[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (661)

Search Parameters:
Keywords = time lapse

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
16 pages, 1584 KiB  
Review
Advancements and Challenges in Preimplantation Genetic Testing for Aneuploidies: In the Pathway to Non-Invasive Techniques
by Ana del Arco de la Paz, Carla Giménez-Rodríguez, Aikaterini Selntigia, Marcos Meseguer and Daniela Galliano
Genes 2024, 15(12), 1613; https://doi.org/10.3390/genes15121613 - 17 Dec 2024
Viewed by 430
Abstract
The evolution of preimplantation genetic testing for aneuploidy (PGT-A) techniques has been crucial in assisted reproductive technologies (ARTs), improving embryo selection and increasing success rates in in vitro fertilization (IVF) treatments. Techniques ranging from fluorescence in situ hybridization (FISH) to next-generation sequencing (NGS) [...] Read more.
The evolution of preimplantation genetic testing for aneuploidy (PGT-A) techniques has been crucial in assisted reproductive technologies (ARTs), improving embryo selection and increasing success rates in in vitro fertilization (IVF) treatments. Techniques ranging from fluorescence in situ hybridization (FISH) to next-generation sequencing (NGS) have relied on cellular material extraction through biopsies of blastomeres at the cleavage stage on day three or from trophectoderm (TE) cells of the blastocyst. However, this has raised concerns about its potential impact on embryo development. As a result, there has been growing interest in developing non-invasive techniques for detecting aneuploidies, such as the analysis of blastocoel fluid (BF), spent culture medium (SCM), and artificial intelligence (AI) models. Non-invasive methods represent a promising advancement in PGT-A, offering the ability to detect aneuploidies without compromising embryo viability. This article reviews the evolution and principles of PGT-A, analyzing both traditional techniques and emerging non-invasive approaches, while highlighting the advantages and challenges associated with these methodologies. Furthermore, it explores the transformative potential of these innovations, which could optimize genetic screening and significantly improve clinical outcomes in the field of assisted reproduction. Full article
Show Figures

Figure 1

Figure 1
<p>Evolution of PGT-A techniques.</p>
Full article ">Figure 2
<p>Overview of FISH methodology for chromosomal analysis.</p>
Full article ">Figure 3
<p>Overview of traditional PGT-A techniques following the introduction of WGA.</p>
Full article ">Figure 4
<p>Overview of non-invasive PGT-A methods.</p>
Full article ">
16 pages, 5214 KiB  
Article
Bichromatic Splicing Detector Allows Quantification of THRA1 and THRA2 Splicing Isoforms in Single Cells by Fluorescent Live-Cell Imaging
by Eugenio Graceffo, Elisa Pedersen, Marta Rosário, Heiko Krude and Markus Schuelke
Int. J. Mol. Sci. 2024, 25(24), 13512; https://doi.org/10.3390/ijms252413512 - 17 Dec 2024
Viewed by 292
Abstract
Thyroid hormone receptor alpha (THRα) is a nuclear hormone receptor that binds triiodothyronine (T3) and acts as an important transcription factor in development, metabolism, and reproduction. The coding gene, THRA, has two major splicing isoforms in mammals, THRA1 and THRA2 [...] Read more.
Thyroid hormone receptor alpha (THRα) is a nuclear hormone receptor that binds triiodothyronine (T3) and acts as an important transcription factor in development, metabolism, and reproduction. The coding gene, THRA, has two major splicing isoforms in mammals, THRA1 and THRA2, which encode THRα1 and THRα1, respectively. The better characterized isoform, THRα1, is a transcriptional stimulator of genes involved in cell metabolism and growth. The less well-characterized isoform, THRα2, lacks the ligand-binding domain (LBD) and may act as an inhibitor of THRα1 activity. Thus, the ratio of THRα1 to THRα2 isoforms is critical for transcriptional regulation in various tissues and during development and may be abnormal in a number of thyroid hormone resistance syndromes. However, the complete characterization of the THRα isoform expression pattern in healthy human tissues, and especially the study of changes in the ratio of THRα1 to THRα2 in cultured patient cells, has been hampered by the lack of suitable tools to detect the isoform-specific expression patterns. Therefore, we developed a plasmid pCMV-THRA-RFP-EGFP splicing detector that allows the visualization and quantification of the differential expression of THRA1 and THRA2 splicing isoforms in living single cells during time-lapse and perturbation experiments. This tool enables experiments to further characterize the role of THRα2 and to perform high-throughput drug screening. Molecules that modify THRA splicing may be developed into drugs for the treatment of thyroid hormone resistance syndromes. Full article
(This article belongs to the Special Issue Thyroid Hormone and Molecular Endocrinology)
Show Figures

Figure 1

Figure 1
<p><span class="html-italic">THRA</span> isoforms and mechanism of action of the <span class="html-italic">pCMV-THRA-RFP-EGFP</span> splicing detector. (<b>a</b>) Schematic representation of the 3′ ends of the <span class="html-italic">THRA1</span> splicing isoform mRNA encoding THR<span class="html-italic">α</span>1 and of the <span class="html-italic">THRA2</span> splicing isoform mRNA encoding THR<span class="html-italic">α</span>2. The orange spheres represent the T3 ligand, the gray solid rectangles represent the exons common to both isoforms, and the light blue and dark blue solid rectangles represent the isoform-specific exons. The schematic shows that T3 can bind to THR<span class="html-italic">α</span>1 but not to THR<span class="html-italic">α</span>2. Adapted from [<a href="#B5-ijms-25-13512" class="html-bibr">5</a>]: (<b>b</b>) Schematic representation of local T3-responsive gene expression based on the relative abundance of THR<span class="html-italic">α</span>1 versus THR<span class="html-italic">α</span>2. Given the same amount of local T3, cell types that express more THR<span class="html-italic">α</span>1 will have higher T3-responsive gene expression levels compared to cell types that express more THR<span class="html-italic">α</span>2. Reproduced with permission from [<a href="#B5-ijms-25-13512" class="html-bibr">5</a>]: (<b>c</b>) Schematic representation of the <span class="html-italic">THRA</span>-specific gene sequence of the <span class="html-italic">pCMV-THRA-RFP-EGFP</span> splicing detector. Endogenous splicing mechanisms will produce either an N-terminally truncated THRα1 labeled with RFP or a truncated THRα2 labeled with EGFP at their C-termini.</p>
Full article ">Figure 2
<p>Cloning steps for the generation of the <span class="html-italic">pCMV-THRA-RFP-EGFP</span> splicing detector. (<b>a</b>) Schematic representation of the <span class="html-italic">THRA</span> region between exon 7 and the 3′ UTR region of exon 10 that we chose to integrate into our <span class="html-italic">pCMV-THRA-RFP-EGFP</span> splicing detector. <span class="html-italic">THRA1</span>-specific regions are shown in light blue, while <span class="html-italic">THRA2</span>-specific regions are shown in dark blue. (<b>b</b>) Agarose gel electrophoresis showing the correct 7300 bp PCR amplification product of <span class="html-italic">THRA</span> with flanking <span class="html-italic">Bgl</span>II and <span class="html-italic">Not</span>I sites. (<b>c</b>) Agarose gel electrophoresis showing the 7300 bp <span class="html-italic">THRA</span> insert and the 4000 bp linearized expression vector. (<b>d</b>) Schematic representation of the final <span class="html-italic">pCMV-THRA-RFP-EGFP</span> splicing detector with a CMV promoter to drive gene expression.</p>
Full article ">Figure 3
<p>Validation of the <span class="html-italic">pCMV-THRA-RFP-EGFP</span> splicing detector in COS1 and N2A neuroblastoma cells. (<b>a</b>) Phase contrast and fluorescence images of COS1 and N2A neuroblastoma cells 48 h after transfection with the <span class="html-italic">pCMV-THRA-RFP-EGFP</span> splicing detector (SD). The RFP signal is a proxy for THRα1 and is shown in magenta. The EGFP signal is a proxy for THRα2 and is shown in green. Colocalization is shown in white. For each condition, experiments were performed in triplicate and 5 fields of view (POVs) were analyzed. One representative image per condition is shown. Scale bar: 100 µm. (<b>b</b>) Efficiency of <span class="html-italic">pCMV-THRA-RFP-EGFP</span> splicing detector transfection measured by EGFP<sup>+</sup> and/or RFP<sup>+</sup> cells in COS1 (black dots) and N2A neuroblastoma (blue dots) cells. Results of five POVs from three independent experiments per condition (n = 15 POVs per condition), presented as mean plus single dots. (<b>c</b>) Quantification of RFP (THRα1 proxy) and EGFP (THRα2 proxy) signals in COS1 and N2A neuroblastoma cells, shown as the percentage of mean fluorescence intensity after background subtraction. We observed a strong predominance of EGFP signal in N2A neuroblastoma cells and a slightly higher percentage of RFP compared to EGFP signal in COS1 cells. Results of five POVs from three independent experiments per condition (n ≥ 12 POVs per condition), presented as mean plus single dots. Kruskal–Wallis test and Dunn’s correction for multiple comparisons. ns = not significant; **** = <span class="html-italic">p</span> &lt; 0.001. (<b>d</b>) Percentage composition of RFP<sup>+</sup> and/or EGFP<sup>+</sup> cells. RFP<sup>+</sup>-only cells (EGFP not detected) are shown in magenta, EGFP<sup>+</sup>-only cells (RFP not detected) are shown in green, and RFP<sup>+</sup> and EGFP<sup>+</sup> cells (both colors detected at various ratios) are shown in white. The bar graph shows a strong predominance of RFP<sup>+</sup> in COS1 and a strong predominance of EGFP<sup>+</sup> in N2A neuroblastoma cells. Results of five POVs from three independent experiments per condition (n = 15 POVs per condition), shown as mean ± standard deviation. (<b>e</b>) Normalized RNA levels of endogenous <span class="html-italic">THRA1</span> (light blue) and endogenous <span class="html-italic">THRA2</span> (dark blue) in naïve COS1 cells and RFP (proxy for THRα1, in magenta) and EGFP (proxy for THRα2, in green) in COS1 cells transfected with the <span class="html-italic">pCMV-THRA-RFP-EGFP</span> splicing detector (SD). As expected, CMV-promoted RFP and EGFP RNA levels were much higher compared to endogenously promoted <span class="html-italic">THRA1</span> and <span class="html-italic">THRA2</span> and the endogenous relative abundance of <span class="html-italic">THRA1</span> and <span class="html-italic">THRA2</span> was mirrored by the relative RFP-to-EGFP abundance. n ≥ 3. Results are expressed as mean ± standard deviation. (<b>f</b>) Normalized RNA levels of endogenous <span class="html-italic">THRA1</span> (light blue) and endogenous <span class="html-italic">THRA2</span> (dark blue) in naïve N2A neuroblastoma cells and RFP (in magenta) and EGFP (in green) in N2A cells transfected with the <span class="html-italic">pCMV-THRA-RFP-EGFP</span> splicing detector (SD). CMV-promoted RFP and EGFP RNA levels were orders of magnitude higher than endogenously promoted <span class="html-italic">THRA1</span> and <span class="html-italic">THRA2</span>; however, the relative <span class="html-italic">THRA1</span>-to-<span class="html-italic">THRA2</span> abundance in naïve samples was reflected by the relative RFP-to-EGFP abundance in transfected samples. n = 3. Results are expressed as mean ± standard deviation.</p>
Full article ">Figure 4
<p>The <span class="html-italic">pCMV-THRA-RFP-EGFP</span> splicing detector can detect splicing perturbations in COS1 cells. (<b>a</b>) Phase contrast and fluorescence images of COS1 cells 48 h after transfection with the <span class="html-italic">pCMV-THRA-RFP-EGFP</span> splicing detector (SD). Cells were treated with either scrambled control ASO (scramble + SD) or <span class="html-italic">THRA2</span>-specific ASO (ASO + SD) 24 h prior to transfection. RFP signal is a proxy for THRα1 and is shown in magenta. EGFP signal is a proxy for THRα2 and is shown in green. Colocalization is shown in white. For each condition, experiments were performed in triplicate and 5 fields of view (POVs) were collected. One representative image per condition is shown. Scale bar: 100 µm. (<b>b</b>) Schematic representation of the mechanism of action of the <span class="html-italic">THRA2</span>-specific antisense oligonucleotides (ASOs). The ASO (in orange) is designed to bind to the region between <span class="html-italic">THRA</span> intron 9 and exon 10, sterically preventing <span class="html-italic">THRA2</span> splicing. (<b>c</b>) Quantification of RFP (THRα1 proxy) and EGFP (THRα2 proxy) signals, shown as a percentage of the mean fluorescence intensity after background subtraction. ASO treatment effectively blocked <span class="html-italic">THRA2</span> splicing, resulting in a loss of EGFP signal in ASO + SD samples and a statistically significant increase in RFP compared to the <span class="html-italic">pCMV-THRA-RFP-EGFP</span> splicing detector-only control. Scramble ASO showed similar levels of RFP and EGFP fluorescence compared to the <span class="html-italic">CMV-THRA-RFP-EGFP</span> splicing detector-only control. Results obtained from five POVs of three independent experiments per condition (n = 15 from five POVs randomly selected OVs per condition), presented as mean plus single data points. Kruskal–Wallis test and Dunn’s correction for multiple comparisons. ns, not significant; ** = <span class="html-italic">p</span> &lt; 0.1. (<b>d</b>) Phase contrast and fluorescence images of COS1 cells pretreated with fluorescently labeled scrambled control ASOs. Images show the CY5 signal (cyan) as a proxy for tagged scramble ASO incorporation. (<b>e</b>) Efficiency of ASO uptake measured by the percentage of CY5<sup>+</sup> cells. Results obtained from five POVs randomly selected from three independent experiments (n = 5 POVs), shown as mean plus single data points. (<b>f</b>) Normalized RNA levels of <span class="html-italic">THRA</span> exon 7, used as proxy of <span class="html-italic">pCMV-THRA-RFP-EGFP</span> splicing detector expression. No statistically significant difference was observed between the <span class="html-italic">pCMV-THRA-RFP-EGFP</span> splicing detector-only control (SD) and ASO-treated samples (scramble + SD; ASO + SD). All <span class="html-italic">pCMV-THRA-RFP-EGFP</span> splicing detector-transfected samples (SD; scramble + SD; ASO + SD) showed higher levels of exon 7 compared to naïve. n = 3 biological replicates (n = 3 C<sub>t</sub> values for each biological replicate). Ordinary one-way ANOVA and Fisher’s LSD tests. ns, not significant; * = <span class="html-italic">p</span> &lt; 0.05. (<b>g</b>) Normalized RNA levels of <span class="html-italic">THRA</span> exon 3, used as proxy for total endogenous <span class="html-italic">THRA</span> RNA expression. No statistically significant difference was observed between naïve and all transfected and treated samples (SD; scramble + SD; ASO + SD). n = 3 biological replicates (n = 3 C<sub>t</sub> values for each biological replicate). Ordinary one-way ANOVA and Fisher’s LSD tests. ns, not significant.</p>
Full article ">Figure 5
<p><span class="html-italic">pCMV-THRA-RFP-EGFP</span> splicing detector allows in vitro live time-lapse visualization and quantification of THRα1 and THRα2 isoforms in N2A neuroblastoma cells. (<b>a</b>) Phase contrast and fluorescence merged images of N2A neuroblastoma cells 24 h after transfection with the <span class="html-italic">pCMV-THRA-RFP-EGFP</span> splicing detector. The RFP signal is a proxy for THRα1 and is shown in magenta. The EGFP signal is a proxy for THRα2 and is shown in green. Co-expression is shown in white. Images were collected every hour for 16 h. Scale bar: 100 µm. (<b>b</b>) Single cell quantification of RFP (THRα1 proxy) and EGFP (THRα2 proxy) signals over time, shown as mean fluorescence intensity after background subtraction. The plots show the variability in THRα1 and THRα2 expression patterns in different single cells of the same cell line. n = 10 single cells.</p>
Full article ">Figure 6
<p><span class="html-italic">pCMV-THRA-RFP-EGFP</span> splicing detector allows in vitro live time-lapse visualization and quantification of THRα1 and THRα2 isoforms in COS1 cells. (<b>a</b>) Phase contrast and fluorescence merged images of COS1 cells 24 h after transfection with the <span class="html-italic">pCMV-THRA-RFP-EGFP</span> splicing detector. The RFP signal is a proxy for THRα1 and is shown in magenta. The EGFP signal is a proxy for THRα2 and is shown in green. Co-expression is shown in white. Images were collected every hour for 48 h. Representative images every 3 h are shown. Scale bar: 100 µm. (<b>b</b>) Single cell quantification of RFP (THRα1 proxy) and EGFP (THRα2 proxy) signals over time, shown as mean fluorescence intensity after background subtraction. The plots show the variability in THRα1 and THRα2 expression patterns in different single cells of the same cell line. n = 10 single cells.</p>
Full article ">Figure 7
<p>The <span class="html-italic">pCMV-THRA-RFP-EGFP</span> splicing detector allows in vitro live time-lapse visualization and quantification of THRα1 and THRα2 isoforms in COS1 cells during cell division. (<b>a</b>) Quantification of RFP (THRα1 proxy) and EGFP (THRα2 proxy) signals over time. The plot highlights an increase in THRα1 levels right before cell division and a subsequent drop immediately thereafter. n = 1 single cell (ROI shown in cyan). (<b>b</b>) Magnified view of <a href="#ijms-25-13512-f006" class="html-fig">Figure 6</a>a phase contrast and fluorescence merged images of single COS1 cell undergoing cell division. The RFP signal is a proxy for THRα1 and is shown in magenta. The EGFP signal is a proxy for THRα2 and is shown in green. Co-expression is shown in white. Scale bar: 50 µm.</p>
Full article ">
18 pages, 16639 KiB  
Article
Improving Object Detection for Time-Lapse Imagery Using Temporal Features in Wildlife Monitoring
by Marcus Jenkins, Kirsty A. Franklin, Malcolm A. C. Nicoll, Nik C. Cole, Kevin Ruhomaun, Vikash Tatayah and Michal Mackiewicz
Sensors 2024, 24(24), 8002; https://doi.org/10.3390/s24248002 - 14 Dec 2024
Viewed by 943
Abstract
Monitoring animal populations is crucial for assessing the health of ecosystems. Traditional methods, which require extensive fieldwork, are increasingly being supplemented by time-lapse camera-trap imagery combined with an automatic analysis of the image data. The latter usually involves some object detector aimed at [...] Read more.
Monitoring animal populations is crucial for assessing the health of ecosystems. Traditional methods, which require extensive fieldwork, are increasingly being supplemented by time-lapse camera-trap imagery combined with an automatic analysis of the image data. The latter usually involves some object detector aimed at detecting relevant targets (commonly animals) in each image, followed by some postprocessing to gather activity and population data. In this paper, we show that the performance of an object detector in a single frame of a time-lapse sequence can be improved by including spatio-temporal features from the prior frames. We propose a method that leverages temporal information by integrating two additional spatial feature channels which capture stationary and non-stationary elements of the scene and consequently improve scene understanding and reduce the number of stationary false positives. The proposed technique achieves a significant improvement of 24% in mean average precision ([email protected]:0.95) over the baseline (temporal feature-free, single frame) object detector on a large dataset of breeding tropical seabirds. We envisage our method will be widely applicable to other wildlife monitoring applications that use time-lapse imaging. Full article
(This article belongs to the Special Issue Sensors and Sensing Technologies for Object Detection and Recognition)
Show Figures

Figure 1

Figure 1
<p>An example annotated image from the RI petrel dataset.</p>
Full article ">Figure 2
<p>Comparison of the effect of colour correction on the difference mask, <math display="inline"><semantics> <msub> <mi>D</mi> <mi>M</mi> </msub> </semantics></math>. (<b>a</b>) Sample image from camera SWC3. (<b>b</b>) Corresponding <math display="inline"><semantics> <msub> <mi>T</mi> <mrow> <msub> <mi>A</mi> <mn>12</mn> </msub> <mi>R</mi> <mi>G</mi> <mi>B</mi> </mrow> </msub> </semantics></math> (before colour correction). (<b>c</b>) Corresponding <math display="inline"><semantics> <msubsup> <mi>T</mi> <mrow> <msub> <mi>A</mi> <mn>12</mn> </msub> <mi>R</mi> <mi>G</mi> <mi>B</mi> </mrow> <mo>′</mo> </msubsup> </semantics></math> (after colour correction). (<b>d</b>) <math display="inline"><semantics> <msub> <mi>D</mi> <mi>M</mi> </msub> </semantics></math> using uncorrected <math display="inline"><semantics> <msub> <mi>T</mi> <mrow> <msub> <mi>A</mi> <mn>12</mn> </msub> <mi>R</mi> <mi>G</mi> <mi>B</mi> </mrow> </msub> </semantics></math>. (<b>e</b>) <math display="inline"><semantics> <msub> <mi>D</mi> <mi>M</mi> </msub> </semantics></math> using colour-corrected <math display="inline"><semantics> <msubsup> <mi>T</mi> <mrow> <msub> <mi>A</mi> <mn>12</mn> </msub> <mi>R</mi> <mi>G</mi> <mi>B</mi> </mrow> <mo>′</mo> </msubsup> </semantics></math>.</p>
Full article ">Figure 3
<p>Modified Squeeze-and-Excitation block for input-aware <math display="inline"><semantics> <msub> <mi>T</mi> <msub> <mi>A</mi> <mn>12</mn> </msub> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>D</mi> <mi>M</mi> </msub> </semantics></math> channel weightings. Input <span class="html-italic">X</span> is the output of two convolutional layers with a kernel size of 3 × 3 and stride 1 × 1, with an intermediate ReLU layer. For <math display="inline"><semantics> <msub> <mi>F</mi> <mrow> <mi>s</mi> <mi>q</mi> </mrow> </msub> </semantics></math>, global average pooling is used across the channel dimension of <span class="html-italic">X</span>, and <math display="inline"><semantics> <msub> <mi>F</mi> <mrow> <mi>e</mi> <mi>x</mi> </mrow> </msub> </semantics></math> is a feed-forward network with a sigmoid output layer (to produce a scaling for each channel between 0 and 1). <math display="inline"><semantics> <msub> <mi>F</mi> <mrow> <mi>s</mi> <mi>c</mi> <mi>a</mi> <mi>l</mi> <mi>e</mi> </mrow> </msub> </semantics></math> denotes the multiplication between the output of <math display="inline"><semantics> <msub> <mi>F</mi> <mrow> <mi>e</mi> <mi>x</mi> </mrow> </msub> </semantics></math> and the input channels <span class="html-italic">X</span> to give <math display="inline"><semantics> <msub> <mi>X</mi> <mrow> <mi>o</mi> <mi>u</mi> <mi>t</mi> </mrow> </msub> </semantics></math>.</p>
Full article ">Figure 4
<p>Sample images from the 10 cameras that comprised our dataset.</p>
Full article ">Figure 4 Cont.
<p>Sample images from the 10 cameras that comprised our dataset.</p>
Full article ">Figure 5
<p>Box plots depicting bounding-box area distribution for each object category, where the area is normalised by the respective image’s dimensions.</p>
Full article ">Figure 6
<p>Class occurrence across each set (normalised by image count).</p>
Full article ">Figure 7
<p>Distribution of class “Adult” across each set.</p>
Full article ">Figure 8
<p>Distribution of class “Chick” across each set.</p>
Full article ">Figure 9
<p>Distribution of class “Egg” across each set.</p>
Full article ">Figure 10
<p>Visualisation of predictions during the day with a confidence threshold of 0.25.</p>
Full article ">Figure 11
<p>Visualisation of predictions during the night with a confidence threshold of 0.25.</p>
Full article ">Figure 12
<p>Visualisation of the <math display="inline"><semantics> <msub> <mi>T</mi> <msub> <mi>A</mi> <mn>12</mn> </msub> </msub> </semantics></math> (<b>b</b>) and <math display="inline"><semantics> <msub> <mi>D</mi> <mi>M</mi> </msub> </semantics></math> (<b>c</b>) channels after weighting for a given image (<b>a</b>), all with ground truth annotations.</p>
Full article ">Figure A1
<p>Illustration of the data augmentation pipeline for object detection for YOLOv7.</p>
Full article ">
21 pages, 9319 KiB  
Article
Exploring Imaging Applications of a Red-Emitting π-Acceptor (π-A) Pyrene-Benzothiazolium Dye
by Chathura S. Abeywickrama, Enya Huang, Wenhui Yan, Michael A. Vrionides, Paaramitha Warushavithana, Kristen A. Johnson, Robert V. Stahelin, Yi Pang, Tomoyasu Mani and Kaveesha J. Wijesinghe
Biosensors 2024, 14(12), 612; https://doi.org/10.3390/bios14120612 - 13 Dec 2024
Viewed by 488
Abstract
Bright biocompatible fluorescent imaging dyes with red to near-infrared (NIR) emissions are ideal candidates for fluorescence microscopy applications. Pyrene–benzothiazolium hemicyanine dyes are a new class of lysosome-specific probes reported on recently. In this work, we conduct a detailed implementation study for a pyrene–benzothiazolium [...] Read more.
Bright biocompatible fluorescent imaging dyes with red to near-infrared (NIR) emissions are ideal candidates for fluorescence microscopy applications. Pyrene–benzothiazolium hemicyanine dyes are a new class of lysosome-specific probes reported on recently. In this work, we conduct a detailed implementation study for a pyrene–benzothiazolium derivative, BTP, to explore its potential imaging applications in fluorescence microscopy. The optical properties of BTP are studied in intracellular environments through advanced fluorescence microscopy techniques, with BTP exhibiting a noticeable shift toward blue (λem ≈ 590 nm) emissions in cellular lysosomes. The averaged photon arrival time (AAT)-based studies exhibit two different emissive populations of photons, indicating the probe’s dynamic equilibrium between two distinctively different lysosomal microenvironments. Here, BTP is successfully utilized for time-lapse fluorescence microscopy imaging in real-time as a ‘wash-free’ imaging dye with no observed background interference. BTP exhibits an excellent ability to highlight microorganisms (i.e., bacteria) such as Bacillus megaterium through fluorescence microscopy. BTP is found to be a promising candidate for two-photon fluorescence microscopy imaging. The two-photon excitability of BTP in COS-7 cells is studied, with the probe exhibiting an excitation maximum at λTP ≈ 905 nm. Full article
(This article belongs to the Special Issue Advanced Fluorescence Biosensors)
Show Figures

Figure 1

Figure 1
<p>Absorbance (<b>a</b>) and emission (<b>b</b>) of BTP (1 × 10<sup>−5</sup> M) in different solvents at room temperature. For the emission spectra (<b>b</b>) acquisition, BTP was excited at λ<sub>ex</sub> ≈ 490 nm, and the emissions were collected from 510 nm to 800 nm.</p>
Full article ">Figure 2
<p>Fluorescence confocal microscopy images of MG-63 cells with BTP (1 µM) in the presence of different organelle markers. Figures (<b>a</b>–<b>l</b>) represent BTP only (<b>a</b>–<b>c</b>), with MitoView<sup>TM</sup> green (<b>d</b>–<b>f</b>), Hoechst 33342 (<b>g</b>–<b>i</b>), LysoTracker<sup>TM</sup> green DND-26 (<b>j</b>–<b>l</b>), ER-Tracker<sup>TM</sup> green (<b>m</b>–<b>o</b>), and GolgiTrack<sup>TM</sup> green (<b>p</b>–<b>r</b>). All co-staining experiments were merged with a bright field to show cell boundaries. The staining procedure and the excitation/emission parameters are described in the Methods section. The scale bar represents 10 µm.</p>
Full article ">Figure 3
<p>Fluorescence confocal microscopy images of MG-63 cells co-stained with Hoechst 33342 (<b>a</b>), MitoView<sup>TM</sup> green (<b>b</b>), and BTP (<b>c</b>). Figures (<b>d</b>–<b>f</b>) represent the bright field (<b>d</b>), a merged multichannel image (<b>e</b>), and the composite image with the bright field (<b>f</b>). The staining procedure and the excitation/emission parameters are described in the Methods section. The scale-bar represents 10 µm.</p>
Full article ">Figure 4
<p>Fluorescence confocal microscopy images of MG-63 cells stained with BTP (1 µM) and excited in the wavelength range 490–565 nm. The emissions were collected from 550 nm to 700 nm.</p>
Full article ">Figure 5
<p>Fluorescence confocal microscopy images of MG-63 cells stained with BTP (1 µM) and excited at a laser wavelength of 520 nm. The emissions were collected from 530 nm to 720 nm at 10 nm intervals.</p>
Full article ">Figure 6
<p>Fluorescence confocal microscopy analysis of MG-63 cells stained with LysoTracker<sup>TM</sup> red DND-99 (<b>b</b>,<b>c</b>) and BTP (<b>e</b>,<b>f</b>) to study their emitted photon population based on the averaged arrival time (AAT). Figures (<b>a</b>,<b>d</b>) illustrate the emitted photon distribution based on the AAT for LysoTracker<sup>TM</sup> red DND-99 (<b>a</b>) and BTP (<b>d</b>).</p>
Full article ">Figure 7
<p>Time-lapse fluorescence confocal microscopy images of MG-63 cells stained with BTP (1 µM). Cells were pre-stained with Hoechst 33342 and BTP was added at the time point t = 0 min. Cells were sequentially excited at 520 nm and the emissions were collected from 550 nm to 700 nm.</p>
Full article ">Figure 8
<p>(<b>a</b>–<b>f</b>) Fluorescence microscopy images of <span class="html-italic">Bacillus megaterium</span> cells stained with BTP (5 µM) for 30 min. Images were acquired by exciting the stained bacterial cells with standard Cy3 filter settings (580–620 nm) for the emission collection. The scale bar indicates 5 microns.</p>
Full article ">Figure 9
<p>The two-photon excitation spectra of BTP (10 µM) in ethanol at room temperature (<b>a</b>) and the emission intensity of the two-photon fluorescence as a function of excitation wavelength (<b>b</b>). The emission spectra were collected from 580 nm to 720 nm and the excitations were performed from 900 nm to 1140 nm.</p>
Full article ">Figure 10
<p>(<b>a</b>,<b>b</b>) Two-photon fluorescence microscopy images obtained for the COS-7 cells stained with BTP (1 µM) while being excited in the 895–995 nm wavelength range. The emissions were collected from 520 nm to 700 nm.</p>
Full article ">Scheme 1
<p>The architecture of donor–π–acceptor and π-acceptor type probes and the chemical structure of BTP.</p>
Full article ">Scheme 2
<p>A representative summary showing the chemical structures of different lysosomes targeting two-photon probes developed within the past decades (<b>top</b>) and a summary of pyrene-based two-photon excitable probes developed to date for cell imaging applications (<b>bottom</b>) [<a href="#B27-biosensors-14-00612" class="html-bibr">27</a>,<a href="#B36-biosensors-14-00612" class="html-bibr">36</a>,<a href="#B46-biosensors-14-00612" class="html-bibr">46</a>,<a href="#B47-biosensors-14-00612" class="html-bibr">47</a>,<a href="#B48-biosensors-14-00612" class="html-bibr">48</a>,<a href="#B49-biosensors-14-00612" class="html-bibr">49</a>,<a href="#B50-biosensors-14-00612" class="html-bibr">50</a>,<a href="#B51-biosensors-14-00612" class="html-bibr">51</a>,<a href="#B52-biosensors-14-00612" class="html-bibr">52</a>,<a href="#B53-biosensors-14-00612" class="html-bibr">53</a>].</p>
Full article ">
28 pages, 6728 KiB  
Article
Ice-Jam Flooding of the Peace–Athabasca Delta, Canada: Insights from Recent Notable Spring Breakup Events and Implications for Strategic Flow Releases from Upstream Dams
by Spyros Beltaos
Geosciences 2024, 14(12), 335; https://doi.org/10.3390/geosciences14120335 - 7 Dec 2024
Viewed by 398
Abstract
Ice jamming is the primary mechanism that can generate overland flooding and recharge the isolated basins of the Peace–Athabasca Delta (PAD), a valuable ecosystem of international importance and the ancient homeland of the Indigenous Peoples of the region. Focusing on the regulated Peace [...] Read more.
Ice jamming is the primary mechanism that can generate overland flooding and recharge the isolated basins of the Peace–Athabasca Delta (PAD), a valuable ecosystem of international importance and the ancient homeland of the Indigenous Peoples of the region. Focusing on the regulated Peace River and the Peace Sector of the delta, which has been experiencing a drying trend in between rare ice-jam floods over the last ~50 years, this study describes recent notable breakup events, associated observational data, and numerical applications to determine river discharge during the breakup events. Synthesis and interpretation of this material provide a new physical understanding that can inform the ongoing development of a protocol for strategic flow releases toward enhancing basin recharge in years when major ice jams are likely to form near the PAD. Additionally, several recommendations are made for future monitoring activities and improvements in proposed antecedent criteria for early identification of “promising” breakup events. Full article
(This article belongs to the Section Hydrogeology)
Show Figures

Figure 1

Figure 1
<p>Plan view of the lower Peace River and Peace Sector of the Peace–Athabasca Delta. Common ice jam lodgment sites (or “toes”) are shown in the upper portion of the figure. Also shown are sites of Water Survey of Canada hydrometric gauges, of which the records have been used in this study.</p>
Full article ">Figure 2
<p>Plan view of Peace River and Peace–Athabasca Delta (showing only the northern portion of the Athabasca River). The river distance from the W.A.C. Bennett dam is marked at 100 km intervals. The Slave River begins at the MOP and flows in a generally northward direction (from [<a href="#B2-geosciences-14-00335" class="html-bibr">2</a>], with changes).</p>
Full article ">Figure 3
<p>Overview of the extent of 2014 flooding discernible during aerial monitoring in Wood Buffalo National Park. Adapted from [<a href="#B26-geosciences-14-00335" class="html-bibr">26</a>] with permission from Parks Canada.</p>
Full article ">Figure 4
<p>Views of the western end of Lake Athabasca on April 20 (<b>left</b>) and 25 (<b>right</b>), 2018, showing the development of an open lead and early melt-out in the upper reach of RdR (triple channel). See image attribution at <a href="https://www.openstreetmap.org/copyright" target="_blank">https://www.openstreetmap.org/copyright</a>—accessed 12 August 2024.</p>
Full article ">Figure 5
<p>Sequence of images from 1 May 2018 mobilization and run of the ice cover at PP. Time sequence: 2024 h (stationary ice), 2027 h, 2033 h, 2040 h, 2047 h, and 2120 h. Photo times can also be seen by zooming in to the upper left corner of each image. Flow direction: right to left.</p>
Full article ">Figure 6
<p>Schematic illustration of spatiotemporal variations in ice conditions in the lower Peace River during the 2018 pre-breakup and breakup seasons, as revealed by time-lapse cameras. Conditions during darkness (~2200 h to 0400 h) are estimated. The “ice run” icon does not differentiate between sheet ice and rubble, which typically follows moving ice sheets. The partial jam in the Slave River formed over a large eddy area near the right bank, but rubble kept moving farther out and closer to the left bank.</p>
Full article ">Figure 7
<p>Water level variation in the lower Peace and upper Slave Rivers, as captured by five pressure loggers and WSC gauges. The RdR logger was placed next to the WSC gauge on Rivière des Rochers, located ~600 m upstream from the MOP. The L. Athabasca stages are from the gauge at Fort Chipewyan. The flat logger segments signify that the logger was still above water and merely indicating its own elevation.</p>
Full article ">Figure 8
<p>Variation in PP discharge in early 2018 May, as estimated by different approaches. The WSC data points represent daily mean values and are plotted at noon each day. The local ice cover moved out in late 1 May, though backwater effects likely persisted during the following days. The blue arrow marks the last day with ice-related backwater, as assessed by the WSC.</p>
Full article ">Figure 9
<p>Mean November discharge at Hudson’s Hope and below Peace Canyon Dam, 1960 to 2023. The Hudson’s Hope WSC gauge operation was discontinued in August 2019. The Canyon Dam data points were derived from BC Hydro’s Station 001 daily flows and can be downloaded from <a href="https://rivers.alberta.ca/" target="_blank">https://rivers.alberta.ca/</a>—accessed 1 December 2024. Neither station was affected by ice.</p>
Full article ">Figure 10
<p>Variation in snow on the ground and mean air temperature at the Grand Prairie met station No. 3072921. Note that the snow depletion is coincident with mild weather spells in January and February; 7.1 mm of rain was recorded on 17 January, when the minimum/maximum temperatures amounted to −20/+0.4 °C.</p>
Full article ">Figure 11
<p>The appearance of highly deteriorated ice cover at the upstream end of Moose Island shortly before final breakup: 30 April 2018 (<b>left</b>, ice moved out later that day or in early 1 May); 4 May 2020 (<b>middle</b>, ice moved out on 5 May); and 4 May 2022 (<b>right</b>, ice moved out on 5 May). The Sentinel images have been enhanced using the B04 band. A similarly mottled ice surface appears on several 5 May photos at this and other sites within the PAD reach [<a href="#B24-geosciences-14-00335" class="html-bibr">24</a>]. See satellite image attribution at <a href="https://www.openstreetmap.org/copyright" target="_blank">https://www.openstreetmap.org/copyright</a>—accessed 12 August 2024.</p>
Full article ">Figure 12
<p>Variation in water level at the PP gauge (No. 07KC001) during the passage of javes on 5 May 2022. Unpublished WSC data, provided on request.</p>
Full article ">Figure 13
<p>End-of-winter ice thickness at PP versus Fort Chipewyan degree-days of frost, 1959–2022. Based on raw WSC data and assessed according to the procedure described in [<a href="#B22-geosciences-14-00335" class="html-bibr">22</a>]. Regulation commenced in 1968, and the reservoir was full in 1971.</p>
Full article ">Figure 14
<p>Time series of Fort Chipewyan DDF and PP HF (CGVD28), 1959–2022. The regulation commenced in 1968, and the reservoir was full in 1971. The red square markers indicate LIJFs.</p>
Full article ">Figure 15
<p>Average celerity of breakup front (CB) between ~Sunny Valley and ~MOP, plotted versus freezeup level at PP (<b>a</b>) and versus FC-DDF (<b>b</b>), for all years for which relevant data are available (promising events: 1996, 1997, 2003, 2007, 2014, 2018, 2020; unpromising events: 2004, 2015, 2016, 2017, 2019, 2021). Red square markers identify LIJFs. From [<a href="#B19-geosciences-14-00335" class="html-bibr">19</a>], with changes.</p>
Full article ">Figure 16
<p>Maximum daily mean breakup discharge at PP plotted versus Fort Chipewyan degree-days of Frost (<b>a</b>) and versus Grand Prairie Oct-Apr solid precipitation (<b>b</b>) for the regulation period 1972–2022 (reservoir filling years 1968–1971 are excluded). Pearson correlation coefficient <span class="html-italic">r</span>~0.63 for both graphs.</p>
Full article ">
16 pages, 1141 KiB  
Review
Towards Clinical Application: Calcium Waves for In Vitro Qualitative Assessment of Propagated Primary Human Corneal Endothelial Cells
by Xiao Yu Ng, Gary Peh, Fernando Morales-Wong, Rami Gabriel, Poh Loong Soong, Kun-Han Lin and Jodhbir S. Mehta
Cells 2024, 13(23), 2012; https://doi.org/10.3390/cells13232012 - 5 Dec 2024
Viewed by 658
Abstract
Corneal endothelium cells (CECs) regulate corneal hydration between the leaky barrier of the corneal endothelium and the ionic pumps on the surface of CECs. As CECs do not regenerate, loss of CECs leads to poor vision and corneal blindness. Corneal transplant is the [...] Read more.
Corneal endothelium cells (CECs) regulate corneal hydration between the leaky barrier of the corneal endothelium and the ionic pumps on the surface of CECs. As CECs do not regenerate, loss of CECs leads to poor vision and corneal blindness. Corneal transplant is the only treatment option; however, there is a severe shortage of donor corneas globally. Cell therapy using propagated primary human CECs is an alternative approach to corneal transplantations, and proof of functionality is crucial for validating such CECs. Expression markers like Na-K-ATPase and ZO-1 are typical but not specific to CECs. Assessing the barrier function of the expanded CECs via electrical resistance (i.e., TEER and Ussing’s chamber) involves difficult techniques and is thus impractical for clinical application. Calcium has been demonstrated to affect the paracellular permeability of the corneal endothelium. Its absence alters morphology and disrupts apical junctions in bovine CECs, underscoring its importance. Calcium signaling patterns such as calcium waves affect the rate of wound healing in bovine CECs. Therefore, observing calcium waves in expanded CECs could provide valuable insights into their health and functional integrity. Mechanical or chemical stimulations, combined with Ca2+-sensitive fluorescent dyes and time-lapse imaging, can be used to visualize these waves, which could potentially be used to qualify expanded CECs. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Schematic showing the structure of the cornea with 3 cellular layers and 2 acellular layers of basement membrane. (<b>B</b>) A figure of a confluent monolayer of corneal endothelial cells in culture. Schematic created with BioRender.com.</p>
Full article ">Figure 2
<p>Calcium wave propagation from a single-cell mechanical stimulation captured on OptioQUANT platform to assess and demonstrate intact cell–cell interactions in human CECs. Schematic created with BioRender.com.</p>
Full article ">
19 pages, 6558 KiB  
Article
Real-Time Observation of Clickable Cyanotoxin Synthesis in Bloom-Forming Cyanobacteria Microcystis aeruginosa and Planktothrix agardhii
by Rainer Kurmayer and Rubén Morón Asensio
Toxins 2024, 16(12), 526; https://doi.org/10.3390/toxins16120526 - 5 Dec 2024
Viewed by 575
Abstract
Recently, the use of click chemistry for localization of chemically modified cyanopeptides has been introduced, i.e., taking advantage of promiscuous adenylation (A) domains in non-ribosomal peptide synthesis (NRPS), allowing for the incorporation of clickable non-natural amino acids (non-AAs) into their peptide products. In [...] Read more.
Recently, the use of click chemistry for localization of chemically modified cyanopeptides has been introduced, i.e., taking advantage of promiscuous adenylation (A) domains in non-ribosomal peptide synthesis (NRPS), allowing for the incorporation of clickable non-natural amino acids (non-AAs) into their peptide products. In this study, time-lapse experiments have been performed using pulsed feeding of three different non-AAs in order to observe the synthesis or decline of azide- or alkyne-modified microcystins (MCs) or anabaenopeptins (APs). The cyanobacteria Microcystis aeruginosa and Planktothrix agardhii were grown under maximum growth rate conditions (r = 0.35–0.6 and 0.2–0.4 (day−1), respectively) in the presence of non-AAs for 12–168 h. The decline of the azide- or alkyne-modified MC or AP was observed via pulse-feeding. In general, the increase in clickable MC/AP in peptide content reached a plateau after 24–48 h and was related to growth rate, i.e., faster-growing cells also produced more clickable MC/AP. Overall, the proportion of clickable MC/AP in the intracellular fraction correlated with the proportion observed in the dissolved fraction. Conversely, the overall linear decrease in clickable MC/AP points to a rather constant decline via dilution by growth instead of a regulated or induced release in the course of the synthesis process. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Mean (±SE) proportion of natural and clickable MC in total MC (cellular fraction, composed of four MC structural variants: DAsp-MC-YR, MC-YR, DAsp-MC-LR, MC-LR) during time-lapse experiments using pulsed feeding of non-natural amino acids (non-AAs) in order to observe the build up (<b>a</b>,<b>b</b>) or decline (<b>c</b>,<b>d</b>) of azide- or alkyne-modified MC in <span class="html-italic">M. aeruginosa</span> strain Hofbauer. Control cells were grown and processed under identical conditions but without non-AA substrate and could not show any clickable MC synthesis.</p>
Full article ">Figure 2
<p>Mean (±SE) proportion of natural and clickable AP in total AP (cellular fraction, composed of four AP structural variants: unknown AP, AP-C, AP-B, AP-A) during time-lapse experiments using pulsed feeding of non-natural amino acids (non-AAs) in order to observe the build up (<b>a</b>,<b>b</b>) or decline (<b>c</b>,<b>d</b>) of azide- or alkyne modified AP in <span class="html-italic">P. agardhii</span> strain no371/1. Control cells were grown and processed under identical conditions but without non-AA substrate and could not show any clickable MC synthesis.</p>
Full article ">Figure 3
<p>Relationship between growth rate (day<sup>−1</sup>) and (<b>a</b>–<b>c</b>) the clickable MC net production rate (day<sup>−1</sup>) in <span class="html-italic">M. aeruginosa</span> strain Hofbauer (calculated from ln(x + 1) MC-LR equivalents in ng/mL) and (<b>d</b>–<b>f</b>) the clickable AP net production rate (day<sup>−1</sup>) in <span class="html-italic">P. agardhii</span> strain no371/1 (calculated from ln(x + 1) AP-B equivalents in ng/mL) during time-lapse experiments using feeding of non-natural amino acids (non-AAs) in order to observe the build up of azide- or alkyne-modified MC/AP. Details of linear regression curves are as follows: (<b>a</b>) MC-Phe-AzR (y = −0.08 + 0.73x, R<sup>2</sup> = 0.95, <span class="html-italic">p</span> &lt; 0.0001), (<b>b</b>) MC-Prop-LysR (y = −0.47 + 1.58x, R<sup>2</sup> = 0.95, <span class="html-italic">p</span> &lt; 0.0001), (<b>c</b>) MC-Prop-TyrR (y = −0.13 + 0.86x, R<sup>2</sup> = 0.99, <span class="html-italic">p</span> &lt; 0.0001), (<b>d</b>) AP-Phe-Az (not significant, <span class="html-italic">p</span> = 0.33), (<b>e</b>) AP-Prop-Lys (y = 0.29 + 0.39x, R<sup>2</sup> = 0.21, <span class="html-italic">p</span> = 0.099), (<b>f</b>) AP-Prop-Tyr (y = −0.28 + 1.32x, R<sup>2</sup> = 0.65, <span class="html-italic">p</span> = 0.0005), where y is MC/AP production rate (day<sup>−1</sup>) and x is growth rate (day<sup>−1</sup>).</p>
Full article ">Figure 4
<p>Proportion of individual clickable MC in <span class="html-italic">M. aeruginosa</span> (<b>a</b>–<b>c</b>) or clickable AP in <span class="html-italic">P. agardhii</span> (<b>d</b>–<b>e</b>) in total MC/AP (cellular fraction) during time-lapse experiments using pulsed feeding of non-natural amino acids (non-AAs) in order to observe the decline of (<b>a</b>) MC-Phe-AzR, (<b>b</b>) MC-Prop-LysR, (<b>c</b>) MC-Prop-TyrR in <span class="html-italic">M. aeruginosa,</span> or (<b>d</b>) AP-Phe-Az, (<b>e</b>) AP-Prop-Lys, (<b>f</b>) AP-Prop-Tyr in <span class="html-italic">P. agardhii</span> strain no371/1. Using growth rates, the theoretical decline of clickable MC/AP was calculated (black symbols, straight line) and compared to the observed decline (colored symbols, dotted line). Note that the scale at the <span class="html-italic">y</span>-axis is different, as production efficiency differs between non-AAs (<a href="#toxins-16-00526-f001" class="html-fig">Figure 1</a> and <a href="#toxins-16-00526-f002" class="html-fig">Figure 2</a>).</p>
Full article ">Figure 5
<p>Workflow of time-lapse experiments using pulsed feeding of non-AAs for real-time observation of clickable MC/AP synthesis in bloom-forming cyanobacteria (the workflow was the same for both <span class="html-italic">M. aeruginosa</span> and <span class="html-italic">P. agardhii</span>): (<b>a</b>) time-lapse build up experiments; (<b>b</b>) time-lapse decline experiments. Created with BioRender.com. Note that the labeling of clickable peptides via chemo-selective reaction with fluorophore and high-resolution microscopy and flow-cytometry analysis using Alexa Fluor488 will be reported in a follow-up article.</p>
Full article ">Figure 6
<p>Chemical structures of non-AA molecules used for clickable microcystin (MC) synthesis in <span class="html-italic">M. aeruginosa</span> and for clickable anabaenopeptin (AP) synthesis in <span class="html-italic">P. agardhii</span>: (<b>a</b>) 4-Azido-L-phenylalanine (Phe-Az, MW 206.20 g/mol), (<b>b</b>) N-Propargyl-L-Lysine (Prop-Lys, MW 228.25 g/mol), (<b>c</b>) O-Propargyl-L-tyrosine (Prop-Tyr, MW 219.24 g/mol).</p>
Full article ">
20 pages, 5171 KiB  
Article
Quantification of Nearshore Sandbar Seasonal Evolution Based on Drone Pseudo-Bathymetry Time-Lapse Data
by Evangelos Alevizos
Remote Sens. 2024, 16(23), 4551; https://doi.org/10.3390/rs16234551 - 4 Dec 2024
Viewed by 996
Abstract
Nearshore sandbars are dynamic features that characterize shallow morphobathymetry and vary over a wide range of geometries and temporal lifespans. Nearshore sandbars influence beach geometry by altering the energy of incoming waves; thus, monitoring the evolution of sandbars is a fundamental approach in [...] Read more.
Nearshore sandbars are dynamic features that characterize shallow morphobathymetry and vary over a wide range of geometries and temporal lifespans. Nearshore sandbars influence beach geometry by altering the energy of incoming waves; thus, monitoring the evolution of sandbars is a fundamental approach in effective coastal planning. Due to several natural and technical limitations related to shallow seafloor mapping, there is a significant gap in the availability of high-resolution, shallow bathymetric data for monitoring the dynamic behaviour of nearshore sandbars effectively. This study introduces a novel image-processing technique that produces time series of pseudo-bathymetric data by utilizing multi-temporal (monthly) drone imagery, and it provides an assessment of local morphodynamics at a sandy beach in the southeast Mediterranean. The technique is called standardized-ratio bathymetric index (SRBI), and it transforms natural-colour drone imagery to pseudo-bathymetric data by applying an empirical formula used for satellite-derived bathymetry. This technique correlates well with laser altimetry depth measurements; however, it does not require in situ depth data for implementation. The resulting pseudo-bathymetric data allows for extracting cross-shore profiles and delineating the sandbar crest with 4 m horizontal accuracy. Stacking of temporal profiles allowed for the quantification of the sandbar’s crest and trough changes at different alongshore sections. The main findings suggest that the nearshore crescentic sandbar at Episkopi Beach (north Crete) shows strong seasonality regarding net offshore migration that is promoted by enhanced wave action during winter months. In addition, the crescentic sandbar is susceptible to morphology arrestment during prolonged weeks of low wave action. The average migration rate during winter is 10 m.month−1, with some sections exhibiting a maximum of 60 m.month−1. This study (a) offers a novel remote-sensing approach, suitable for nearshore seafloor monitoring with low computational complexity, (b) reveals sandbar geometry and temporal change in superior detail compared to other observational methods, and (c) advances knowledge about nearshore sandbar monitoring in the Mediterranean region. Full article
(This article belongs to the Section Remote Sensing in Geology, Geomorphology and Hydrology)
Show Figures

Figure 1

Figure 1
<p>Overview of the study area. Example RGB orthomosaic overlaid on Google Earth basemap. The red square shows the location of the study area on the island of Crete. The white lines mark the cross-shore profile positions, which are examined in the Results section.</p>
Full article ">Figure 2
<p>Workflow diagram followed for data processing and analysis.</p>
Full article ">Figure 3
<p>(<b>A</b>) Linear relation (<span class="html-italic">R</span><sup>2</sup> = 0.97, <span class="html-italic">p</span> &lt; 0.001) between the natural logarithms of Green and Red bands from 19 points with increasing distance from the coastline (<a href="#remotesensing-16-04551-f0A1" class="html-fig">Figure A1</a>, <a href="#app1-remotesensing-16-04551" class="html-app">Appendix A</a>); (<b>B</b>) Linear relation (<span class="html-italic">R</span><sup>2</sup> = 0.88, <span class="html-italic">p</span> &lt; 0.001) between 203 bathymetric points derived from ICESAT-2 LiDAR data (25 February 2023) and corresponding SRBI values from 17 February 2023 orthomosaics. Red dotted lines indicate the regression trend.</p>
Full article ">Figure 4
<p>Significant wave height (Hs) and direction from (<b>A</b>) November 2022 to June 2023; (<b>B</b>) July 2023 to November 2023. The red stars indicate the date of the drone surveys. Only wave directions between 0–90° and 270–360° azimuth are presented.</p>
Full article ">Figure 5
<p>Temporal pseudo-bathymetric models based on SRBI grids of the study area. The black dotted line indicates the crest of the intermediate sandbar. The white rectangles correspond to the zoomed-in areas shown in <a href="#remotesensing-16-04551-f006" class="html-fig">Figure 6</a>. Please note that the inner and outer bars are not always detected because (<a href="#remotesensing-16-04551-f006" class="html-fig">Figure 6</a>a) the outer bar is mainly in the seaward side of the area and is only partially captured in the mosaics, and (<a href="#remotesensing-16-04551-f006" class="html-fig">Figure 6</a>b) the inner bar is often welded with the shallow platform and does not show a clear morphology.</p>
Full article ">Figure 6
<p>Close-up frames of characteristic morpho-bathymetric features in natural colour and SRBI grids: (<b>a</b>,<b>b</b>) Sand-waves within the trough of a large crescentic bar, April 2023; (<b>c</b>,<b>d</b>) Rip-channel and trough of crescentic bar segment, November 2022; (<b>e</b>,<b>f</b>) Integration of intermediate and inner crescentic bar segments, May 2023. The exact positions of the frames are shown in <a href="#remotesensing-16-04551-f005" class="html-fig">Figure 5</a>.</p>
Full article ">Figure 7
<p>Temporal stacks of the cross-shore profiles (C1–C10, <a href="#remotesensing-16-04551-f001" class="html-fig">Figure 1</a>). Contours relate to SRBI values (0.3 step). Numbers 1–12 correspond to the survey month, as presented in <a href="#remotesensing-16-04551-t001" class="html-table">Table 1</a>.</p>
Full article ">Figure 8
<p>Differential distance of sandbar crest from the coastline between consecutive months.</p>
Full article ">Figure 9
<p>(<b>A</b>) Temporal bar crest positions overlaid on SRBI range mosaic (the largest absolute difference in pixel values during the 1-year monitoring period), with bright hues indicating large variability. (<b>B</b>) Points p1–p4 show the temporal variability of the SRBI at four exemplary locations.</p>
Full article ">Figure A1
<p>The orange points correspond to the locations used for extracting the band logarithm values in <a href="#remotesensing-16-04551-f003" class="html-fig">Figure 3</a>A. The green points correspond to the tracks of the ICESAT-2 LiDAR data used in <a href="#remotesensing-16-04551-f003" class="html-fig">Figure 3</a>B (dataset labels in white text). All points are overlaid on the 17 February 2023 drone RGB orthomosaic.</p>
Full article ">
17 pages, 3382 KiB  
Communication
Progressive Cachexia: Tuberculosis, Cancer, or Thyrotoxicosis? Disease-Directed Therapy and Atypical Courses of Autoimmune and Malignant Thyroid Diseases in a High Specialization Era: Case-Control Study with a Critical Literature Review
by Przemyslaw Zdziarski and Zbigniew Sroka
Biomedicines 2024, 12(12), 2722; https://doi.org/10.3390/biomedicines12122722 - 28 Nov 2024
Viewed by 858
Abstract
Background. Critical and progressive cachexia may be observed in numerous medical disciplines, but in patients with various diseases, several pathways overlap (endocrine, inflammatory and kidney diseases, heart failure, cancer). Methods. Unlike numerous cohort studies that examine thyroid cancer and risk factors, a different [...] Read more.
Background. Critical and progressive cachexia may be observed in numerous medical disciplines, but in patients with various diseases, several pathways overlap (endocrine, inflammatory and kidney diseases, heart failure, cancer). Methods. Unlike numerous cohort studies that examine thyroid cancer and risk factors, a different method was used to avoid bias and analyze the sequence of events, i.e., the pathway. A case-control analysis is presented on patients with initial immune-mediated thyroiditis complicated by cachexia, presenting pulmonary pathology coexisting with opportunistic infection, and ultimately diagnosed with cancer (TC—thyroid cancer, misdiagnosed as lung cancer). Results. Contrary to other patients with lung cancer, the presented patients were not active smokers and exclusively women who developed cachexia with existing autoimmune processes in the first phase. Furthermore, the coexistence of short overall survival without cancer progression in the most seriously ill patients, as well as correlation with sex (contrary to history of smoking) and predisposition to mycobacterial disease, are very suggestive. Although we describe three different autoimmune conditions (de Quervain’s, Graves’, and atrophic thyroiditis), disturbances in calcium and metabolic homeostasis, under the influence of hormonal and inflammatory changes, are crucial factors of cachexia and prognosis. Conclusions. The unique sequence sheds light on immune-mediated thyroid disease as a subclinical paraneoplastic process modified by various therapeutic regimens. However, it is also associated with cachexia, systemic consequences, and atypical sequelae, which require a holistic approach. The differential diagnosis of severe cachexia, adenocarcinoma with pulmonary localization, and tuberculosis reactivation requires an analysis of immunological and genetic backgrounds. Contrary to highly specialized teams (e.g., lung cancer units), immunotherapy and general medicine in aging populations require a multidisciplinary, holistic, and inquiring approach. The lack of differentiation, confusing biases, and discrepancies in the literature are the main obstacles to statistical research, limiting findings to correlations of common factors only. Time-lapse case studies such as this one may be among the first to build evidence of a pathway and an association between inflammatory and endocrine imbalances in cancer cachexia. Full article
Show Figures

Figure 1

Figure 1
<p>Initial patient selection. After initial selection, a small amount of patients was qualified, but contrary to most retrospective analyses of patients with thyroid cancer, in our clinical model, AITD preceded oncogenesis and may be with different types of AITD (i.e., de Quervain thyroiditis, Graves’ disease, Hashimoto/atrophic thyroiditis).</p>
Full article ">Figure 2
<p>Flowchart of clinical data collection and time-lapse analysis. Patients with autoimmune thyroid disease (AITD) were the starting point. The case-control study includes patient histories with well-characterized and differentiated autoimmune thyroid disease (AITD) complicated with infectious and neoplastic processes. TC was the sixth cancer in women; it was not observed in men. However, this could be apparent because the initial group consisted of patients with autoimmunity, which is more common in women with no difference between multiparous and childless. Comparing our AITDs where hyperthyroidism, hypothyroidism, or both occurred at different times, no clear effect of hypothyroidism and elevated TSH can be seen. AITD—autoimmune thyroid disease, PFS—progression-free survival, OS—overall survival, TSH—thyroid-stimulating hormone, FT3—free triiodothyronine, FT4—free thyroxine, CT—computer tomography.</p>
Full article ">Figure 3
<p>Modification of unique balance between pro- and anticancerous factors (i.e., hormonal and inflammatory signal, respectively) by microbiota (mycobacteria) and steroids. BRAF-BRAFV600E mutation; PTC—papillary thyroid cancer oncogene (RET/PTC), gks—glucocorticoids, MHC—Major Histocompatibility Complex, CTLA4—Cytotoxic T Lymphocyte Antigen-4; TG—thyreoglobulin, TSHR—thyreotropin receptor, TNF—cachectin, PFS—progression-free survival, OS—overall survival. The red symbol indicates the inhibitory effect; the green symbol indicates the stimulating effect.</p>
Full article ">
13 pages, 881 KiB  
Article
Image Processing Application for Pluripotent Stem Cell Colony Migration Quantification
by Timofey Chibyshev, Olga Krasnova, Alina Chabina, Vitaly V. Gursky, Irina Neganova and Konstantin Kozlov
Mathematics 2024, 12(22), 3584; https://doi.org/10.3390/math12223584 - 15 Nov 2024
Viewed by 494
Abstract
Human pluripotent stem cells (hPSCs) attract tremendous attention due to their unique properties. Manual extraction of trajectories of cell colonies in experimental image time series is labor intensive and subjective, thus the aim of the work was to develop a computer semi-automated protocol [...] Read more.
Human pluripotent stem cells (hPSCs) attract tremendous attention due to their unique properties. Manual extraction of trajectories of cell colonies in experimental image time series is labor intensive and subjective, thus the aim of the work was to develop a computer semi-automated protocol for colony tracking. The developed procedure consists of three major stages, namely, image registration, object detection and tracking. Registration using discrete Fourier transform and tracking based on the solution of a linear assignment problem was implemented as console programs in the Python 3 programming language using a variety of packages. Object detection was implemented as a multistep procedure in the ProStack in-house software package. The procedure consists of more than 40 elementary operations that include setting of several biologically relevant parameters, image segmentation and performing of quantitative measurements. The developed procedure was applied to the dataset containing bright-field images from time-lapse recording of the human embryonic cell line H9. The detection step took about 6 h for one image time series with a resolution of 2560 by 2160; about 1 min was required for image registration and trajectories extraction. The developed procedure was effective in detecting and analyzing the time series of images with “good” and “bad” phenotypes. The differences between phenotypes in the distance in pixels between the starting and finishing positions of trajectories, in the path length along the trajectory, and the mean instant speed and mean instant angle of the trajectories were identified as statistically significant by Mann–Whitney and Student’s tests. The measured area and perimeter of the detected colonies differed, on average, for different phenotypes throughout the entire time period under consideration. This result confirms previous findings obtained by analyzing static images. Full article
(This article belongs to the Special Issue Image Processing and Machine Learning with Applications)
Show Figures

Figure 1

Figure 1
<p>Solution overview. (<b>A</b>) Overall scheme. (<b>B</b>) Detection procedure.</p>
Full article ">Figure 2
<p>The trajectories obtained with developed procedure. (<b>A</b>,<b>B</b>) Examples of extracted trajectories for “good” and “bad” phenotypes, respectively. The identifier is printed next to the starting and ending points of the trajectory. The movement direction with increase in the frame number is visualized with a color gradient from blue to yellow. (<b>C</b>) Histogram of the direction angle for trajectories for “good” and “bad” phenotypes.</p>
Full article ">Figure 3
<p>Statistical significance of the difference in area between phenotypes. (<b>A</b>) Area, (<b>B</b>) perimeter, (<b>C</b>) Feret diameter, (<b>D</b>) minor axis, (<b>E</b>) shape factor, (<b>F</b>) average intensity. <span class="html-italic">p</span>-value annotation legend volume: ns (not significant): 5.00 <math display="inline"><semantics> <mrow> <mo>×</mo> <mspace width="3.33333pt"/> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>2</mn> </mrow> </msup> </mrow> </semantics></math> &lt; <span class="html-italic">p</span> ≤ 1.00, *: 1.00 <math display="inline"><semantics> <mrow> <mo>×</mo> <mspace width="3.33333pt"/> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>2</mn> </mrow> </msup> </mrow> </semantics></math> &lt; <span class="html-italic">p</span> ≤ 5.00 <math display="inline"><semantics> <mrow> <mo>×</mo> <mspace width="3.33333pt"/> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>2</mn> </mrow> </msup> </mrow> </semantics></math> ***: 1.00 <math display="inline"><semantics> <mrow> <mo>×</mo> <mspace width="3.33333pt"/> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>4</mn> </mrow> </msup> </mrow> </semantics></math> &lt; <span class="html-italic">p</span> ≤ 1.00 <math display="inline"><semantics> <mrow> <mo>×</mo> <mspace width="3.33333pt"/> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>3</mn> </mrow> </msup> </mrow> </semantics></math> ****: <span class="html-italic">p</span> ≤ 1.00 <math display="inline"><semantics> <mrow> <mo>×</mo> <mspace width="3.33333pt"/> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>4</mn> </mrow> </msup> </mrow> </semantics></math>.</p>
Full article ">
17 pages, 3783 KiB  
Article
Effects of the Interaction Between Time-on-Task and Task Load on Response Lapses
by Jingqiang Li, Yanru Zhou and Tianci Hao
Behav. Sci. 2024, 14(11), 1086; https://doi.org/10.3390/bs14111086 - 13 Nov 2024
Viewed by 773
Abstract
To investigate the interaction effects of prolonged working periods and different task loads on response lapses, focusing on the mechanisms of delayed responses and error lapses. Professionals such as pilots, truck drivers, and nurses often face extended work hours and fluctuating task loads. [...] Read more.
To investigate the interaction effects of prolonged working periods and different task loads on response lapses, focusing on the mechanisms of delayed responses and error lapses. Professionals such as pilots, truck drivers, and nurses often face extended work hours and fluctuating task loads. While these factors individually affect performance, their interaction and its impact on response lapses remain unclear. Twenty participants completed the Uchida–Kraepelin (U–K) Psychological Test and a dual-task version with functional near-infrared spectroscopy. Independent variables were time-on-task and task load. Dependent variables included measures of fatigue, arousal, workload, task performance (delayed and error rates), and brain functional connectivity. Both time-on-task and task load significantly affected cerebral connectivity, response lapses, workload (frustration level), fatigue, and arousal. Arousal levels significantly decreased and reaction times increased after 60 min of work. Cognitive resource regulation became challenging after 90 min under high task load levels. A decline in the connection between the prefrontal and occipital cortex during high-load tasks was observed. The findings provide insight into the mechanisms of response lapses under different task load levels and can inform strategies to mitigate these lapses during extended work periods. This study’s findings can be applied to improve work schedules and fatigue management in industries like aviation, transportation, and healthcare, helping reduce response lapses and errors during extended work periods under high task load conditions. Full article
Show Figures

Figure 1

Figure 1
<p>The psychological mechanism theory and performances of response lapses (adapted from [<a href="#B10-behavsci-14-01086" class="html-bibr">10</a>,<a href="#B11-behavsci-14-01086" class="html-bibr">11</a>]).</p>
Full article ">Figure 2
<p>Prefrontal and occipital cortical channels.</p>
Full article ">Figure 3
<p>Experimental protocol.</p>
Full article ">Figure 4
<p>Uchida–Kraepelin test interface.</p>
Full article ">Figure 5
<p>Delayed response and error rates for two-task load levels.</p>
Full article ">Figure 6
<p>Response lapse rates for two-task load levels.</p>
Full article ">Figure 7
<p>Total number of questions with two task load levels.</p>
Full article ">Figure 8
<p>Brain functional connectivity at different TOT points.</p>
Full article ">Figure 9
<p>Subjective experience before and after two task load levels.</p>
Full article ">Figure 10
<p>Subjective evaluation of workloads for two task load levels.</p>
Full article ">
12 pages, 2745 KiB  
Article
Single-Shot Time-Lapse Target-Oriented Velocity Inversion Using Machine Learning
by Katerine Rincon, Ramon C. F. Araújo, Moisés M. Galvão, Samuel Xavier-de-Souza, João M. de Araújo, Tiago Barros and Gilberto Corso
Appl. Sci. 2024, 14(21), 10047; https://doi.org/10.3390/app142110047 - 4 Nov 2024
Viewed by 596
Abstract
In this study, we used machine learning (ML) to estimate time-lapse velocity variations in a reservoir region using seismic data. To accomplish this task, we needed an adequate training set that could map seismic data to velocity perturbation. We generated a synthetic seismic [...] Read more.
In this study, we used machine learning (ML) to estimate time-lapse velocity variations in a reservoir region using seismic data. To accomplish this task, we needed an adequate training set that could map seismic data to velocity perturbation. We generated a synthetic seismic database by simulating reservoirs of varying velocities using a 2D velocity model typical of the Brazilian pre-salt ocean bottom node (OBN) acquisition, located in the Santos basin, Brazil. The largest velocity change in the injector well was around 3% of the empirical velocity model, which mimicked a realistic scenario. The acquisition geometry was formed by the geometry of 1 shot and 49 receivers. For each synthetic reservoir, the corresponding seismic data were obtained by estimating a one-shot forward-wave propagation using acoustic approximation. We studied the reservoir illumination to optimize the input data of the ML inversion. We split the set of synthetic reservoirs into two subsets: training (80%) and testing (20%) sets. We point out that the ML inversion was restricted to the reservoir zone, which means that it was inversion-oriented to a target. We obtained a good similarity between true and ML-inverted reservoir anomalies. The similarity diminished for a situation with non-repeatability noise. Full article
(This article belongs to the Section Earth Sciences)
Show Figures

Figure 1

Figure 1
<p>P-wave velocity model and acquisition geometry used to model the synthetic database. The rectangle indicates the reservoir region.</p>
Full article ">Figure 2
<p>Illumination map produced by a single seismic source. This color map should be interpreted as areas where the energy of the seismic wave pass by and is captured by the receptors: (<b>a</b>) traces of sequence numbers from 11 to 20 (inclusive), (<b>b</b>) the 5 smallest offset traces, (<b>c</b>) traces from 21 to 30, and (<b>d</b>) traces from 31 to 40. The yellow rectangle indicates the target reservoir region as in <a href="#applsci-14-10047-f001" class="html-fig">Figure 1</a>. The illumination map reveals that this simple acquisition geometry provides reasonable information to invert the target region.</p>
Full article ">Figure 3
<p>Block diagram of the proposed ML inversion methodology. The input set consists of the seismic time-lapse (TL) difference and the output consists of the inverted reservoir anomaly. The red dashed rectangles indicate the subsets of time-lapse differences considered to calculate input and target data: we extract from seismic differences the time window concentrating most of the reflection energy coming from the reservoir, and from the velocity differences we consider the spatial region of the reservoir.</p>
Full article ">Figure 4
<p>Proposed neural network architecture. The input of the ML is the seismic time-lapse difference within the reservoir time window and the output is the velocity anomaly of the target region.</p>
Full article ">Figure 5
<p>Loss curves computed on the training and validation subsets during the training of the neural network employed for velocity inversion. The loss function is calculated with velocities in the scale of the 4D anomalies (0–100 m/s range).</p>
Full article ">Figure 6
<p>Inversion results of individual samples for the perfect repeatability scenario. The yellow pattern in the panels represent the velocity anomaly, the horizontal and vertical dimensions reproduce the reservoir region indicated in the rectangle of <a href="#applsci-14-10047-f001" class="html-fig">Figure 1</a>. Comparison of true (first row) and inverted (second row) reservoir anomaly for five samples at key points of the SSIM distribution on the test subset: minimum (worst case), 25th percentile, median, 75th percentile, and maximum (best case). The last two rows compare, respectively, the central vertical and central horizontal velocity profiles of the velocity anomalies.</p>
Full article ">Figure 7
<p>Inversion results of individual samples for a scenario with non-repeatability, modeled by randomly moving the receivers in the lateral directions, with maximum perturbations equal to ±0.1 m. The yellow pattern in the panels represent the velocity anomaly; the horizontal and vertical dimensions reproduce the reservoir region indicated by the rectangle in <a href="#applsci-14-10047-f001" class="html-fig">Figure 1</a>. Comparison of true (first row) and predicted (second row) time-lapse velocity anomalies in the target region for specific samples of a test dataset contaminated with geometry non-repeatability noise. The 4D noise was modeled by randomly shifting the receivers in the lateral direction, with maximum perturbations equal to ±0.1 m. The shown samples are located at key percentiles of the SSIM distribution on the referred dataset: minimum (worst case), three quartiles, and maximum (best case). The last two rows compare, respectively, the central vertical and central horizontal velocity profiles of the velocity anomalies.</p>
Full article ">Figure 8
<p>Inversion results of individual samples for a scenario with non-repeatability, modeled by randomly moving the receivers in the lateral directions, with maximum perturbations equal to ±0.5 m. The yellow pattern in the panels represent the velocity anomaly; the horizontal and vertical dimensions reproduce the reservoir region indicated by the rectangle in <a href="#applsci-14-10047-f001" class="html-fig">Figure 1</a>. Comparison of true (first row) and predicted (second row) time-lapse velocity anomalies in the target region for specific samples of a testing dataset contaminated with geometry non-repeatability noise. The illustrated samples are located at key percentiles of the SSIM distribution on the referred dataset: minimum (worst case), three quartiles, and maximum (best case). The last two rows compare, respectively, the central vertical and central horizontal velocity profiles of the velocity anomalies.</p>
Full article ">Figure 9
<p>Spatial distribution of <math display="inline"><semantics> <mrow> <mi>Δ</mi> <mi>v</mi> </mrow> </semantics></math> prediction errors for the test scenarios with (<b>a</b>) perfect repeatability, (<b>b</b>) ±0.1 m geometry 4D noise, and (<b>c</b>) ±0.5 m geometry 4D noise.</p>
Full article ">
18 pages, 2075 KiB  
Article
Quantitative Standardized Expansion Assay: An Artificial Intelligence-Powered Morphometric Description of Blastocyst Expansion and Zona Thinning Dynamics
by Danilo Cimadomo, Samuele Trio, Tamara Canosi, Federica Innocenti, Gaia Saturno, Marilena Taggi, Daria Maria Soscia, Laura Albricci, Ben Kantor, Michael Dvorkin, Anna Svensson, Thomas Huang, Alberto Vaiarelli, Gianluca Gennarelli and Laura Rienzi
Life 2024, 14(11), 1396; https://doi.org/10.3390/life14111396 - 30 Oct 2024
Viewed by 1603
Abstract
Artificial intelligence applied to time-lapse microscopy may revolutionize embryo selection in IVF by automating data collection and standardizing the assessments. In this context, blastocyst expansion dynamics, although being associated with reproductive fitness, have been poorly studied. This retrospective study (N = 2184 blastocysts [...] Read more.
Artificial intelligence applied to time-lapse microscopy may revolutionize embryo selection in IVF by automating data collection and standardizing the assessments. In this context, blastocyst expansion dynamics, although being associated with reproductive fitness, have been poorly studied. This retrospective study (N = 2184 blastocysts from 786 cycles) exploited both technologies to picture the association between embryo and inner-cell-mass (ICM) area in µm2, the ICM/Trophectoderm ratio, and the zona pellucida thickness in µm (zp-T) at sequential blastocyst expansion stages, with (i) euploidy and (ii) live-birth per transfer (N = 548 transfers). A quantitative-standardized-expansion-assay (qSEA) was also set-up; a novel approach involving automatic annotations of all expansion metrics every 30 min across 5 h following blastulation. Multivariate regressions and ROC curve analyses were conducted. Aneuploid blastocysts were slower, expanded less and showed thicker zp. The qSEA outlined faster and more consistent zp thinning processes among euploid blastocysts, being more or as effective as the embryologists in ranking euploid embryo as top-quality of their cohorts in 69% of the cases. The qSEA also outlined faster and more consistent blastocyst expansion and zp thinning dynamics among euploid implanted versus not implanted blastocysts, disagreeing with embryologists’ priority choice in about 50% of the cases. In conclusion, qSEA is a promising objective, quantitative, and user-friendly strategy to predict embryo competence that now deserves prospective validations. Full article
(This article belongs to the Special Issue Obstetrics and Gynecology Medicine: Go From Bench to Bedside)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Definition of the features and timings under investigation. The AI-powered software CHLOE™ (Fairtility, Tel Aviv, Israel) was adopted to automatically annotate the time of blastulation (tB); the time of expanding blastocyst (tEB) (according to the definitions of the ESHRE TLT working group); and the time of biopsy (t-biopsy; i.e., the end of the video, when trophectoderm biopsy was performed) in hours post insemination (hpi). The same software annotated the area of embryo including the zp (zp-A; green circle); the area of the embryo proper (emb-A; yellow circle); the thickness of the zona pellucida (zp-T; calculated as the largest distance between the emb-A and the zp-A edges; orange line); the area of the ICM (ICM-A; purple shade); and the ratio between the area of ICM and the area of the trophectoderm (ICM/TE ratio). All of these metrics were calculated by the software at the median focal plane as the proportions of video frames occupied by each feature under investigation (single pixel = 300 µm; whole wells’ area = 90,000 µm<sup>2</sup>) at each blastulation timing. The quantitative standardized expansion assay (qSEA) was also automatically generated for each embryo by annotating the zp-A, emb-A, and the zp-T every 30 min across the 5 h following the tB. These data were then clustered according to blastocyst chromosomal constitution (euploid versus aneuploid) and reproductive competence (transferred euploid blastocysts that resulted in a LB versus transferred euploid blastocysts that did not result in a LB). This process generated six expansion maps (like the example with the blue and red lines for the two different outcomes) that were scrutinized to assess putative differences. Scale bar, 100 µm.</p>
Full article ">Figure 2
<p>(<b>A</b>) The zp-A (area of the embryo including the zona pellucida in µm<sup>2</sup>) qSEA (quantitative standardized expansion assay) outlined a larger expansion among euploid blastocysts that resulted in a live birth (LB) (green line) versus euploid blastocysts that did not result in a LB (orange line), which became significant 150 min following the time of blastulation (tB). The stars (*) identify the significant datapoints showing the mean ± SD in the two groups at that timing. (<b>B</b>) Receiver Characteristics Operating (ROC) curve analyses outlined a significant association between the zp-A qSEA with a LB after euploid blastocyst transfer unadjusted, adjusted for blastocyst morphology, and adjusted for blastocyst morphology and time of biopsy (t-biopsy). AUC, area under the curve.</p>
Full article ">Figure 3
<p>(<b>A</b>) The emb-A (area of the embryo proper in µm<sup>2</sup>) qSEA (quantitative standardized expansion assay) outlined a large expansion among euploid blastocysts that resulted in a live birth (LB) (green line) versus euploid blastocysts that did not result in a LB (orange line), which became significant already 180 min following the time of blastulation (tB). The stars (*) identify the significant datapoints showing the mean ± SD in the two groups at that timing. (<b>B</b>) Receiver Operating Characteristics (ROC) curve analyses outlined a significant association between the emb-A qSEA with a LB after euploid blastocyst transfer unadjusted, adjusted for blastocyst morphology, and adjusted for blastocyst morphology and time of biopsy (t-biopsy). AUC, area under the curve.</p>
Full article ">Figure 4
<p>(<b>A</b>) The zp-T (thickness of the zona pellucida in µm) qSEA (quantitative standardized expansion assay) outlined a more consistent zona thinning among euploid blastocysts that resulted in a live birth (LB) (green line) versus euploid blastocysts that did not result in a LB (orange line), which became significant already 180 min following the time of blastulation (tB). The stars (*) identify the significant datapoints showing the mean ± SD in the two groups at that timing. (<b>B</b>) Receiver Characteristics Operating (ROC) curve analyses outlined a significant association between the zp-T qSEA with a LB after euploid blastocyst transfer unadjusted, adjusted for blastocyst morphology, and adjusted for blastocyst morphology and time of biopsy (t-biopsy). AUC, area under the curve.</p>
Full article ">
15 pages, 3124 KiB  
Article
A Microfluidic Design for Quantitative Measurements of Shear Stress-Dependent Adhesion and Motion of Dictyostelium discoideum Cells
by Sepideh Fakhari, Clémence Belleannée, Steve J. Charrette and Jesse Greener
Biomimetics 2024, 9(11), 657; https://doi.org/10.3390/biomimetics9110657 - 27 Oct 2024
Viewed by 903
Abstract
Shear stress plays a crucial role in modulating cell adhesion and signaling. We present a microfluidic shear stress generator used to investigate the adhesion dynamics of Dictyostelium discoideum, an amoeba cell model organism with well-characterized adhesion properties. We applied shear stress and [...] Read more.
Shear stress plays a crucial role in modulating cell adhesion and signaling. We present a microfluidic shear stress generator used to investigate the adhesion dynamics of Dictyostelium discoideum, an amoeba cell model organism with well-characterized adhesion properties. We applied shear stress and tracked cell adhesion, motility, and detachment using time-lapse videomicroscopy. In the precise shear conditions generated on-chip, our results show cell migration patterns are influenced by shear stress, with cells displaying an adaptive response to shear forces as they alter their adhesion and motility behavior. Additionally, we observed that DH1-10 wild-type D. discoideum cells exhibit stronger adhesion and resistance to shear-induced detachment compared to phg2 adhesion-defective mutant cells. We also highlight the influence of cell density on detachment kinetics. Full article
(This article belongs to the Special Issue Biological Attachment Systems and Biomimetics)
Show Figures

Figure 1

Figure 1
<p>Microfluidic device design. (<b>A</b>) CAD model of the microfluidic array, featuring 2 inlets, 1 outlet, and a cell adhesion chamber, designed to ensure controlled and uniform shear stress, provide ample space for cell mobility, and enable short-term and long-term experiments. (<b>B</b>) Microfluidic chip fabricated based on the CAD design in (<b>A</b>). (<b>C</b>) Three-dimensional numerical model of the device with a semi-spherical amoeba model fixed in the middle of the cell adhesion chamber. The shape of the cell model is shown zoomed in for clarity. (<b>D</b>) Velocity magnitude along the device at a flow rate of 5 mL h<sup>−1</sup> at inlet 1, showing a uniform distribution of velocity across the device. (<b>E</b>) Velocity magnitude along a line through the cell adhesion chamber width (Y-direction), highlighting the uniformity of shear stress distribution; the green box shows the width of the field of view in (<b>A</b>). (<b>F</b>) Difference between the analytically calculated wall shear stress and the numerically calculated average shear stress applied to the amoeba cell surface.</p>
Full article ">Figure 2
<p>Analysis of <span class="html-italic">D. discoideum</span> cell motility under shear stress. (<b>A</b>) Detachment of a single DH1-10 cell at 10× magnification, at 10 mL h<sup>−1</sup>; the arrow in the last frame indicates the flow direction. (<b>B</b>) Migration of a single DH1-10 and single <span class="html-italic">phg2</span> cell under fluid flow over a period of 40 min. (<b>C</b>) Cell migration of 10 cells under shear stress. (<b>D</b>,<b>E</b>) Mean directionality of cell movement as a function of applied shear stress, indicating to what extent migration is aligned with flow direction. Directionality is defined as the angle between the flow direction and the cell movement direction over a period of 40 min. Therefore, cos(θ) = 1 indicates fully biased cell movement in the flow direction, cos(θ) = 0 indicates cell movement perpendicular to fluid flow, and cos(θ) = −1 indicates cell movement opposite to fluid flow direction. Error bars represent the standard error. Significance levels in the figures are represented as follows: <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 (***), and not significant (ns) for <span class="html-italic">p</span> &gt; 0.05.</p>
Full article ">Figure 3
<p>Analysis of cell detachment under various shear stress conditions. (<b>A</b>) Primary data showing the microscope raw image at time, t = 0 for DH1-10 cells. (<b>B</b>) Same t = 0 data after image treatment, including background subtraction and conversion to a binary image. (<b>C</b>) Treated image of shear stress chambe r after t = 40 min. Flow for images (<b>A</b>–<b>C</b>) was 10 mL h<sup>−1</sup>. (<b>D</b>–<b>F</b>) Comparative analysis of cell detachments for all flow rates and cell types. (D) Cell detachment curves for <span class="html-italic">Dictyostelium discoideum</span> DH1-10 in medium at three flow rates of Q = 10 mL h<sup>−1</sup> (blue), Q = 5 mL h<sup>−1</sup> (green), and Q = 2 mL h<sup>−1</sup> (red), illustrating increased detachment with higher shear stresses from increasing fluid flow rates. (<b>E</b>) Cell detachment curves for the <span class="html-italic">phg2</span> adhesion-defective mutant in medium at three flow rates of Q = 10 mL h<sup>−1</sup> (blue), Q = 5 mL h<sup>−1</sup> (green), and Q = 2 mL h<sup>−1</sup> (red), showing a rapid increase in detachment levels. (<b>F</b>) Differential response of DH1-10 and <span class="html-italic">phg2</span> final detachment percentages (after 40 min) for flow rates of 2, 5, and 10 mL h<sup>−1</sup>. Figures denote significance as <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 (****), and <span class="html-italic">p</span> &gt; 0.05 (ns).</p>
Full article ">Figure 4
<p>Simulation of shear stresses on an amoeba and accounting for the influence of upstream cells. (<b>A</b>) Schematic of the simulation showing a test cell (dark blue), from which the shear stresses are obtained, and up to 7 upstream cells (light blue) that are separated by distance d. (<b>B</b>) Shear stress (τ) as a function of the total distance (d<sub>tot</sub>) to the most distant amoeba with data points for the number of cells equal to 0, 2, 4, and 7.</p>
Full article ">Figure 5
<p>Detachment percentages of (<b>A</b>) DH1-10 wild-type cells and (<b>B</b>) <span class="html-italic">phg2</span> mutant cells at different initial cell densities under flow rates of 2 mL h<sup>−1</sup>, 5 mL h<sup>−1</sup>, and 10 mL h<sup>−1</sup>.</p>
Full article ">
20 pages, 11125 KiB  
Article
QSOX1 Modulates Glioblastoma Cell Proliferation and Migration In Vitro and Invasion In Vivo
by Reetika Dutt, Colin Thorpe and Deni S. Galileo
Cancers 2024, 16(21), 3620; https://doi.org/10.3390/cancers16213620 - 26 Oct 2024
Viewed by 1138
Abstract
Background: Quiescin Sulfhydryl Oxidase 1 (QSOX1) is an enzyme that catalyzes the oxidation of free thiols to generate disulfide bonds in a variety of proteins, including the cell surface and extracellular matrix. QSOX1 has been reported to be upregulated in a number [...] Read more.
Background: Quiescin Sulfhydryl Oxidase 1 (QSOX1) is an enzyme that catalyzes the oxidation of free thiols to generate disulfide bonds in a variety of proteins, including the cell surface and extracellular matrix. QSOX1 has been reported to be upregulated in a number of cancers, and the overexpression of QSOX1 has been correlated with aggressive cancers and poor patient prognosis. Glioblastoma (GBM) brain cancer has been practically impossible to treat effectively, with cells that rapidly invade normal brain tissue and escape surgery and other treatment. Thus, there is a crucial need to understand the multiple mechanisms that facilitate GBM cell invasion and to determine if QSOX1 is involved. Methods and Results: Here, we investigated the function of QSOX1 in human glioblastoma cells using two cell lines derived from T98G cells, whose proliferation, motility, and invasiveness has been shown by us to be dependent on disulfide bond-containing adhesion and receptor proteins, such as L1CAM and the FGFR. We lentivirally introduced shRNA to attenuate the QSOX1 protein expression in one cell line, and a Western blot analysis confirmed the decreased QSOX1 expression. A DNA content/cell cycle analysis using flow cytometry revealed 27% fewer knockdown cells in the S-phase of the cell cycle, indicating a reduced proliferation. A cell motility analysis utilizing our highly quantitative SuperScratch time-lapse microscopy assay revealed that knockdown cells migrated more slowly, with a 45% decrease in migration velocity. Motility was partly rescued by the co-culture of knockdown cells with control cells, indicating a paracrine effect. Surprisingly, knockdown cells exhibited increased motility when assayed using a Transwell migration assay. Our novel chick embryo orthotopic xenograft model was used to assess the in vivo invasiveness of knockdown vs. control cells, and tumors developed from both cell types. However, fewer invasive knockdown cells were observed after about a week. Conclusions: Our results indicate that an experimental reduction in QSOX1 expression in GBM cells leads to decreased cell proliferation, altered in vitro migration, and decreased in vivo invasion. Full article
(This article belongs to the Special Issue Invasion in Glioblastoma)
Show Figures

Figure 1

Figure 1
<p>Cell appearance and Western blot analysis. (<b>A</b>) Phase-contrast images of uninfected T98G, control T98G/pLKO.1, and QSOX1 knockdown T98G/sh86 cells. Bar, 100 μm. (<b>B</b>) Western blot analysis of T98G/pLKO.1 and T98G/sh86 cell lysates. Human anti-QSOX1 antibody was used for probing QSOX1 expression and anti-GAPDH antibody was used as loading control. (<b>C</b>) Quantitation of Western blot showing relative expression levels of QSOX1a and QSOX1b in T98G/pLKO.1 and T98G/sh86 cells. Expression levels are normalized to GAPDH levels. The original Western blot figure can be found in <a href="#app1-cancers-16-03620" class="html-app">Supplementary Figure S1</a>.</p>
Full article ">Figure 2
<p>Cell cycle/DNA content analysis of GBM cells. (<b>A</b>) Histograms generated by ModFit LT software depicting various stages of the cell cycle in T98G/pLKO.1 and T98G/sh86 cells. S-phase is depicted by the striped region between the red G1 and red G2 peaks and is an indicator of cell proliferation. (<b>B</b>) Average percentage of cells in the S-phase for T98G/pLKO.1 and T98G/sh86 cells. In total, 50,000 cells per cell type were analyzed per experiment. Graph depicts data from 3 separate experiments. <span class="html-italic">p</span>-value &lt; 0.05.</p>
Full article ">Figure 3
<p><span class="html-italic">SuperScratch</span> Assay images for cell motility. Phase-contrast images of T98G/pLKO.1 (<b>left</b>) and T98G/sh86 (<b>right</b>) cells at the start (0 h) and at the end (20 h). Bottom cell tracks row shows paths taken by tracked cells in red superimposed on 0 h images. Bar, 100 μm.</p>
Full article ">Figure 4
<p><span class="html-italic">SuperScratch</span> Assay measurement of cell velocity. Graph shows the overall average velocity of cells (microns/minute) over the 20 h period. In total, 100 total cells per cell type were analyzed from 3 separate experiments (10 cells/field of view). <span class="html-italic">p</span>-value &lt; 0.001.</p>
Full article ">Figure 5
<p><span class="html-italic">SuperScratch</span> Assay for paracrine effect. Fluorescence images of co-culture experiments of 25% T98G/sh86/DiI + 75% T98G/pLKO.1 (<b>left</b>) and 25% T98G/sh86/DiI + 75% T98G/sh86 (<b>right</b>) at the start (0 h) and at the end (20 h) of time-lapse image collection. T98G/sh86/DiI cells appear as white. Track points row shows paths traveled by tracked cells over the course of the experiment in red superimposed on 0 h images. Bar, 100 μm.</p>
Full article ">Figure 6
<p><span class="html-italic">SuperScratch</span> Assay measurement of paracrine effect. Graph showing the overall average velocity of cells (microns/minute) over the 20 h period. In total, 60–90 total cells per cell type were analyzed from 3 separate experiments (10 cells/field of view). <span class="html-italic">p</span>-value &lt; 0.001.</p>
Full article ">Figure 7
<p>Transwell cell migration. Light micrographs of migrated T98G pLKO.1 and T98G sh86 stained with crystal violet and visualized on the underside of Transwell inserts at end of experiment. Bars: high magnification, 100 μm; low magnification, 500 μm.</p>
Full article ">Figure 8
<p>Quantitation of Transwell cell migration. Graph showing the number of migrated cells per field of view on the underside of the Transwell inserts. Cells were counted from 24 fields of view per cell type (2 separate experiments × 3 replicate Transwell filters per experiment × 4 fields of view per Transwell filter). <span class="html-italic">p</span>-value &lt; 0.05.</p>
Full article ">Figure 9
<p>Embryonic chick brain microinjection. Images acquired through dissecting microscope during microinjection process. (<b>A</b>) Embryonic day E5 chick embryo inside shell viewed through a hole cut at the top of the egg. (<b>B</b>) E5 embryo being held by its amnion membrane showing visible optic tectum (OT). (<b>C</b>) Embryo held by amnion membrane with OT being injected with GBM cells mixed with dye. (<b>D</b>) E5 embryo immediately after injection of OT with cells showing extent of ventricle.</p>
Full article ">Figure 10
<p>Tissue preparation for visualization. (<b>A</b>) E10 whole brain post injection. (<b>B</b>) Brain embedded in agar. (<b>C</b>) Embedded brain being sectioned on a vibratome. (<b>D</b>) Sectioned whole brain slice for mounting and visualization. OT, optic tectum.</p>
Full article ">Figure 11
<p>E10 chick brain sections with T98G/pLKO.1 and T98G/sh86 tumors. Invading cells labeled with Vybrant DiO are indicated by white arrows. Images are maximum intensity projections from confocal z-stacks. OT, optic tectum; T, tumor; P, pial surface; V, ventricular surface. Bar, 500 μm.</p>
Full article ">Figure 12
<p>E13 chick brain sections with T98G/pLKO.1 and T98G/sh86 tumors. Invading cells labeled with Vybrant DiO are indicated by white arrows. Images are maximum intensity projections from confocal z-stacks. OT, optic tectum; T, tumor; P, pial surface; V, ventricular surface. Bar, 500 μm.</p>
Full article ">
Back to TopTop