <span class="html-italic">C</span>. <span class="html-italic">gigas</span> was marked with a grey arrow.</p>
Full article ">Figure 2
<p><span class="html-italic">Cg</span>CHIP is highly expressed in oyster haemocytes and alters the proportion of their three subpopulations. (<b>A</b>) The mRNA transcripts of <span class="html-italic">Cg</span>CHIP in the indicated tissues and haemocytes examined by qRT-PCR, normalized to <span class="html-italic">Cg</span>EF1-α. Hep: hepatopancreas; Man: mantle; Gon: gonad; Amu: adductor muscle; Lap: labial palp; Gil: gill; Hae: haemocytes. <span class="html-italic">p</span>-values, <sup>a</sup> <span class="html-italic">p</span> &gt; 0.05, <sup>b</sup> <span class="html-italic">p</span> &lt; 0.05, and <sup>c</sup> <span class="html-italic">p</span> &lt; 0.01, were calculated using a one-way ANOVA with Dunnett’s correction for multiple comparisons. (<b>B</b>) Relative temporal levels of <span class="html-italic">Cg</span>CHIP mRNA in haemocytes with or without <span class="html-italic">V. splendidus</span> infection examined by qRT-PCR, normalized to <span class="html-italic">Cg</span>EF1-α. Error bars show mean ± standard deviation. <span class="html-italic">p</span>-values, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, were calculated using a two-tailed, unpaired <span class="html-italic">t</span>-test. Error bars show mean ± standard deviation (<span class="html-italic">n</span> = 3). (<b>C</b>) SDS-PAGE analysis showed the recombinant-<span class="html-italic">Cg</span>CHIP (r<span class="html-italic">Cg</span>CHIP) proteins. Lane M: protein molecular marker; Lane 1: negative control (without IPTG induction); Lane 2: induced recombinant protein with IPTG; Lane 3: purified r<span class="html-italic">Cg</span>CHIP protein. (<b>D</b>) The specificity of the <span class="html-italic">Cg</span>CHIP polyclonal antibody determined by Western blotting. Lane M: protein molecular marker; Lane 1: in vitro recombinant proteins; Lane 2: haemocyte lysate. (<b>E</b>) Transcriptome data analysis shows the mRNA transcripts of <span class="html-italic">Cg</span>CHIP in the three haemocyte subpopulations (<span class="html-italic">n</span> = 7). <span class="html-italic">p</span>-values, * <span class="html-italic">p</span> &lt; 0.05, were calculated using a one-way ANOVA with Dunnett’s correction for multiple comparisons. ns indicates no significant difference. (<b>F</b>) Haemocytes collected from oyster haematocoel, and morphology observed under confocal. (<b>G</b>) Haemocytes observed following Giemsa staining. (<b>H</b>) Three subpopulations of haemocytes morphologically identified and separated as agranulocytes (A), semi-granulocytes (SG), and granulocytes (G), by flow cytometry. (<b>I</b>) Representative immunofluorescence image shows the localization of <span class="html-italic">Cg</span>CHIP (green) in haemocytes and the nuclei stained with DAPI (blue). The localization region marked with yellow circles. (<b>J</b>) Bar graph shows the mean fluorescence intensity of <span class="html-italic">Cg</span>CHIP in the three haemocyte subpopulations. The per cell compartment was outlined, and the fluorescence intensity of positive signals within per cell was measured using ImageJ software. For each haemocyte subpopulation, the mean fluorescence value of ten cells from five fields were calculated as one replicate, and there were three replicates (<span class="html-italic">n</span> = 3). Abbreviations: Ara: agranulocytes; Semi-gra: semi-granulocytes; and Gra: granulocytes. (<b>K</b>) The percentages of three subpopulations in total haemocytes measured by flow cytometry (<span class="html-italic">n</span> = 3). (<b>L</b>) The bar graph shows the percentage of three haemocyte subpopulations (<span class="html-italic">n</span> = 3). <span class="html-italic">p</span>-values, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, were calculated using a two-tailed, unpaired <span class="html-italic">t</span>-test.</p>
Full article ">Figure 3
<p><span class="html-italic">Cg</span>CHIP targets the <span class="html-italic">Cg</span>Runx protein. (<b>A</b>) CHIP was known to interact with Runx1 as predicted by UbiBrowser 2.0 program. (<b>B</b>) The typical and conserved ubiquitination sites within the Runx protein. The Runt domain is marked with a red box. A Met-1 ubiquitination site is marked with a blue asterisk. Four conserved Lys ubiquitination sites are labeled with a red triangle. (<b>C</b>) Co-IP-based interaction detection of <span class="html-italic">Cg</span>CHIP and <span class="html-italic">Cg</span>Runx in oyster haemocytes. (<b>D</b>) Docking model analysis of <span class="html-italic">Cg</span>CHIP and <span class="html-italic">Cg</span>Runx. (<b>E</b>) The binding coefficients of <span class="html-italic">Cg</span>CHIP and <span class="html-italic">Cg</span>Runx protein interaction sites. (<b>F</b>) Ubiquitination activity of <span class="html-italic">Cg</span>CHIP detected with Western blotting in vitro. (<b>G</b>) <span class="html-italic">Cg</span>Runx ubiquitination assessed by Western blotting. (<b>H</b>) The levels of <span class="html-italic">Cg</span>Runx in oyster haemocytes treated with MG132 (20 μM), quantified by Western blotting.</p>
Full article ">Figure 4
<p><span class="html-italic">Cg</span>CHIP enhances the ubiquitination and degradation of <span class="html-italic">Cg</span>Runx. (<b>A</b>) Representative immunofluorescence image shows the localization of <span class="html-italic">Cg</span>Runx (green) in haemocytes and the nuclei stained with DAPI (blue). (<b>B</b>) Transcriptome data analysis shows the mRNA transcripts of <span class="html-italic">Cg</span>Runx in the three haemocyte subpopulations (<span class="html-italic">n</span> = 7). <span class="html-italic">p</span>-values, * <span class="html-italic">p</span> &lt; 0.05, were calculated using a one-way ANOVA with Dunnett’s correction for multiple comparisons. ns indicates no significant difference. (<b>C</b>) Bar graph shows the mean fluorescence intensity of <span class="html-italic">Cg</span>Runx in the three haemocyte subpopulations. (<b>D</b>) An injection cartoon of dsRNA in the interference assay. (<b>E</b>) The RNAi efficiency of <span class="html-italic">Cg</span>CHIP in haemocytes quantified via qRT-PCR, normalized to <span class="html-italic">Cg</span>EF1-α. Error bars show mean ± standard deviation (<span class="html-italic">n</span> = 3). <span class="html-italic">p</span>-values, ** <span class="html-italic">p</span> &lt; 0.01, were calculated using a two-tailed, unpaired <span class="html-italic">t</span>-test. (<b>F</b>) Protein abundance of <span class="html-italic">Cg</span>CHIP (RNAi efficiency) and <span class="html-italic">Cg</span>Runx examined with Western blotting. (<b>G</b>) Gray analysis of protein band, normalized to β-Tubulin and Histone H3, respectively. <span class="html-italic">p</span>-values, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, were calculated using a two-tailed, unpaired <span class="html-italic">t</span>-test.</p>
Full article ">Figure 5
<p><span class="html-italic">Cg</span>CHIP inhibits agranulocyte differentiation. (<b>A</b>) Schematic of the induced differentiation in cultured agranulocytes. (<b>B</b>) Representative flow cytometry dot-plots show the gated semi-granulocyte and granulocyte populations differentiated from agranulocytes using the agranulocyte differentiation protocol. (<span class="html-italic">n</span> = 3). (<b>C</b>) The bar graph shows the percentage of differentiated agranulocytes (<span class="html-italic">n</span> = 3). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, determined by a two-tailed Student’s <span class="html-italic">t</span>-test. (<b>D</b>,<b>E</b>) Protein expression levels of the proliferative marker <span class="html-italic">Cg</span>PCNA, immature agranulocyte marker <span class="html-italic">Cg</span>Integrin α4, and mature granulocyte marker <span class="html-italic">Cg</span>AATase, in agranulocytes. β-Tubulin was used as an internal control. Error bars show mean ± standard deviation (<span class="html-italic">n</span> = 3). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, determined by a two-tailed Student’s <span class="html-italic">t</span>-test.</p>
Full article ">Figure 6
<p><span class="html-italic">Cg</span>CHIP inhibits granulocyte proliferation. (<b>A</b>) Representative flow cytometry peak diagrams show the proliferation rate of gated EdU labeling agranulocytes in total agranulocytes. (<b>B</b>) The bar graph shows the proliferation rate of agranulocytes (<span class="html-italic">n</span> = 3). (<b>C</b>) Representative flow cytometry peak diagrams showing the proliferation rate of gated EdU labeling granulocytes in total granulocytes. (<b>D</b>) The bar graph shows the proliferation rate of granulocytes (<span class="html-italic">n</span> = 3). * <span class="html-italic">p</span> &lt; 0.05, determined by a two-tailed Student’s <span class="html-italic">t</span> test. (<b>E</b>) Schematic of granulocyte isolation for cell cycle and Western blotting analyses. (<b>F</b>) The percentage changes of granulocytes in different cell cycle phases. (<b>G</b>) The bar graph shows the percentage of agranulocytes in different cell cycle phases (<span class="html-italic">n</span> = 3). Error bars show mean ± standard deviation (<span class="html-italic">n</span> = 3). <span class="html-italic">p</span>-values, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, were calculated using a one-way ANOVA with Dunnett’s correction for multiple comparisons. (<b>H</b>,<b>I</b>) Protein expression levels of proliferative genes <span class="html-italic">Cg</span>Cyclin B1 and <span class="html-italic">Cg</span>CDK2 in granulocytes. β-Tubulin was used as an internal control. Error bars show mean ± standard deviation (<span class="html-italic">n</span> = 3). The data shown are representative of three independent experiments. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, determined by a two-tailed Student’s <span class="html-italic">t</span> test.</p>
Full article ">Figure 7
<p><span class="html-italic">Cg</span>CHIP attenuates phagocytosis in a <span class="html-italic">Cg</span>Runx-dependent manner. (<b>A</b>,<b>B</b>) Representative flow cytometry peak diagrams show the gated phagocytic haemocytes that are defined according to the red positive signal of latex beads. Phagocytic rate in haemocytes was defined by the percentage of phagocytic haemocytes taking in latex beads in total haemocytes. Error bars show mean ± standard deviation (<span class="html-italic">n</span> = 3). <span class="html-italic">p</span>-values were calculated using a one-way ANOVA with Dunnett’s correction for multiple comparisons. The asterisk * and ** indicated a significant difference at <span class="html-italic">p</span> &lt; 0.05 and extremely significant difference at <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 8
<p>A graphical abstract. A conceptual framework for the ubiquitination and degradation of <span class="html-italic">Cg</span>Runx mediated by <span class="html-italic">Cg</span>CHIP, which inhibits the differentiation of agranulocytes and the proliferation of granulocytes in the Pacific oyster <span class="html-italic">C</span>. <span class="html-italic">gigas</span>.</p>
Full article ">
10 pages, 3158 KiB  
Article
Role of Pulsed Electromagnetic Field on Alveolar Bone Remodeling during Orthodontic Retention Phase in Rat Models
by Hafiedz Maulana, Yuyun Yueniwati, Nur Permatasari and Hadi Suyono
Dent. J. 2024, 12(9), 287; https://doi.org/10.3390/dj12090287 - 9 Sep 2024
Viewed by 524
Abstract
Alveolar bone remodeling during the retention phase is essential for successful orthodontic treatment. Pulsed electromagnetic field (PEMF) therapy is an adjunctive therapy for bone-related diseases that induces osteogenesis and prevents bone loss. This study aimed to examine the role of PEMF exposure during [...] Read more.
Alveolar bone remodeling during the retention phase is essential for successful orthodontic treatment. Pulsed electromagnetic field (PEMF) therapy is an adjunctive therapy for bone-related diseases that induces osteogenesis and prevents bone loss. This study aimed to examine the role of PEMF exposure during the retention phase of orthodontic treatment in alveolar bone remodeling. A total of 36 male Wistar rats were divided into control, PEMF 7, and PEMF 14 groups; a 50 g force nickel–titanium closed-coil spring was inserted to create mesial movement in the first molar for 21 d. Furthermore, the spring was removed, and the interdental space was filled with glass ionomer cement. Concurrently, rats were exposed to a PEMF at 15 Hz with a maximum intensity of 2.0 mT 2 h daily, for 7 and 14 days. Afterwards, the cements were removed and the rats were euthanized on days 1, 3, 7, and 14 to evaluate the expression of Wnt5a mRNA and the levels of RANKL, OPG, ALP, and Runx2 on the tension side. The data were analyzed with ANOVA and post hoc tests, with p < 0.05 declared statistically significant. PEMF exposure significantly upregulated Wnt5a mRNA expression, OPG and ALP levels, and Runx2 expression, and decreased RANKL levels in the PEMF 7 and 14 groups compared to the control group (p < 0.05). This study showed that PEMF exposure promotes alveolar bone remodeling during the orthodontic retention phase on the tension side by increasing alveolar bone formation and inhibiting resorption. Full article
Show Figures

Figure 1

Figure 1
<p>Research design. (<b>A</b>) PEMF stimulation phases, (<b>B</b>) orthodontic appliance installation, (<b>C</b>) post orthodontic tooth movement, (<b>D</b>) absorption of GCF sample with paper points, and (<b>E</b>) sampling region (white arrow) for RT-PCR. PEMF: pulsed electromagnetic field, GCF: gingival crevicular fluid, RT-PCR: real-time polymerase chain reaction.</p>
Full article ">Figure 2
<p>PEMF stimulator. (<b>A</b>) The PEMF device and rats were kept in a special fiber cage, placed between a Helmholtz coil and exposed 2 h/day. (<b>B</b>) The waveform was square with a burst width of 5 ms, burst wait of 60 ms, pulse width of 0.2 ms, pulse wait of 0.02 ms, pulse rise of 0.3 μs, and pulse fall of 2.0 μs.</p>
Full article ">Figure 3
<p>The histogram of Wnt5a mRNA expression. *: <span class="html-italic">p</span> &lt; 0.05, significant compared with control group. PEMF: pulsed electromagnetic field.</p>
Full article ">Figure 4
<p>The histogram of RANKL, OPG, and ALP levels. *: <span class="html-italic">p</span> &lt; 0.05, significant compared with control group; #: <span class="html-italic">p</span> &lt; 0.05, significant compared with PEMF 7 group. PEMF: pulsed electromagnetic field, RANKL: receptor activator of nuclear factor-kappa B ligand, OPG: osteoprotegerin, ALP: alkaline phosphatase.</p>
Full article ">Figure 5
<p>Histogram and immunohistochemical image of Runx2 expression. Runx2 positive-osteoblast (black arrow) and the direction of tooth movement (blue arrow). *: <span class="html-italic">p</span> &lt; 0.05, significant compared with control group; #: <span class="html-italic">p</span> &lt; 0.05, significant compared with PEMF 7 group. PEMF: pulsed electromagnetic field, T: tooth, PDL: periodontal ligament, AB: alveolar bone, Runx2: runt-related transcription factor 2.</p>
Full article ">
11 pages, 1597 KiB  
Review
Minimal Requirements for Cancer Initiation: A Comparative Consideration of Three Prototypes of Human Leukemia
by Toshiyuki Hori
Cancers 2024, 16(17), 3109; https://doi.org/10.3390/cancers16173109 - 9 Sep 2024
Viewed by 933
Abstract
Even if its completed form is complex, cancer originates from one or two events that happened to a single cell. A simplified model can play a role in understanding how cancer initiates at the beginning. The pathophysiology of leukemia has been studied in [...] Read more.
Even if its completed form is complex, cancer originates from one or two events that happened to a single cell. A simplified model can play a role in understanding how cancer initiates at the beginning. The pathophysiology of leukemia has been studied in the most detailed manner among all human cancers. In this review, based on milestone papers and the latest research developments in hematology, acute promyelocytic leukemia (APL), chronic myeloid leukemia (CML), and acute myeloid leukemia (AML) with RUNX1-RUNX1T1 are selected to consider minimal requirements for cancer initiation. A one-hit model can be applied to the initiation of APL and CML whereas a two-hit model is more suitable to the initiation of AML with RUNX1-RUNX1T1 and other AMLs. Even in cancer cells with multiple genetic abnormalities, there must be a few mutant genes critical for the mutant clone to survive and proliferate. Such genes should be identified and characterized in each case in order to develop individualized target therapy. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) A model of a normal bone marrow showing the myeloid maturation pathway. HPSC indicated by grey circles are located close to a stroma cell in niche. (<b>B</b>) A model of APL. Red cells represent the mutant clone.</p>
Full article ">Figure 2
<p>A model of chronic phase CML. Red cells represent the mutant clone.</p>
Full article ">Figure 3
<p>A model of the pre-leukemia stage of AML. Red cells represent the mutant clone.</p>
Full article ">Figure 4
<p>A model of the overt stage of AML. Red cells represent the mutant clone.</p>
Full article ">Figure 5
<p>(<b>A</b>) A scheme of a normal colon crypt. Stem cells at the bottom of the crypt give rise to progeny of epithelial cells which proliferate and mature. Terminally differentiated cells are pushed away into the bowel lumen. (<b>B</b>) Red mutant cells proliferate rapidly to form a polyp but can be pushed out in the meantime.</p>
Full article ">
15 pages, 5256 KiB  
Article
Nested-PCR vs. RT-qPCR: A Sensitivity Comparison in the Detection of Genetic Alterations in Patients with Acute Leukemias
by Flávia Melo Cunha de Pinho Pessoa, Marcelo Braga de Oliveira, Igor Valentim Barreto, Anna Karolyna da Costa Machado, Deivide Sousa de Oliveira, Rodrigo Monteiro Ribeiro, Jaira Costa Medeiros, Aurélia da Rocha Maciel, Fabiana Aguiar Carneiro Silva, Lívia Andrade Gurgel, Kaira Mara Cordeiro de Albuquerque, Germison Silva Lopes, Ricardo Parente Garcia Vieira, Jussara Alencar Arraes, Meton Soares de Alencar Filho, André Salim Khayat, Maria Elisabete Amaral de Moraes, Manoel Odorico de Moraes Filho and Caroline Aquino Moreira-Nunes
DNA 2024, 4(3), 285-299; https://doi.org/10.3390/dna4030019 - 6 Sep 2024
Viewed by 584
Abstract
The detection of genetic alterations in patients with acute leukemias is essential for the targeting of more specific and effective therapies. Therefore, the aim of this study was to compare the sensitivity of Nested-PCR and RT-qPCR techniques in the detection of genetic alterations [...] Read more.
The detection of genetic alterations in patients with acute leukemias is essential for the targeting of more specific and effective therapies. Therefore, the aim of this study was to compare the sensitivity of Nested-PCR and RT-qPCR techniques in the detection of genetic alterations in patients with acute leukemias. This study included samples from 117 patients treated at the Fortaleza General Hospital. All samples were submitted to analysis using the Nested-PCR and the RT-qPCR techniques. Acute Myeloid Leukemia (AML) patients’ samples were submitted to the analysis of the following alterations: FLT3-ITD, RUNX1::RUNX1T1, CBFB::MYH11 and PML::RARA; meanwhile, BCR::ABL1, TCF3::PBX1, KMT2A::AFF1, ETV6::RUNX1, and STIL::TAL1 fusions were investigated in the Acute Lymphoblastic Leukemia (ALL) patients’ samples. Throughout the study, 77 patients were diagnosed with AML and 40 with ALL. Among the 77 AML patients, FLT3-ITD, RUNX1::RUNX1T1, PML::RARA, and CBFB::MYH11 were detected in 4, 7, 10 and 8 patients, respectively. Among the 40 ALL patients, the presence of 23 patients with BCR::ABL1 translocation and 9 patients with TCF3::PBX1 translocation was observed through the RT-qPCR methodology. Overall, the present study demonstrated that the RT-qPCR technique presented a higher sensitivity when compared to the Nested-PCR technique at the time of diagnosis of the acute leukemia samples studied. Full article
Show Figures

Figure 1

Figure 1
<p>Molecular alteration frequency in research participants. This figure illustrates how many times each fusion was detected in the analyzed patients.</p>
Full article ">Figure 2
<p>Amplification plots of Acute Myeloid Leukemia patients. This figure illustrates the AML amplification plots in which the researched alterations were detected. (<b>A</b>) Amplification plots indicating the detection of <span class="html-italic">FLT3-ITD</span> (blue and red curves are from BM samples and the orange and green ones are from PB samples). (<b>B</b>) Amplification plots indicating the detection of <span class="html-italic">PML::RARA</span> (pink and orange curves are from BM samples and the green and blue ones are from PB samples). (<b>C</b>) Amplification plots indicating the detection of <span class="html-italic">CBFB::MYH11</span> (blue curves are from BM samples and the green ones are from PB samples). (<b>D</b>) Amplification plots indicating the detection of <span class="html-italic">RUNX1::RUNX1T1</span> (orange curves are from BM samples and the green ones are from PB samples).</p>
Full article ">Figure 3
<p>Amplification plots of Acute Lymphoblastic Leukemia samples. All amplification plots in this figure show the results of two samples in duplicate (bone marrow and peripheral blood) of two ALL patients. The figure’s left side shows the results of two patients samples in which <span class="html-italic">BCR::ABL1</span> alteration was detected; meanwhile, the figure’s right side demonstrates the amplification plots of two patients’ samples in which <span class="html-italic">TCF3::PBX1</span> was detected. (<b>A</b>) <span class="html-italic">BCR::ABL1</span> amplification plots (green and orange curves are from BM samples and the pink and blue ones are from PB samples). (<b>B</b>) <span class="html-italic">ABL1</span> amplification plots of two <span class="html-italic">BCR::ABL1</span> patients (blue and green curves are from BM samples and the red and gray ones are from PB samples). (<b>C</b>) <span class="html-italic">ACTB</span> amplification plots of two <span class="html-italic">BCR::ABL1</span> patients (pink and purple curves are from BM samples and the green and blue ones are from PB samples). (<b>D</b>) <span class="html-italic">TCF3::PBX1</span> amplification plots (green and pink curves are from BM samples and the purple and orange ones are from PB samples). (<b>E</b>) <span class="html-italic">ABL1</span> amplification plots of two <span class="html-italic">TCF3::PBX1</span> patients (pink and orange curves are from BM samples and the purple and blue ones are from PB samples). (<b>F</b>) <span class="html-italic">ACTB</span> amplification plots of two <span class="html-italic">TCF3::PBX1</span> patients (pink and purple curves are from BM samples and the orange and blue ones are from PB samples).</p>
Full article ">Figure 4
<p>Nested-PCR results of Acute Lymphoblastic Leukemia samples. These Nested-PCR results are from the same patient’s samples analyzed in the amplification plot figure. (<b>A</b>) Shows the <span class="html-italic">BCR::ABL1</span> (1 and 7—bone marrow; 4 and 10—peripheral blood), <span class="html-italic">GAPDH</span> (2 and 8—bone marrow; 5 and 11—peripheral blood), and <span class="html-italic">HPRT</span> (3 and 9—bone marrow; 6 and 12—peripheral blood) detection in ALL patients. (<b>B</b>) Shows the TCF3::PBX1 (1 and 7—bone marrow; 4 and 10—peripheral blood), <span class="html-italic">GAPDH</span> (2 and 8—bone marrow; 5 and 11—peripheral blood), and <span class="html-italic">HPRT</span> (3 and 9—bone marrow; 6 and 12—peripheral blood) detection in ALL patients.</p>
Full article ">
Back to TopTop