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

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

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (219)

Search Parameters:
Keywords = notch formations

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
16 pages, 1466 KiB  
Review
Reviewing the Genetic and Molecular Foundations of Congenital Spinal Deformities: Implications for Classification and Diagnosis
by Diana Samarkhanova, Maxat Zhabagin and Nurbek Nadirov
J. Clin. Med. 2025, 14(4), 1113; https://doi.org/10.3390/jcm14041113 - 9 Feb 2025
Viewed by 580
Abstract
Congenital spinal deformities (CSDs) are rare but severe conditions caused by abnormalities in vertebral development during embryogenesis. These deformities, including scoliosis, kyphosis, and lordosis, significantly impair patients’ quality of life and present challenges in diagnosis and treatment. This review integrates genetic, molecular, and [...] Read more.
Congenital spinal deformities (CSDs) are rare but severe conditions caused by abnormalities in vertebral development during embryogenesis. These deformities, including scoliosis, kyphosis, and lordosis, significantly impair patients’ quality of life and present challenges in diagnosis and treatment. This review integrates genetic, molecular, and developmental insights to provide a comprehensive framework for classifying and understanding CSDs. Traditional classification systems based on morphological criteria, such as failures in vertebral formation, segmentation, or mixed defects, are evaluated alongside newer molecular-genetic approaches. Advances in genetic technologies, including whole-exome sequencing, have identified critical genes and pathways involved in somitogenesis and sclerotome differentiation, such as TBX6, DLL3, and PAX1, as well as key signaling pathways like Wnt, Notch, Hedgehog, BMP, and TGF-β. These pathways regulate vertebral development, and their disruption leads to skeletal abnormalities. The review highlights the potential of molecular classifications based on genetic mutations and developmental stage-specific defects to enhance diagnostic precision and therapeutic strategies. Early diagnosis using non-invasive prenatal testing (NIPT) and emerging tools like CRISPR-Cas9 gene editing offer promising but ethically complex avenues for intervention. Limitations in current classifications and the need for further research into epigenetic and environmental factors are discussed. This study underscores the importance of integrating molecular genetics into clinical practice to improve outcomes for patients with CSDs. Full article
Show Figures

Figure 1

Figure 1
<p>Classification of the spinal deformities based on the nature of vertebral development failure.</p>
Full article ">Figure 2
<p>SHH and BMP signaling in sclerotome development. SHH protein initiates sclerotome formation and its transformation into cartilage after secretion from the notochord and neural tube floor plate. During early sclerotome development, SHH is antagonized by BMP signals from the LPM. Sclerotome development is supported by the secretion of Noggin from the notochord, which inhibits BMP and promotes cartilage development in conjunction with SHH. PAX1 and PAX9 activate the <span class="html-italic">BAPX1</span> gene, which is necessary for axial skeleton development. DM—dermomyotome, IM—intermediate mesoderm, N—notochord, NT—neural tube, LPM—lateral plate mesoderm.</p>
Full article ">Figure 3
<p>Crosstalk between Wnt, Notch, TGFβ, BMP, and Hedgehog signaling pathways. β-catenin activates the Notch signaling pathway. GSK3β phosphorylates NICD and activates Notch signaling, while Notch degrades β-catenin and negatively regulates its stability. As for the Hedgehog pathway, Wnt inhibits it by Gli3. The expression of ligands and components, such as Wnts, LRP5, Axin, BMP2, and TGF-β, is determined by TGF-β/BMP signaling and Wnt signaling, and these two signaling pathways are connected by the interaction between Smad7 and Axin.</p>
Full article ">
17 pages, 1629 KiB  
Review
Atherosclerosis: A Comprehensive Review of Molecular Factors and Mechanisms
by Vasiliki Tasouli-Drakou, Ian Ogurek, Taha Shaikh, Marc Ringor, Michael V. DiCaro and KaChon Lei
Int. J. Mol. Sci. 2025, 26(3), 1364; https://doi.org/10.3390/ijms26031364 - 6 Feb 2025
Viewed by 1399
Abstract
Atherosclerosis, a condition characterized by the accumulation of lipids and a culprit behind cardiovascular events, has long been studied. However, in recent years, there has been an increase in interest in its initiation, with researchers shifting focus from traditional pathways involving the vascular [...] Read more.
Atherosclerosis, a condition characterized by the accumulation of lipids and a culprit behind cardiovascular events, has long been studied. However, in recent years, there has been an increase in interest in its initiation, with researchers shifting focus from traditional pathways involving the vascular infiltration of oxidized lipids and towards the novel presence of chronic inflammatory pathways. The accumulation of pro-inflammatory cytokines, in combination with the activation of transcription factors, creates a positive feedback loop that drives the creation and progression of atherosclerosis. From the upregulation of the nod-like receptor protein 3 (NLRP3) inflammasome and the Notch and Wnt pathways to the increased expression of VEGF-A and the downregulation of connexins Cx32, Cx37, and Cx40, these processes contribute further to endothelial dysfunction and plaque formation. Herein, we aim to provide insight into the molecular pathways and mechanisms implicated in the initiation and progression of atherosclerotic plaques, and to review the risk factors associated with their development. Full article
(This article belongs to the Section Molecular Pathology, Diagnostics, and Therapeutics)
Show Figures

Figure 1

Figure 1
<p>Figure showing the creation of the NLRP3 inflammasome and pro-inflammatory cytokines IL-1β and IL-18. The attachment of oxLDL on the CD36–TLR4–TLR6 signaling complex activates NF-κB, which translocates to the nucleus and binds to NF-κB-dependent promoters for the activation of target genes. The produced NLRP3 binds with an apoptosis-associated speck-like protein (ASC) and procaspase-1 to create an NLRP3 inflammasome. The NLRP3 inflammasome cleaves pro-caspase-1 into its active form caspase-1, which in turn cleaves pro-IL-1β and pro-IL-18 to their active isomers IL-1β and IL-18, respectively. It is important to note that the created IL-1β can bind to IL-1R on the extracellular surface of macrophages to further induce the formation of more IL-1β and IL-18, thus creating a pro-inflammatory environment [<a href="#B27-ijms-26-01364" class="html-bibr">27</a>].</p>
Full article ">Figure 2
<p>Figure summarizing the pathophysiology of atherosclerosis. From the infiltration of LDLs to the creation of a calcified fibrous cap and its rupture, multiple factors contribute to the initiation and progression of atherosclerosis.</p>
Full article ">Figure 3
<p>Intracellular portion of the Notch pathway. Once the intracellular domain of the Notch receptor (NICD) has been cleaved by the enzyme γ-secretase, it translocates to the nucleus to bind to transcription factors such as Mastermind-like proteins (MAMLs) and CSL for the activation of target transcription genes [<a href="#B62-ijms-26-01364" class="html-bibr">62</a>].</p>
Full article ">Figure 4
<p>Summary of the inflammatory feedback mechanism in the initiation of atherosclerosis. Atherosclerosis is not just a collection of individual events but a positive feedback loop in which each step upregulates the previous one. The infiltration of LDLs into the tunica intima and the creation of lipid-rich macrophages in conjunction with inflammatory cytokines stimulate the progression of atherosclerosis.</p>
Full article ">
10 pages, 2009 KiB  
Article
Morphometry of the Scapular Notch and Its Clinical Implication in Suprascapular Nerve Entrapment
by Jhonatan Duque-Colorado, Oscar Andrés Alzate-Mejia and Mariano del Sol
Diagnostics 2025, 15(3), 346; https://doi.org/10.3390/diagnostics15030346 - 2 Feb 2025
Viewed by 592
Abstract
Background/Objectives: The aim of the present study was to evaluate the relationship between the type of scapular notch (SN), the morphometry of the SN, and the area of the suprascapular nerve (SSN). In addition to determining whether scapular notches other than Type [...] Read more.
Background/Objectives: The aim of the present study was to evaluate the relationship between the type of scapular notch (SN), the morphometry of the SN, and the area of the suprascapular nerve (SSN). In addition to determining whether scapular notches other than Type VI, according to the classification of Rengachary, can generate a predisposition to SSN entrapment neuropathy. Methods: One hundred and sixty-nine dry scapulae were examined, the scapular notches were classified, according to the classification of Rengachary, and for each SN, the superior transverse diameter (STD), longitudinal diameter (LD), and area of the SN were determined. The SSN was dissected in five shoulders and its area was calculated. The data were analyzed in the statistical software SPSS. Results: The values for the STD, LD, and area of the SN showed significant differences between the types of scapular notches (p < 0.0001). Along the same lines, a considerable positive correlation (r = 0.79; p < 0.0001) was established between the area of the SN and the STD. Similarly, a very strong positive correlation (r = 0.87; p < 0.0001) was established between the area of the SN and the LD. This indicated that, as the STD and the LD increase, the area of the SN increases. Conclusions: Although different studies have reported an association between SN Type VI and the compression of the SSN by the formation of a bony hole that reduces the area of the notch, we have found that SN Type IV presented a smaller area among the types of notches and a smaller area than the SSN, which exposes the SSN to be closer to or in contact with the superior transverse ligament of the scapula, potentially subjecting the nerve to greater pressure and potentially resulting in SSN entrapment. This is evidence that should be considered in the clinical diagnosis of patients with entrapment neuropathy, since the type of SN and the area of the SSN can be determined by ultrasound, which contributes to a more accurate preoperative evaluation and diagnosis. Full article
(This article belongs to the Special Issue Advances in Human Anatomy)
Show Figures

Figure 1

Figure 1
<p>Types of scapular notches as classified by Rengachary et al. [<a href="#B4-diagnostics-15-00346" class="html-bibr">4</a>]. (<b>A</b>) Type I. (<b>B</b>) Type II. (<b>C</b>) Type III. (<b>D</b>) Type IV. (<b>E</b>) Type V. (<b>F</b>) Type VI.</p>
Full article ">Figure 2
<p>Morphometric measurements. (<b>A</b>) Superior transverse diameter represented by the red line and longitudinal diameter represented by the blue line. (<b>B</b>) Area of the scapular notch, Represented by the red-highlighted section.</p>
Full article ">Figure 3
<p>Graphical representation of the Kruskal–Wallis one-way ANOVA. The same letter in different scapular notch (SN) types means no significant differences. (<b>A</b>) SN type and superior transverse diameter (STD); (<b>B</b>) SN type and longitudinal diameter (LD); (<b>C</b>) SN type and area of the SN.</p>
Full article ">Figure 4
<p>Correlation between measurements. (<b>A</b>) Area of the SN and superior transverse diameter (STD); (<b>B</b>) area of the SN and longitudinal diameter (LD).</p>
Full article ">Figure 5
<p>Detection of area by virtual gray channel at transverse section of the suprascapular nerve.</p>
Full article ">
19 pages, 4936 KiB  
Article
Mid-Term Outcomes of a Rectangular Stem Design with Metadiaphyseal Fixation and a 135° Neck–Shaft Angle in Reverse Total Shoulder Arthroplasty
by Yacine Ameziane, Laurent Audigé, Christian Schoch, Matthias Flury, Hans-Kaspar Schwyzer, Alessandra Scaini, Emanuele Maggini and Philipp Moroder
J. Clin. Med. 2025, 14(2), 546; https://doi.org/10.3390/jcm14020546 - 16 Jan 2025
Viewed by 601
Abstract
Background/Objectives: Classical reverse shoulder arthroplasty (RSA) with a high neck–shaft angle (NSA) of 155° has shown satisfactory outcomes. However, newer RSA designs aim to improve results by modifying the stem design. This study evaluates the 5-year outcomes of a stem design featuring [...] Read more.
Background/Objectives: Classical reverse shoulder arthroplasty (RSA) with a high neck–shaft angle (NSA) of 155° has shown satisfactory outcomes. However, newer RSA designs aim to improve results by modifying the stem design. This study evaluates the 5-year outcomes of a stem design featuring a rectangular metadiaphyseal fixation and a 135° NSA. Methods: This prospective bicentric case series included and longitudinally followed up patients that were treated for cuff arthropathy, massive irreparable rotator cuff tears, or eccentric osteoarthritis using a non-cemented rectangular metadiaphyseal fixation stem with a 135° NSA (Univers Revers, Arthrex, Naples, FL, USA). Subjective and objective functional outcome scores (Constant–Murley Score (CS), Shoulder Pain and Disability Index (SPADI), and Subjective Shoulder Value (SSV)), range of motion (ROM), radiographic outcome, adverse events, complications, and quality of life were investigated. Results: This study enrolled 132 patients (59% female, mean age 75 years, SD 6). At the 5-year follow-up, subjective and objective outcomes significantly improved compared to baseline: CS (32.9 to 71.7, p < 0.001), SPADI (38.7 to 86.2, p < 0.001), and SSV (43.0 to 84.1, p < 0.001). ROM improved in flexion (80° to 142.4°, p < 0.001), abduction (71.5° to 130.2°, p < 0.001), internal rotation (p < 0.001), internal rotation at 90° abduction (12.7° to 45.0°, p < 0.001), and abduction strength (0.8 kg to 5.2 kg, p < 0.001). External rotation remained unchanged (32.1° to 32.0°, p = 0.125), but external rotation at 90° abduction improved (20.9° to 52.7°, p < 0.001). No signs of implant migration, subsidence, shift, tilt, alignment loss, or wear were observed, but scapular bone spur formation (11%), scapular notching grade 1 (10%), bone resorption (10%), and partial humeral radiolucent lines (1%) were reported. Conclusions: Rectangular stems with metadiaphyseal fixation and a 135° neck–shaft angle in RSA consistently improve shoulder function, showing no aseptic loosening and minimal radiological changes at 5 years. Full article
Show Figures

Figure 1

Figure 1
<p>Example at the five-year follow-up after reverse shoulder arthroplasty with 135° NSA and a cementless metadiaphyseal fixation stem (Univers Revers, Arthrex, Naples, FL).</p>
Full article ">Figure 2
<p>Flow chart—enrolled patients and follow-up.</p>
Full article ">Figure 3
<p>Signs of diaphyseal bone resorption at the 5-year follow-up after RTSA implantation.</p>
Full article ">Figure 4
<p>Reverse shoulder arthroplasty at the 5-year follow-up showing grade 1 scapular notching and consecutive scapular bone spur formation.</p>
Full article ">Figure 5
<p>Kaplan–Meier curve: implant survival rate, including number at risk at specific time points. Numbers in parentheses indicate revisions conducted prior to the 2-year follow-up and at the 2- and 5-year follow-ups.</p>
Full article ">Figure 6
<p>Longitudinal changes of subjective and objective outcome parameters and ROM from baseline to the five-year follow-up. (<b>a</b>) Longitudinal changes of subjective quality of life parameters from baseline to five-year follow-up. (<b>b</b>) Longitudinal changes of the SPADI, the Constant-Murley Score and the Subjective Shoulder-Value from baseline to five-year follow-up. (<b>c</b>) Longitudinal changes of ROM from baseline to five-year follow-up. (<b>d</b>) Longitudinal changes of ROM in terms of maximal internal rotation heights from baseline to five-year follow-up.</p>
Full article ">Figure 6 Cont.
<p>Longitudinal changes of subjective and objective outcome parameters and ROM from baseline to the five-year follow-up. (<b>a</b>) Longitudinal changes of subjective quality of life parameters from baseline to five-year follow-up. (<b>b</b>) Longitudinal changes of the SPADI, the Constant-Murley Score and the Subjective Shoulder-Value from baseline to five-year follow-up. (<b>c</b>) Longitudinal changes of ROM from baseline to five-year follow-up. (<b>d</b>) Longitudinal changes of ROM in terms of maximal internal rotation heights from baseline to five-year follow-up.</p>
Full article ">
43 pages, 2331 KiB  
Review
Molecular Mechanisms of Dietary Compounds in Cancer Stem Cells from Solid Tumors: Insights into Colorectal, Breast, and Prostate Cancer
by Alexandru Filippi, Teodora Deculescu-Ioniță, Ariana Hudiță, Oana Baldasici, Bianca Gălățeanu and Maria-Magdalena Mocanu
Int. J. Mol. Sci. 2025, 26(2), 631; https://doi.org/10.3390/ijms26020631 - 13 Jan 2025
Viewed by 1024
Abstract
Cancer stem cells (CSC) are known to be the main source of tumor relapse, metastasis, or multidrug resistance and the mechanisms to counteract or eradicate them and their activity remain elusive. There are different hypotheses that claim that the origin of CSC might [...] Read more.
Cancer stem cells (CSC) are known to be the main source of tumor relapse, metastasis, or multidrug resistance and the mechanisms to counteract or eradicate them and their activity remain elusive. There are different hypotheses that claim that the origin of CSC might be in regular stem cells (SC) and, due to accumulation of mutations, these normal cells become malignant, or the source of CSC might be in any malignant cell that, under certain environmental circumstances, acquires all the qualities to become CSC. Multiple studies indicate that lifestyle and diet might represent a source of wellbeing that can prevent and ameliorate the malignant phenotype of CSC. In this review, after a brief introduction to SC and CSC, we analyze the effects of phenolic and non-phenolic dietary compounds and we highlight the molecular mechanisms that are shown to link diets to CSC activation in colon, breast, and prostate cancer. We focus the analysis on specific markers such as sphere formation, CD surface markers, epithelial–mesenchymal transition (EMT), Oct4, Nanog, Sox2, and aldehyde dehydrogenase 1 (ALDH1) and on the major signaling pathways such as PI3K/Akt/mTOR, NF-κB, Notch, Hedgehog, and Wnt/β-catenin in CSC. In conclusion, a better understanding of how bioactive compounds in our diets influence the dynamics of CSC can raise valuable awareness towards reducing cancer risk. Full article
(This article belongs to the Special Issue Molecular Mechanisms of Dietary Compounds in Cancer Management)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic representation of major mechanisms of action in cases of phenolic and non-phenolic dietary compounds on stemness pathways, such as Notch, Wnt/β-catenin, and Sonic Hedgehog pathways; insights into breast, prostate, and colorectal cancer. In active CSC, (i) after heterodimerization of Notch with JAG/DLL and proteolysis with ADAM and γ-secretase, a soluble fragment NICS is liberated in cytoplasm and translocated into nucleus; (ii) Wnt binds to Frizzled and its co-receptor LRP5/6 stimulating the complex formation (axin, Dsh, GSK-3β, APC) that will lead to β-catenin translocation into nucleus; (iii) Shh binds to PTCH which activate SMO and Gli1/2 is translocated into nucleus [<a href="#B290-ijms-26-00631" class="html-bibr">290</a>,<a href="#B293-ijms-26-00631" class="html-bibr">293</a>]. The phenolic and non-phenolic compounds can hinder the activity of these pathways and, eventually, will block gene transcription and further the processes associated with cell stemness, such as self-renewal and proliferation. Legend: JAG, Jagged ligand; DLL, Delta-like ligand; ADAM, a disintegrin and metalloproteinase; NICD, Notch intracellular domain; CBL, Casitas B-lineage Lymphoma transcription factor; Dsh, dishevelled; APC, adenomatous polyposis coli; GSK-3β, glycogen synthase kinase-3; LRP5/6, low-density lipoprotein receptor-related protein-5/-6; TEF, thyrotrophic embryonic factor; TCF, T-cell factor; PTCH, patched; Shh, sonic hedgehog; SMO, smoothened; SuFu, suppressor of fused; Gli1/2, glioma-associated oncoprotein family. Created with BioRender.com (accessed on 23 December 2024).</p>
Full article ">Figure 2
<p>Schematic representation of major mechanisms of action in case of phenolic and non-phenolic dietary compounds on stemness markers such as NF-κB level, PI3K/Akt/mTOR, and STAT3 pathways; insights into breast, prostate, and colorectal cancer [<a href="#B290-ijms-26-00631" class="html-bibr">290</a>,<a href="#B294-ijms-26-00631" class="html-bibr">294</a>,<a href="#B298-ijms-26-00631" class="html-bibr">298</a>,<a href="#B300-ijms-26-00631" class="html-bibr">300</a>]. Legend: HER2, human epidermal growth factor receptor 2; Ras, protein similar to the one coded by Rat sarcoma virus; TNFα, tumor necrosis factor α, TNFR1, tumor necrosis factor receptor; TRAF, TNF Receptor-Associated Factor 2; NIK, NF-κB-inducing kinase; IκB, inhibitor of nuclear factor kappa B; NF-κB, nuclear factor kappa-light-chain-enhancer of activated B cells; RelA, member of NF-κB transcription factors (p65); RTK, receptor tyrosine kinase; PI3K, phosphoinositide 3-kinase; Akt, protein kinase B; mTORC1, mammalian target of rapamycin complex 1; HIF-1α, hypoxia-inducible factor 1α; S6K1, S6 kinase 1; 4EBP-1, eukaryotic translation initiation factor 4E-binding protein 1; SREBP, sterol regulatory element-binding protein; IL, interleukin; ILR, interleukin receptor; JAK, Janus kinase; Src, protein similar to the one coded by Rous sarcoma virus, non-receptor tyrosine kinase; STAT3, signal transducer and activator of transcription 3. Created with BioRender.com (accessed on 23 December 2024).</p>
Full article ">Figure 3
<p>Phenolic and non-phenolic action on CSC that manifest mesenchymal phenotype in breast, prostate, and colorectal cancer [<a href="#B302-ijms-26-00631" class="html-bibr">302</a>,<a href="#B303-ijms-26-00631" class="html-bibr">303</a>]. Created with BioRender.com (accessed on 23 December 2024).</p>
Full article ">
19 pages, 1902 KiB  
Article
A Study on the Fracture of Brittle Heterogeneous Materials Using Non-Extensive Statistical Mechanics and the Energy Distribution Function
by Dimos Triantis, Ilias Stavrakas, Ermioni D. Pasiou and Stavros K. Kourkoulis
Materials 2025, 18(2), 335; https://doi.org/10.3390/ma18020335 - 13 Jan 2025
Viewed by 421
Abstract
The fracture process of heterogeneous materials is studied here in the framework of the discipline of Non-Extensive Statistical Mechanics. Acoustic emission data provided by an experimental protocol with concrete specimens, plain or fiber-reinforced, under bending are taken advantage of. This innovation of the [...] Read more.
The fracture process of heterogeneous materials is studied here in the framework of the discipline of Non-Extensive Statistical Mechanics. Acoustic emission data provided by an experimental protocol with concrete specimens, plain or fiber-reinforced, under bending are taken advantage of. This innovation of the study lies in the fact that the analysis of the acoustic activity is implemented in terms of the energy content of the acoustic signals rather than of their interevent time or their interevent distance. The Energy Distribution Functions were properly fitted using the expression proposed by Shcherbakov, Kuksenko and Chmelet. This study reveals that the loading and fracture processes of the specific materials are definitely non-additive and non-extensive. It is concluded that the presence of notches is crucial since it assigns non-additivity and non-extensivity from relatively low loading levels due to the early formation of the fracture process zone around the crown of the notch. The values of the Tsallis entropic index, q, that were determined are in very good agreement with the respective ones obtained in previous studies by means of different analysis tools. Finally, a clear correlation between the index q and the average energy content of the acoustic signals is highlighted for the whole range of values of the energy content of the acoustic signals. Full article
Show Figures

Figure 1

Figure 1
<p>The temporal evolution of the load applied for typical specimens of the four classes of beams tested. The lower embedded sketch schematically describes the geometry of the specimens and the position of the acoustic sensors (empty circles indicate the sensors positioned at the rear vertical face of the beams), which are numbered from 1 to 8. The upper embedded photo shows a P<span class="html-italic">l</span>C-class specimen during testing.</p>
Full article ">Figure 2
<p>The energy content of the acoustic hits recorded during four typical experiments (one from each class) in juxtaposition to the temporal evolution of the applied load vs. the time to failure (t<sub>f</sub>-t) parameter, adopting logarithmic scales for both axes: (<b>a</b>) P<span class="html-italic">l</span>C class; (<b>b</b>) M<span class="html-italic">e</span>F class; (<b>c</b>) P<span class="html-italic">p</span>F class; (<b>d</b>) P<span class="html-italic">o</span>F class.</p>
Full article ">Figure 3
<p>The Cumulative Distribution Function for the energy content of the acoustic hits. Experimental data are denoted by empty markers while the curves fitted to the data by means of Equation (3) are denoted by continuous lines. (<b>a</b>) P<span class="html-italic">l</span>C class; (<b>b</b>) M<span class="html-italic">e</span>F class; (<b>c</b>) P<span class="html-italic">p</span>F class; (<b>d</b>) P<span class="html-italic">o</span>F class.</p>
Full article ">Figure 4
<p>The Cumulative Distribution Function for the energy content of the acoustic hits recorded during the four typical experiments considered. Empty circular markers and dotted lines represent the acoustic hits recorded during the interval t &lt; t* (Group I), while filled markers and continuous lines represent the acoustic hits recorded during the interval t* &lt; t &lt; t<sub>f</sub> (Group II). (<b>a</b>) P<span class="html-italic">l</span>C class; (<b>b</b>) M<span class="html-italic">e</span>F class; (<b>c</b>) P<span class="html-italic">p</span>F class; (<b>d</b>) P<span class="html-italic">o</span>F class.</p>
Full article ">Figure 5
<p>The dependence of the entropic index q on the average energy content of the acoustic hits, for the overall duration of the four tests analyzed and, also, for the stages before and after the maximization of the applied load.</p>
Full article ">Figure 6
<p>(<b>a</b>) The Cumulative Distribution Function for the energy content of the acoustic hits recorded during typical experiments of the present protocol; (<b>b</b>) qualitative representation of the data for the Cumulative Distribution Function for the energy, based on <a href="#materials-18-00335-f002" class="html-fig">Figure 2</a> of the milestone paper by Shcherbakov et al. [<a href="#B43-materials-18-00335" class="html-bibr">43</a>].</p>
Full article ">Figure 7
<p>The evolution of the entropic index q in terms of the applied load (normalized over its maximum value) in the case of intact (un-notched) plain concrete beams loaded under quasi-static three-point bending [<a href="#B66-materials-18-00335" class="html-bibr">66</a>].</p>
Full article ">
33 pages, 8068 KiB  
Article
Silencing of Epidermal Growth Factor-like Domain 8 Promotes Proliferation and Cancer Aggressiveness in Human Ovarian Cancer Cells by Activating ERK/MAPK Signaling Cascades
by Yong-Jung Song, Ji-Eun Kim, Lata Rajbongshi, Ye-Seon Lim, Ye-Jin Ok, Seon-Yeong Hwang, Hye-Yun Park, Jin-Eui Lee, Sae-Ock Oh, Byoung-Soo Kim, Dongjun Lee, Hwi-Gon Kim and Sik Yoon
Int. J. Mol. Sci. 2025, 26(1), 274; https://doi.org/10.3390/ijms26010274 - 31 Dec 2024
Viewed by 794
Abstract
Ovarian cancer (OC) is the second most common female reproductive cancer and the most lethal gynecological malignancy worldwide. Most human OCs are characterized by high rates of drug resistance and metastasis, leading to poor prognosis. Improving the outcomes of patients with relapsed and [...] Read more.
Ovarian cancer (OC) is the second most common female reproductive cancer and the most lethal gynecological malignancy worldwide. Most human OCs are characterized by high rates of drug resistance and metastasis, leading to poor prognosis. Improving the outcomes of patients with relapsed and treatment-resistant OC remains a challenge. This study aimed to investigate the role of epidermal growth factor-like domain 8 (EGFL8) in human OC by examining the effects of siRNA-mediated EGFL8 knockdown on cancer progression. EGFL8 knockdown in human OC cells promoted aggressive traits associated with cancer progression, including enhanced proliferation, colony formation, migration, invasion, chemoresistance, and reduced apoptosis. Additionally, knockdown upregulated the expression of epithelial–mesenchymal transition (EMT) markers (Snail, Twist1, Zeb1, Zeb2, and vimentin) and cancer stem cell biomarkers (Oct4, Sox2, Nanog, KLF4, and ALDH1A1), and increased the expression of matrix metallopeptidases (MMP-2 and MMP-9), drug resistance genes (MDR1 and MRP1), and Notch1. Low EGFL8 expression also correlated with poor prognosis in human OC. Overall, this study provides crucial evidence that EGFL8 inhibits the proliferation and cancer aggressiveness of human OC cells by suppressing ERK/MAPK signaling. Therefore, EGFL8 may serve as a valuable prognostic biomarker and a potential target for developing novel human OC therapies. Full article
(This article belongs to the Section Molecular Oncology)
Show Figures

Figure 1

Figure 1
<p>EGFL8 expression and its relationship with prognosis of patients with ovarian cancer (OC). (<b>A</b>) Relative expression levels of EGFL8 mRNA in human OC tissues (<span class="html-italic">n</span> = 426, red box) and normal ovarian tissues (<span class="html-italic">n</span> = 88, blue box) analyzed using GEPIA2 platform based on TCGA and GTEx data. (<b>B</b>) Representative images show EGFL8 protein expression in ovarian stromal cells (arrows) and vascular endothelial cells (arrowheads) of human OC and normal ovarian tissues, obtained from Human Protein Atlas database, as determined by immunohistochemical staining (magnification, 40×). Bar graphs represent area of immunostained regions in each group. (<b>C</b>) Overall survival (OS) and disease-specific survival (DSS) analysis of EGFL8 expression levels in patients with OC. Number of patients at risk (No. at risk) at different time points is displayed on graph. Significant differences between normal and OC groups are indicated by * <span class="html-italic">p</span> &lt; 0.05 and *** <span class="html-italic">p</span> &lt; 0.001. TCGA, The Cancer Genome Atlas; UALCAN, University of Alabama at Birmingham Cancer.</p>
Full article ">Figure 2
<p>EGFL8 protein expression in human OC cells as detected by Western blot analysis. (<b>A</b>) EGFL8 protein expression in various human OC cell lines. (<b>B</b>) Western blot analysis for EGFL8 siRNA transfection efficiency, comparing non-transfected control (Cont), scrambled negative control siRNA (siNC), and EGFL8 siRNA-transfected A2780 and SKOV3 cells. Data represent mean ± standard deviation (SD) of three independent experiments. * <span class="html-italic">p</span> &lt; 0.05 and *** <span class="html-italic">p</span> &lt; 0.001 vs. respective control.</p>
Full article ">Figure 3
<p>Effects of EGFL8 knockdown on cell proliferation. (<b>A</b>) Bar graphs showing proliferation of non-transfected control (control), scrambled negative control siEGFL8 (siNC), and siEGFL8-transfected A2780 and SKOV3 cells on days 1, 2, and 3, as assessed by WST-1 assay. Representative phase-contrast microscopic images depict cell morphology in siNC and siEGFL8-transfected A2780 and SKOV3 cells on days 1, 2, and 3. Scale bars = 150 μm. (<b>B</b>) Representative fluorescence micrographs of Ki-67 staining (blue: DAPI; green: Ki-67) in siNC and siEGFL8-transfected A2780 and SKOV3 cells at 24 and 48 h. Relative percentage of Ki-67<sup>+</sup> cells was calculated by dividing number of cells stained with Ki-67 by total number of cells stained with DAPI. Scale bars = 150 μm. All data are expressed as relative values compared to scrambled control group. Data represent mean ± SD of three independent experiments. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001 vs. respective scrambled control group.</p>
Full article ">Figure 4
<p>Effects of EGFL8 knockdown on colony formation and spheroid growth of human OC cells. (<b>A</b>) Colony formation assay showing siEGFL8 transfection stimulated colony formation in scrambled negative control siEGFL8 (siNC), and siEGFL8-transfected A2780 and SKOV3 cells on day 7 (magnification, 40×). (<b>B</b>) Representative phase-contrast microscopic images of OC cell spheroids generated by 3D cell culture in siNC and siEGFL8-transfected A2780 cells on days 1, 4, 6, 10, 12 and 14, and SKOV3 cells on days 3, 5, and 7. Scale bars = 150 μm. All data are expressed as relative values compared to scrambled control groups. Data represent mean ± SD of three independent experiments. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.05, and *** <span class="html-italic">p</span> &lt; 0.001 vs. respective scrambled control group.</p>
Full article ">Figure 5
<p>Effects of EGFL8 silencing on the migration and invasion of A2780 and SKOV3 cells. (<b>A</b>) Representative phase-contrast microscopy images from wound healing assay showing transfected and non-transfected A2780 and SKOV3 cells migrating into cell-free space. EGFL8 siRNA-transfected cells have significantly increased migratory ability compared with non-transfected control (Cont) and scrambled negative control siRNA (siNC) groups. Scale bars = 650 μm. (<b>B</b>) Representative images of transwell invasion assay with A2780 cells. Data represent mean ± SD of three independent experiments. Scale bars = 75 μm. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001 vs. respective non-transfected control group. <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 vs. respective scrambled control group.</p>
Full article ">Figure 6
<p>Effects of EGFL8 knockdown on apoptosis in human OC cells. Expression of apoptotic proteins in siEGFL8-transfected and non-transfected A2780 (<b>A</b>) and SKOV3 (<b>B</b>) cells was detected by Western blot analysis. Bar graphs depict densitometry quantitation of protein expression normalized to β-actin protein. Data represent mean ± SD of three independent experiments. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001 vs. respective scrambled control group.</p>
Full article ">Figure 7
<p>Effect of EGFL8 on expression of EMT-related markers and MMPs. (<b>A</b>) mRNA levels of key EMT markers in siEGFL8-transfected and non-transfected A2780 and SKOV3 cells as measured by qRT-PCR. (<b>B</b>) Confocal microscopy analysis of vimentin expressed in A2780 cells transfected with EGFL8 siRNA or scrambled siRNAs (blue: DAPI; red: vimentin). Scale bars = 30 μm. Bar graph shows quantification of red mean fluorescence intensity (MFI). Relative integrated mean fluorescence intensity (iMFI) of siEGFL8-transfected A2780 cells was normalized to that of scrambled negative control siEGFL8 (siNC). (<b>C</b>) Analysis of <span class="html-italic">MMP-2</span> and <span class="html-italic">MMP-9</span> expression in siEGFL8-transfected and non-transfected A2780 and SKOV3 cells by qRT-PCR. Bar graphs depict densitometric quantitation of mRNA expression normalized to <span class="html-italic">GAPDH</span>. Data represent mean ± standard deviation (SD) of three independent experiments. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001 vs. respective scrambled control group. EMT, epithelial–mesenchymal transition; MMP, matrix metallopeptidase.</p>
Full article ">Figure 8
<p>Effects of EGFL8 knockdown on expression of cancer stem cell (CSC) markers in human ovarian cancer (OC) cells. (<b>A</b>) Expression levels of <span class="html-italic">CD44</span>, <span class="html-italic">CD117</span>, <span class="html-italic">CD133</span>, <span class="html-italic">Sox2</span>, <span class="html-italic">Oct4</span>, <span class="html-italic">Nanog</span>, and <span class="html-italic">KLF4</span> in siEGFL8-transfected and non-transfected cells were analyzed by qRT-PCR. (<b>B</b>) Flow cytometry analysis of ALDH1A1 expression in siEGFL8-transfected and non-transfected A2780 cells. Data represent mean ± SD of three independent experiments. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001 vs. respective scrambled control group.</p>
Full article ">Figure 9
<p>Effects of EGFL8 knockdown on chemosensitivity of cisplatin, doxorubicin, docetaxel, paclitaxel, curcumin, and rucaparib in A2780 (<b>A</b>) and SKOV3 (<b>B</b>) cells. Cell viability in scrambled control and siEGFL8-transfected human OC cells 24 h after drug treatment, as shown by WST-1 assay. Data represent mean ± SD of three independent experiments. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001, siEGFL8#1 vs. respective scrambled control group. <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05, <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01, and <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001, siEGFL8#2 vs. respective scrambled control group.</p>
Full article ">Figure 10
<p>Effects of EGFL8 knockdown on expression levels of chemoresistance-related genes <span class="html-italic">MDR1</span>, <span class="html-italic">MRP1</span>, and <span class="html-italic">Notch1</span> in scrambled control and EGFL8 siEGFL8-transfected cells as measured by qRT-PCR. Bar graphs show densitometry quantification of mRNA expression normalized to <span class="html-italic">GAPDH</span>. Data represent mean ± SD of three independent experiments. ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001 vs. respective scrambled control group.</p>
Full article ">Figure 11
<p>Effects of EGFL8 knockdown on ERK/MAPK pathway in human OC cells. Western blot analysis of ERK1/2 and phospho-ERK1/2 (pERK1/2) protein expression in scrambled control and siEGFL8-transfected A2780 and SKOV3 cells. Bar graphs display densitometric quantification of protein expression normalized to β-actin. Data presented as mean ± SD of three independent experiments. ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001 compared to respective scrambled control group. <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 compared to U0126-untreated siEGFL8-transfected A2780 and SKOV3 cells.</p>
Full article ">Figure 12
<p>EGFL8 knockdown promotes human ovarian cancer (OC) cell proliferation by activating ERK/MAPK signaling pathway. (<b>A</b>) Results of WST-1 assay showing proliferation rates and representative phase-contrast microscopic images depicting cell morphology in scramble- and siEGFL8-transfected A2780 and SKOV3 cells at 24 h post-transfection. Scale bars = 150 μm. (<b>B</b>) Representative fluorescence micrographs of Ki-67 staining (blue: DAPI; green: Ki-67) in scramble- and siEGFL8-transfected A2780 and SKOV3 cells at 24 h. Scale bars = 150 μm. All data expressed as relative values compared to scrambled control group. Data represent mean ± SD of three independent experiments. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001 vs. respective scrambled control group. <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 compared to U0126-untreated siEGFL8-transfected A2780 and SKOV3 cells.</p>
Full article ">Figure 13
<p>EGFL8 knockdown-mediated upregulation of EMT-regulating transcription factors, metastasis-promoting molecules, and stemness-regulating molecules in human OC cells acts via ERK/MAPK signaling pathway. Expression levels of key genes associated with malignancy in scramble- and siEGFL8-transfected A2780 (<b>A</b>) and SKOV3 (<b>B</b>) cells 24 h post-transfection measured using qRT-PCR. Bar graphs display relative mRNA expression levels of EMT-regulating transcription factors (<span class="html-italic">Twist1</span>, <span class="html-italic">Zeb1</span>, and <span class="html-italic">Zeb2</span>), (<b>B</b>) metastasis-promoting molecules (<span class="html-italic">MMP-2</span>, <span class="html-italic">MMP-9</span>, and <span class="html-italic">Notch1</span>), and stemness-regulating molecules (<span class="html-italic">Oct4</span>, <span class="html-italic">CD44</span>, and <span class="html-italic">CD117</span>) in A2780 and SKOV3 cells. <span class="html-italic">GAPDH</span> used as housekeeping gene for normalization. Data presented as relative values compared to respective scrambled control group and represent mean ± SD of three independent experiments. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001 vs. respective scrambled control group. <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05, <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01, and <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 vs. U0126-untreated siEGFL8-transfected A2780 and SKOV3 cells.</p>
Full article ">Figure 13 Cont.
<p>EGFL8 knockdown-mediated upregulation of EMT-regulating transcription factors, metastasis-promoting molecules, and stemness-regulating molecules in human OC cells acts via ERK/MAPK signaling pathway. Expression levels of key genes associated with malignancy in scramble- and siEGFL8-transfected A2780 (<b>A</b>) and SKOV3 (<b>B</b>) cells 24 h post-transfection measured using qRT-PCR. Bar graphs display relative mRNA expression levels of EMT-regulating transcription factors (<span class="html-italic">Twist1</span>, <span class="html-italic">Zeb1</span>, and <span class="html-italic">Zeb2</span>), (<b>B</b>) metastasis-promoting molecules (<span class="html-italic">MMP-2</span>, <span class="html-italic">MMP-9</span>, and <span class="html-italic">Notch1</span>), and stemness-regulating molecules (<span class="html-italic">Oct4</span>, <span class="html-italic">CD44</span>, and <span class="html-italic">CD117</span>) in A2780 and SKOV3 cells. <span class="html-italic">GAPDH</span> used as housekeeping gene for normalization. Data presented as relative values compared to respective scrambled control group and represent mean ± SD of three independent experiments. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001 vs. respective scrambled control group. <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05, <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01, and <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 vs. U0126-untreated siEGFL8-transfected A2780 and SKOV3 cells.</p>
Full article ">
19 pages, 18132 KiB  
Article
Notch Fatigue Damage Evolution Mechanism of TC21 Alloy with Multilevel Lamellar Microstructures
by Xiaosong Zhou, Xiang Li, Chaowen Huang, Quan Wu and Fei Zhao
Metals 2025, 15(1), 18; https://doi.org/10.3390/met15010018 - 29 Dec 2024
Viewed by 445
Abstract
This study aims to explore the effect of microstructural parameters on the notch fatigue damage behavior of the TC21 alloy. Different levels of lamellar microstructures were achieved through distinct aging temperatures of 550 °C, 600 °C, and 650 °C. The findings reveal that [...] Read more.
This study aims to explore the effect of microstructural parameters on the notch fatigue damage behavior of the TC21 alloy. Different levels of lamellar microstructures were achieved through distinct aging temperatures of 550 °C, 600 °C, and 650 °C. The findings reveal that increasing aging temperature primarily contributes to the augmentation of α colony (αc) thickness, grain boundaries α phase (GBα) thickness, and α fine (αfine) size alongside a reduction in α lath (αlath) thickness and αfine content. The notch alters stress distribution and relaxation effects at the root, enhancing notched tensile strength while weakening plasticity. Moreover, the increased thickness of GBα emerges as a critical factor leading to the increase area of intergranular cleavage fracture. It is noteworthy that more thickness αlath and smaller αfine facilitate deformation coordination and enhance dislocation accumulation at the interface, leading to a higher propensity for micro-voids and micro-cracks to propagate along the interface. Conversely, at elevated aging temperatures, thinner αlath and larger αfine are more susceptible to fracture, resulting in the liberation of dislocations at the interface. The reduction in αlath thickness is crucial for triggering the initiation of multi-system dislocations at the interface, which promotes the development of persistent slip bands (PSBs) and dislocation nets within αlath. This phenomenon induces inhomogeneous plastic deformation and localized hardening, fostering the formation of micro-voids and micro-cracks. Full article
(This article belongs to the Special Issue Structure and Mechanical Properties of Titanium Alloys)
Show Figures

Figure 1

Figure 1
<p>Original microstructure of TC21 alloy (<b>a</b>) SEM image and (<b>b</b>) TEM image.</p>
Full article ">Figure 2
<p>(<b>a</b>) Geometry of notch tensile and high cycle fatigue specimen, where A and B are represented as two reference symbols (Reprinted from Ref. [<a href="#B14-metals-15-00018" class="html-bibr">14</a>]) (<b>b</b>) 3D model images; (<b>c</b>) real sample.</p>
Full article ">Figure 3
<p>Schematic illustration for (<b>a</b>) SEM observation of micro-voids, micro-cracks and deformation characteristics and (<b>b</b>) TEM observation of microstructural features below the primary fatigue crack initiation site.</p>
Full article ">Figure 4
<p>Microstructure of TC21 alloy after 1010 °C/0.5 h solid solution, 800 °C/2 h/AC annealing and different temperatures: (<b>a</b>,<b>d</b>,<b>g</b>) 550 °C, (<b>b</b>,<b>e</b>,<b>h</b>) 600 °C, (<b>c</b>,<b>f</b>,<b>i</b>) 650 °C aging for 5 h/AC; (<b>a</b>–<b>f</b>) SEM images; (<b>g</b>–<b>i</b>) TEM images.</p>
Full article ">Figure 5
<p>Notch tensile properties under different aging temperatures of stress–strain curves.</p>
Full article ">Figure 6
<p>Notch tensile fracture surfaces of TC21 alloys at aging temperatures of 550 °C (<b>a</b>–<b>c</b>), 600 °C (<b>d</b>–<b>f</b>) and 650 °C (<b>g</b>–<b>i</b>): (<b>a</b>,<b>d</b>,<b>g</b>) low magnification of fracture, (<b>b</b>,<b>e</b>,<b>h</b>) tearing ridges and secondary crack; (<b>c</b>,<b>f</b>,<b>d</b>) dimples.</p>
Full article ">Figure 7
<p>The micro-voids and micro-cracks characteristics underneath the main crack initiation site of specimens A1 (<b>a</b>–<b>c</b>) and A2 (<b>d</b>–<b>f</b>): (<b>a</b>) micro-voids and micro-cracks initiated at the torsional deformed α colony; (<b>b</b>) micro-voids and micro-cracks pileups at α<sub>lath</sub>/β<sub>trans</sub> interface and within the β<sub>trans</sub>; (<b>c</b>) micro-voids nucleated at α<sub>fine</sub> plates; (<b>d</b>) micro-voids and micro-cracks initiated at the α colony at notch root; (<b>e</b>) micro-voids and micro-cracks nucleated at α<sub>lath</sub>/β<sub>trans</sub> interface and along the interfaces; (<b>f</b>) micro-cracks nucleated at the α<sub>lath</sub> and within the torsional deformed α<sub>fine</sub>.</p>
Full article ">Figure 8
<p>The micro-voids and micro-cracks characteristics underneath the main crack initiation site of specimens B1 (<b>a</b>–<b>c</b>) and B2 (<b>d</b>–<b>f</b>): (<b>a</b>) micro-voids and micro-cracks initiated at the torsional deformed α colony; (<b>b</b>) micro-voids and secondary cracks formed at α<sub>lath</sub>/β<sub>trans</sub> interface and GBα; (<b>c</b>) micro-cracks nucleated at the GBα; (<b>d</b>) micro-voids and micro-cracks initiated at the α colony at notch root; (<b>e</b>) micro-voids and micro-cracks nucleated at α<sub>lath</sub>/β<sub>trans</sub> interface and α<sub>fine</sub>/β<sub>r</sub> interface; (<b>f</b>) micro-cracks nucleated at the α<sub>lath</sub> and within the torsional deformed β<sub>r</sub>.</p>
Full article ">Figure 9
<p>The micro-voids and micro-cracks characteristics underneath the main crack initiation site of specimens C1 (<b>a</b>–<b>c</b>) and C2 (<b>d</b>–<b>f</b>): (<b>a</b>) micro-voids and micro-cracks initiated at the torsional deformed α colonies; (<b>b</b>) torsional deformed α colony; (<b>c</b>) micro-voids nucleated at α laths; (<b>d</b>) micro-voids and micro-cracks initiated at the deformed α colony at notch root; (<b>e</b>) micro-voids and micro-cracks nucleated at α<sub>lath</sub>/β<sub>trans</sub> interface and α<sub>fine</sub>/β<sub>r</sub> interface; (<b>f</b>) micro-cracks and slip lines nucleated within the α<sub>fine</sub> and β<sub>r</sub>.</p>
Full article ">Figure 10
<p>The TEM images of fatigue crack initiation region of the interrupted specimen A1(σ = 240 MPa, N = 4 × 10<sup>6</sup> cycles) and A2 (σ = 240 MPa, N = 8 × 10<sup>6</sup> cycles): (<b>A1-a–A1-c</b>) high-density dislocations formed in the α<sub>fine</sub>/β<sub>trans</sub> interface; (<b>A1-d–A1-h</b>) the high-resolution transmission electron diffraction (SAED), fast Fourier transform (FFT) patterns and atoms spacing of (d); (<b>A1-i–A1-k</b>) geometric phase analysis (GPA) of (<b>A1</b>-<b>d</b>); (<b>A2</b>-<b>a,A2</b>-<b>b</b>) high-density dislocations formed in the α<sub>fine</sub>/β<sub>trans</sub> interface; (<b>A2</b>-<b>c</b>) ladder-like structures formed in the α<sub>lath</sub>/β<sub>trans</sub> interface; (<b>A2-d–A2-f</b>) twins formed and SAED in the α<sub>fine</sub>; (<b>A2-g–A2-i</b>) the GPA of (<b>A2</b>-<b>d</b>).</p>
Full article ">Figure 11
<p>The TEM images of fatigue crack initiation region of the interrupted specimen B1 (σ = 280 MPa, N = 4 × 10<sup>6</sup> cycles) and B2 (σ = 280 MPa, N = 8 × 10<sup>6</sup> cycles): (<b>B1-a</b>–<b>B1-c</b>) slip lines and ladder-like structure formed in the α<sub>lath</sub>/β<sub>trans</sub> interface; (<b>B1-d</b>) deformed α<sub>fine</sub> and dislocation tangles formed in the α<sub>fine</sub>/β<sub>trans</sub> interface; (<b>B1-e</b>,<b>B1-f</b>) the SAED and FFT) patterns of (<b>B1-d</b>); (<b>B1-g</b>–<b>B1-i</b>) GPA of (<b>B1-d</b>); (<b>B2-a</b>,<b>B2-b</b>) persistent slip bands (PSBs) and dislocation net formed within the α<sub>lath</sub>; (<b>B2-c</b>) zigzag structure formed in the α<sub>fine</sub>; (<b>B2-d</b>–<b>B2-f</b>) the SAED and FFT in the PSBs; (<b>B2-g</b>–<b>B2-i</b>) the GPA of (<b>B2-d</b>).</p>
Full article ">Figure 12
<p>The TEM images of fatigue crack initiation region of the interrupted specimen C1 (σ =230 MPa, N = 4 × 10<sup>6</sup> cycles) and C2 (σ =230 MPa, N = 8 × 10<sup>6</sup> cycles): (<b>C1</b>-<b>a</b>–<b>C1</b>-<b>c</b>) dislocations tangles and pinning within the α<sub>lath</sub>; (<b>C1-d</b>) zigzag structure formed in the α<sub>fine</sub>; (<b>C1-e</b>) the stacking faults and SAED within the α<sub>fine</sub>; (<b>C1-f</b>) the FFT patterns of (<b>C1-e</b>); (<b>C1-g</b>–<b>C1-i</b>) GPA of (<b>C1-e</b>); (<b>C2</b>-<b>a</b>–<b>C2</b>-<b>c</b>) ladder-like structure and slip lines formed in the α<sub>fine</sub>/β<sub>trans</sub> interface; (<b>C2-d</b>) dislocation walls formed within the α<sub>lath</sub>; (<b>C2-e</b>) dislocation lines formed within the α<sub>lath</sub>; (<b>C2-f</b>) deformed α<sub>fine</sub> and dislocation lines formed within the α<sub>lath</sub>.</p>
Full article ">Figure 13
<p>AFM images showing slip steps at aging temperatures 550 °C (<b>a</b>–<b>c</b>) and 650 °C (<b>d</b>–<b>f</b>) within α<sub>lath</sub>: (<b>a</b>,<b>d</b>) 2D AFM images; (<b>b</b>,<b>e</b>) 3D AFM images showing the detail of slip steps formed inner α<sub>lath</sub> and at α<sub>lath</sub>/β<sub>trans</sub> interface; (<b>c</b>,<b>f</b>) the height of slip steps in (<b>a</b>,<b>d</b>).</p>
Full article ">
28 pages, 3739 KiB  
Review
The Interplay of Molecular Factors and Morphology in Human Placental Development and Implantation
by Ioana Vornic, Victor Buciu, Cristian George Furau, Flavia Zara, Dorin Novacescu, Alina Cristina Barb, Alin Adrian Cumpanas, Silviu Constantin Latcu, Ioan Sas, Denis Serban, Talida Georgiana Cut and Cristina Stefania Dumitru
Biomedicines 2024, 12(12), 2908; https://doi.org/10.3390/biomedicines12122908 - 20 Dec 2024
Viewed by 1138
Abstract
The placenta is a vital organ that supports fetal development by mediating nutrient and gas exchange, regulating immune tolerance, and maintaining hormonal balance. Its formation and function are tightly linked to the processes of embryo implantation and the establishment of a robust placental-uterine [...] Read more.
The placenta is a vital organ that supports fetal development by mediating nutrient and gas exchange, regulating immune tolerance, and maintaining hormonal balance. Its formation and function are tightly linked to the processes of embryo implantation and the establishment of a robust placental-uterine interface. Recent advances in molecular biology and histopathology have shed light on the key regulatory factors governing these processes, including trophoblast invasion, spiral artery remodeling, and the development of chorionic villi. This review integrates morphological and molecular perspectives on placental development, emphasizing the roles of cytokines, growth factors, and signaling pathways, such as VEGF and Notch signaling, in orchestrating implantation and placental formation. The intricate interplay between molecular regulation and morphological adaptations highlights the placenta’s critical role as a dynamic interface in pregnancy. This review synthesizes current findings to offer clinicians and researchers a comprehensive understanding of the placenta’s role in implantation, emphasizing its importance in maternal-fetal medicine. By integrating these insights, the review lays the groundwork for advancing diagnostic and therapeutic approaches that can enhance pregnancy outcomes and address related complications effectively. Full article
(This article belongs to the Special Issue Role of Factors in Embryo Implantation and Placental Development)
Show Figures

Figure 1

Figure 1
<p>Development and Maturation of the Villous Tree.</p>
Full article ">Figure 2
<p>Histological section of the decidua at 20× magnification, showing large, polygonal decidual cells (indicated by red arrows) with abundant cytoplasm rich in glycogen and lipids.</p>
Full article ">Figure 3
<p>Histological section of the placenta at 20× magnification, illustrating the the structure of chorionic tertiary villi within the intervillous space. The image highlights various components essential for placental function, including the mesenchymal core (MC), cytotrophoblast cells (C), syncytiotrophoblast cells (S), decidual cells (D), and intervillous space (IV).</p>
Full article ">Figure 4
<p>Illustration of the developing placenta and early uteroplacental circulation. The left side shows an embryo within the amniotic cavity surrounded by the trophoblast and early placental structures. The right side zooms into the placental interface, highlighting the chorionic villi’s branching structure within the intervillous space, where maternal blood circulates to facilitate nutrient and gas exchange.</p>
Full article ">Figure 5
<p>The left image shows a placenta with a magistral pattern of chorionic plate vessels originating from a marginal insertion of the umbilical cord. In this configuration, blood vessels radiate from the edge of the placenta, distributing blood to the chorionic villi. The right image displays the maternal floor of a term placenta, revealing numerous well-defined cotyledons or lobules. These cotyledons, which are the functional units of the basal plate, contain chorionic villi essential for nutrient and gas exchange between maternal and fetal blood within the intervillous space. The organized arrangement of these cotyledons enhances placental efficiency by maximizing the surface area available for maternal-fetal exchange. Macroscopy images, representative of placental anatomy, obtained during the authors’ clinical practice within the Gynecology ward (private collection).</p>
Full article ">
21 pages, 9157 KiB  
Article
Numerical Study of Air Cushion Effect in Notched Disk Water Entry Process Using Structured Arbitrary Lagrangian–Eulerian Method
by Zhe Zhang, Nana Yang, Jinlong Ju, Xingzhi Bai, Houcun Zhou and Wenhua Wu
J. Mar. Sci. Eng. 2024, 12(11), 2012; https://doi.org/10.3390/jmse12112012 - 8 Nov 2024
Viewed by 614
Abstract
This paper presents a novel numerical investigation into the air cushion effect and impact loads during the water entry of notched discs, utilizing the Structured Arbitrary Lagrangian–Eulerian (S-ALE) algorithm in LS-DYNA. Unlike prior studies that focused on smooth or unnotched geometries, the present [...] Read more.
This paper presents a novel numerical investigation into the air cushion effect and impact loads during the water entry of notched discs, utilizing the Structured Arbitrary Lagrangian–Eulerian (S-ALE) algorithm in LS-DYNA. Unlike prior studies that focused on smooth or unnotched geometries, the present study explores how varying notch parameters influence the fluid–solid coupling process during high-speed water entry. The reliability and accuracy of the computational method are validated through grid independence verification and comparisons with experimental data and empirical formulas. Systematic analysis of the effects of notch size, water entry velocity, and entry angle on the evolution of the free surface, impact loads, and structural responses uncovers several novel findings. Notably, increasing the notch diameter significantly enhances the formation and stability of the air cushion, leading to a considerable reduction in peak impact loads—a phenomenon not previously quantified. Additionally, higher water entry Froude numbers are shown to accelerate air cushion compression and formation, markedly affecting free surface morphology and force distribution. The results also reveal that varying the water entry angle alters the air cushion’s morphological characteristics, where larger angles induce a more pronounced but less stable air cushion, influencing the internal structural response differently across regions. Full article
(This article belongs to the Section Ocean Engineering)
Show Figures

Figure 1

Figure 1
<p>Diagram of flat-plate impact into water.</p>
Full article ">Figure 2
<p>(<b>a</b>) Total plate force; (<b>b</b>) centre plate pressure.</p>
Full article ">Figure 3
<p>Air cushion effect of disc impact.</p>
Full article ">Figure 4
<p>(<b>a</b>) The disc forms a notch on the impact surface; (<b>b</b>) disc top view; (<b>c</b>) the disc hits the water, and the free surface is displaced by the air compression surface.</p>
Full article ">Figure 5
<p>Computational domain. (<b>a</b>) Boundary condition diagram. (<b>b</b>) The percentage of the mesh refinement area to the base size. (<b>c</b>) Initial VOF fractions: red represents air, blue represents water, and the board is white (dimensions are measured in disc diameter).</p>
Full article ">Figure 6
<p>(<b>a</b>) Computational efficiency for different grid size; (<b>b</b>) Comparison of pressure at different grid sizes.</p>
Full article ">Figure 7
<p>Time-step independent verification.</p>
Full article ">Figure 8
<p>(<b>a</b>) Punishment function mechanism diagram; (<b>b</b>) impact pressures for disc models with different <math display="inline"><semantics> <mrow> <msub> <mi>p</mi> <mi>f</mi> </msub> </mrow> </semantics></math> values.</p>
Full article ">Figure 9
<p>(<b>a</b>) The time history of the pressure of discs with different <math display="inline"><semantics> <mi>d</mi> </semantics></math> values during the attack process. (<b>b</b>) The standard deviation of peak pressure during disc attacks.</p>
Full article ">Figure 10
<p>(<b>a</b>) Evolution of free surface under different <math display="inline"><semantics> <mrow> <mi>d</mi> <mo>/</mo> <mi>D</mi> </mrow> </semantics></math> conditions; (<b>b</b>) a comparative analysis of the maximum air cushion thickness in relation to varying notch diameters.</p>
Full article ">Figure 11
<p>Schematic of flow field with different notch diameters. (<b>a</b>) Flow field pressure; (<b>b</b>) volume fraction of flow field.</p>
Full article ">Figure 12
<p>The impact force of the <math display="inline"><semantics> <mrow> <mi>d</mi> <mo>=</mo> <mn>60</mn> </mrow> </semantics></math> disc at different speeds.</p>
Full article ">Figure 13
<p>(<b>a</b>) Time series of impulse when <math display="inline"><semantics> <mrow> <mi>d</mi> <mo>=</mo> <mn>60</mn> </mrow> </semantics></math> at different impact velocities. All time series are centred on <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>, the moment when the pressure at the centre of the disc reaches its maximum. (<b>b</b>) The cumulative impulse of different <math display="inline"><semantics> <mi>d</mi> </semantics></math> during the first shock peak compared with the theoretical value [<a href="#B17-jmse-12-02012" class="html-bibr">17</a>,<a href="#B28-jmse-12-02012" class="html-bibr">28</a>,<a href="#B29-jmse-12-02012" class="html-bibr">29</a>].</p>
Full article ">Figure 14
<p>Evolution of the air cushion of a 60 mm disc with a groove diameter for different Fr values: (<b>a</b>) air cushion change at the same time t = 6 ms; (<b>b</b>) comparison of the moments of maximum air cushion thickness.</p>
Full article ">Figure 15
<p>(<b>a</b>) Pressure load distribution of d = 60 mm disc for different Fr values; (<b>b</b>) time course of air cushion thickness variation for different Fr values.</p>
Full article ">Figure 16
<p>(<b>a</b>) Resistance curve with time with different speeds; (<b>b</b>) resistance coefficient curve with time with different speeds.</p>
Full article ">Figure 17
<p>(<b>a</b>) Velocity variation in the disc during water entry at different tilt angles (3°, 5°, 6°); (<b>b</b>) angle change in the disc during water entry with different initial inclination angles.</p>
Full article ">Figure 18
<p>Stress cloud images for a disc with <math display="inline"><semantics> <mrow> <mi>d</mi> <mo>=</mo> <mn>60</mn> <mo> </mo> <mi>mm</mi> </mrow> </semantics></math> and velocity of 2 m/s tilted at different angles (3°, 5°).</p>
Full article ">Figure 19
<p>Time chart of force: (<b>a</b>) velocity: 3 m/s, <math display="inline"><semantics> <mrow> <mi>θ</mi> <mo>=</mo> <mn>3</mn> <mo>°</mo> </mrow> </semantics></math>, distance y = 0 (the midpoint of the <span class="html-italic">Y</span>-axis direction of the disc; the tilt is positive), and load curve at different distances; (<b>b</b>) load curve at a speed of 2 m/s and <math display="inline"><semantics> <mrow> <mi>θ</mi> <mo>=</mo> <mn>3</mn> <mo>°</mo> </mrow> </semantics></math>.</p>
Full article ">Figure 20
<p>Evolution of free liquid surface when the disc with a notch <math display="inline"><semantics> <mrow> <mi>d</mi> <mo>=</mo> <mn>60</mn> <mo> </mo> <mi>mm</mi> </mrow> </semantics></math> is impacted at different attitude angles.</p>
Full article ">
17 pages, 999 KiB  
Review
“A Friend Among Strangers” or the Ambiguous Roles of Runx2
by Kseniia Azarkina, Ekaterina Gromova and Anna Malashicheva
Biomolecules 2024, 14(11), 1392; https://doi.org/10.3390/biom14111392 - 31 Oct 2024
Viewed by 1290
Abstract
The transcription factor Runx2 plays a crucial role in regulating osteogenic differentiation and skeletal development. This factor not only controls the expression of genes involved in bone formation, but also interacts with signaling pathways such as the Notch pathway, which are essential for [...] Read more.
The transcription factor Runx2 plays a crucial role in regulating osteogenic differentiation and skeletal development. This factor not only controls the expression of genes involved in bone formation, but also interacts with signaling pathways such as the Notch pathway, which are essential for body development. However, studies have produced conflicting results regarding the relationship between Runx2 and the Notch pathway. Some studies suggest a synergistic interaction between these molecules, while others suggest an inhibitory one, for example, the interplay between Notch signaling, Runx2, and vitamin D3 in osteogenic differentiation and bone remodeling. The findings suggest a complex relationship between Notch signaling and osteogenic differentiation, with ongoing research needed to clarify the mechanisms involved and resolve existing contradictions regarding role of Notch in this process. Additionally, there is increasing evidence of contradictory roles for Runx2 in various tissues and organs, both under normal conditions and in pathological states. This diversity of roles makes Runx2 a potential therapeutic target, offering new directions for research. In this review, we have discussed the mechanisms of osteogenic differentiation and the important role of Runx2 in this process. We have also examined its relationship with different signaling pathways. However, there are still many uncertainties and inconsistencies in our current understanding of these interactions. Additionally, given that Runx2 is also involved in numerous other events in various tissues, we have tried to comprehensively examine its functions outside the skeletal system. Full article
Show Figures

Figure 1

Figure 1
<p>Two possible options for the interaction of Notch signaling pathway targets (Hes1, Hey1) with Runx2. (<b>a</b>) Hrt transcription factors (Hes1, Hey1) suppress the transcriptional activity of <span class="html-italic">Runx2</span>. (<b>b</b>) Runx2 and Hes1 physically interact and enhance basal and 1,25-(OH)<sub>2</sub>-vitamin D3-induced (VD3) transcription of a gene associated with osteodifferentiation. “Created in BioRender. Malashicheva, A. (2024) <a href="https://BioRender.com/g44l299" target="_blank">https://BioRender.com/g44l299</a>”.</p>
Full article ">Figure 2
<p>Tissue-specific effect of the Runx2 transcription factor and its associations with various signaling pathways and target genes. “Created in BioRender. Malashicheva, A. (2024) <a href="https://BioRender.com/t84w530" target="_blank">https://BioRender.com/t84w530</a>”.</p>
Full article ">
13 pages, 2539 KiB  
Article
Differential Effects of Four Canonical Notch-Activating Ligands on c-Kit+ Cardiac Progenitor Cells
by Matthew Robeson, Steven L. Goudy and Michael E. Davis
Int. J. Mol. Sci. 2024, 25(20), 11182; https://doi.org/10.3390/ijms252011182 - 17 Oct 2024
Viewed by 1050
Abstract
Notch signaling, an important signaling pathway in cardiac development, has been shown to mediate the reparative functions of c-kit+ progenitor cells (CPCs). However, it is unclear how each of the four canonical Notch-activating ligands affects intracellular processes in c-kit+ cells when used as [...] Read more.
Notch signaling, an important signaling pathway in cardiac development, has been shown to mediate the reparative functions of c-kit+ progenitor cells (CPCs). However, it is unclear how each of the four canonical Notch-activating ligands affects intracellular processes in c-kit+ cells when used as an external stimulus. Neonatal c-kit+ CPCs were stimulated using four different chimeric Notch-activating ligands tethered to Dynabeads, and the resulting changes were assessed using TaqMan gene expression arrays, with subsequent analysis by principal component analysis (PCA). Additionally, functional outcomes were measured using an endothelial cell tube formation assay and MSC migration assay to assess the paracrine capacity to stimulate new vessel formation and recruit other reparative cell types to the site of injury. Gene expression data showed that stimulation with Jagged-1 is associated with the greatest pro-angiogenic gene response, including the expression of VEGF and basement membrane proteins, while the other canonical ligands, Jagged-2, Dll-1, and Dll-4, are more associated with regulatory and epigenetic changes. The functional assay showed differential responses to the four ligands in terms of angiogenesis, while none of the ligands produced a robust change in migration. These data demonstrate how the four Notch-activating ligands differentially regulate CPC gene expression and function. Full article
(This article belongs to the Special Issue Notch Signaling Pathways)
Show Figures

Figure 1

Figure 1
<p>Fc-Jagged functionalized beads bind and activate Notch in vitro. (<b>A</b>) Fc-Jagged1 functionalized beads aggregate on the surface of c-kit+ hCPCs compared to nonspecific human IgG functionalized beads. (<b>B</b>,<b>C</b>) Fluorescent YFP is expressed when cells are stimulated with Jagged-1 beads. The response is attenuated when the small molecule Notch inhibitor DAPT (10 µmol/L) is added to cell media. N = 6 per group. (***) <span class="html-italic">p</span> &lt; 0.005, one-way ANOVA.</p>
Full article ">Figure 2
<p>The four canonical Notch ligands drive substantially different gene expression profiles in c-kit+ hCPCs. (<b>A</b>) HES1 expression is upregulated in all four treatment groups, though Jag1 and Jag2 drive the strongest response. HEY1 expression is upregulated by Jag1, Jag2, and Dll4, but not Dll1. N = 4 per group, (**) <span class="html-italic">p</span> &lt; 0.005, (*) <span class="html-italic">p</span> &lt; 0.05, (n.s.) not significant, one-way ANOVA. (<b>B</b>) HES1 protein level increases upon activation of Notch by Jag1/Jag2/Dll1/Dll4. (<b>C</b>) Heat map demonstrating gene expression in response to all 4 ligands. Gene expression was normalized to IgG control beads and presented as fold change over control.</p>
Full article ">Figure 3
<p>PCA reveals clusters of genes that co-vary closely with particular Notch-activating ligands. (<b>A</b>) Genes associated with pro-angiogenic factors co-vary with Jag1 stimulation. (<b>B</b>) Genes associated with ECM remodeling co-vary with Jag1 stimulation. (<b>C</b>) Genes associated with Notch receptors and ligands co-vary with Jag2/Dll1/Dll4 stimulation. (<b>D</b>) Genes associated with Hitone remodeling co-vary with Jag2/Dll1/Dll4 stimulation.</p>
Full article ">Figure 4
<p>Endothelial cell tube formation is enhanced across all treatment groups compared to IgG, but differences between the four canonical ligands are not significant. (<b>A</b>) Representative fluorescent images of endothelial cell tubes after 6 h. (<b>B</b>) Quantified average tube length. N = 16 for each treatment group. (*) <span class="html-italic">p</span> &lt; 0.05, (**) <span class="html-italic">p</span> &lt; 0.005, (****) <span class="html-italic">p</span> &lt; 0.00005, (ns) Not significant, one-way ANOVA.</p>
Full article ">Figure 5
<p>MSC migration in response to conditioned media is upregulated across all treatment groups, though differences between groups are non-significant. (<b>A</b>) MSCs are seeded with serum-free media in the upper well of a transwell insert, with conditioned media positioned below. Migration is quantified after overnight incubation. (<b>B</b>) MSC migration quantified by fluorescence measurement.</p>
Full article ">Figure 6
<p>Illustrated design of the study. Protein G Dynabeads are functionalized overnight by incubation with Fc-conjugated Notch-activating ligands. Following functionalization, c-kit+ CPCs are incubated with functionalized beads in a well plate for 48 h. At the conclusion of this, RNA and conditioned media are collected to perform gene expression analysis via PCR array, endothelial cell tube formation assays, and a transwell assay to assess the migration of MSCs.</p>
Full article ">
10 pages, 6933 KiB  
Article
Role of Coalesced Bainite in Hydrogen Embrittlement of Tempered Martensitic Steels
by Hee-Chang Shin, Sang-Gyu Kim and Byoungchul Hwang
Metals 2024, 14(10), 1171; https://doi.org/10.3390/met14101171 - 15 Oct 2024
Viewed by 902
Abstract
This study investigates the role of coalesced bainite in enhancing the hydrogen embrittlement resistance of tempered martensitic steels. By analyzing the microstructural characteristics and mechanical properties under varying cooling rates, it was found that the presence of coalesced bainite significantly impedes hydrogen accumulation [...] Read more.
This study investigates the role of coalesced bainite in enhancing the hydrogen embrittlement resistance of tempered martensitic steels. By analyzing the microstructural characteristics and mechanical properties under varying cooling rates, it was found that the presence of coalesced bainite significantly impedes hydrogen accumulation at prior austenite grain boundaries. This leads to a transition in the fracture mode from intergranular to transgranular, thereby improving the overall resistance to hydrogen embrittlement in steels. Slow strain rate tests (SSRTs) on both smooth and notched specimens further confirmed that steels cooled at lower rates, which form a higher fraction of coalesced bainite, exhibiting superior hydrogen embrittlement resistance. These findings suggest that optimizing the cooling process to promote coalesced bainite formation could be a valuable strategy for enhancing the performance of tempered martensitic steels in hydrogen-rich environments. Full article
(This article belongs to the Special Issue Recent Insights into Mechanical Properties of Metallic Alloys)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Continuous cooling transformation (CCT) diagram of the investigated steels (M, martensite; B, bainite; A, austenite; F, ferrite; P, pearlite) and (<b>b</b>) the phase fraction plotted as a function of different cooling rates of 3.5 °C/s, 5.3 °C/s, and 7.5 °C/s. Both phase diagrams were calculated by thermodynamic calculation software.</p>
Full article ">Figure 2
<p>SEM micrographs of the steels under different cooling rates of (<b>a</b>) 3.5 °C/s, (<b>b</b>) 5.3 °C/s, and (<b>c</b>) 7.5 °C/s. (<b>d</b>) Magnified view showing coalesced bainite (CB) in the white dotted box for the (<b>a</b>) micrograph. (<b>e</b>,<b>f</b>) Electron channeling contrast imaging (ECCI) micrographs for the steels with a cooling rate of 3.5 °C/s. TB refers to tempered bainite, and TM indicates tempered martensite.</p>
Full article ">Figure 3
<p>(<b>a</b>–<b>c</b>) Inverse pole figure (IPF) and (<b>d</b>) image quality (IQ) maps obtained from electron backscatter diffraction (EBSD) results for the steels with a cooling rate of 3.5 °C/s. The (<b>b</b>,<b>c</b>) IPF maps magnified in the (<b>a</b>) IPF map (TM; tempered martensite, CB; coalesced bainite, TB; tempered bainite). The (<b>e</b>,<b>f</b>) misorientation and image quality (IQ) value distributions of coalesced bainite (CB) relative to surrounding microstructures, which are profiled along blue arrows in (<b>b</b>) and (<b>c</b>), respectively.</p>
Full article ">Figure 4
<p>(<b>a</b>) X-ray diffraction (XRD) profile results and (<b>b</b>–<b>d</b>) Williamson–Hall plots for the steels under different cooling rates of 3.5 °C/s, 5.3 °C/s, and 7.5 °C/s.</p>
Full article ">Figure 5
<p>Slow strain rate test (SSRT) results of non-charged and H-charged tensile specimens for the (<b>a</b>) smooth and (<b>b</b>) notch for the steels under different cooling rates of 3.5 °C/s, 5.3 °C/s, and 7.5 °C/s.</p>
Full article ">Figure 6
<p>(<b>a</b>) Variations of tensile properties and (<b>b</b>) relative index of hydrogen embrittlement resistance measured from the slow strain rate test (SSRT) results for the steels under different cooling rates of 3.5 °C/s, 5.3 °C/s, and 7.5 °C/s.</p>
Full article ">Figure 7
<p>SEM fractographs of the H-charged tensile specimens of (<b>a</b>,<b>b</b>) smooth and (<b>c</b>,<b>d</b>) notch for the steels under different cooling rates of (<b>a</b>,<b>c</b>) 3.5 °C/s and (<b>b</b>,<b>d</b>) 7.5 °C/s.</p>
Full article ">Figure 8
<p>Schematic illustration explaining the fracture mechanism of the smooth and notch tensile specimens.</p>
Full article ">
30 pages, 1907 KiB  
Review
Molecular Signaling Pathways and MicroRNAs in Bone Remodeling: A Narrative Review
by Monica Singh, Puneetpal Singh, Baani Singh, Kirti Sharma, Nitin Kumar, Deepinder Singh and Sarabjit Mastana
Diseases 2024, 12(10), 252; https://doi.org/10.3390/diseases12100252 - 12 Oct 2024
Viewed by 1879
Abstract
Bone remodeling is an intricate process executed throughout one’s whole life via the cross-talk of several cellular events, progenitor cells and signaling pathways. It is an imperative mechanism for regaining bone loss, recovering damaged tissue and repairing fractures. To achieve this, molecular signaling [...] Read more.
Bone remodeling is an intricate process executed throughout one’s whole life via the cross-talk of several cellular events, progenitor cells and signaling pathways. It is an imperative mechanism for regaining bone loss, recovering damaged tissue and repairing fractures. To achieve this, molecular signaling pathways play a central role in regulating pathological and causal mechanisms in different diseases. Similarly, microRNAs (miRNAs) have shown promising results in disease management by mediating mRNA targeted gene expression and post-transcriptional gene function. However, the role and relevance of these miRNAs in signaling processes, which regulate the delicate balance between bone formation and bone resorption, are unclear. This review aims to summarize current knowledge of bone remodeling from two perspectives: firstly, we outline the modus operandi of five major molecular signaling pathways, i.e.,the receptor activator of nuclear factor kappa-B (RANK)-osteoprotegrin (OPG) and RANK ligand (RANK-OPG-RANKL), macrophage colony-stimulating factor (M-CSF), Wnt/β-catenin, Jagged/Notch and bone morphogenetic protein (BMP) pathways in regards to bone cell formation and function; and secondly, the miRNAs that participate in these pathways are introduced. Probing the miRNA-mediated regulation of these pathways may help in preparing the foundation for developing targeted strategies in bone remodeling, repair and regeneration. Full article
Show Figures

Figure 1

Figure 1
<p>The figure shows (a) Wnt signaling in which mesenchymal stem cells are differentiated into osteoblast precursor cells and transformed to osteoblasts. Prompted by Wnt proteins, these transformations are induced however by the inhibition of TNFα-directed DKK-1 and the drug Macitentan. (b) RANK-RANKL-OPG signaling in which hematopoietic stem cells are differentiated into osteoclast precursor cells, transforming them into osteoclasts. This transformation is stimulated by RANKL but inhibited by OPG, both of which are secreted by osteoblasts. (c) M-CSF signaling in which osteoclast precursor cells are transformed into osteoclasts and this transformation is augmented by M-CSF, which is secreted by osteoblasts.</p>
Full article ">Figure 2
<p>The figure shows the process of osteoclastogenesis and osteoblast differentiation by miRNAs through interaction of different signaling pathways.</p>
Full article ">Figure 3
<p>The figure illustrates Notch/Jagged signaling in which mesenchymal stem cells are differentiated into osteoblasts and are transformed into osteocytes through mineralization. Notch activation plays a dual role in this transformation by inducing it but also halting it by targeting Wnt/β-catenin pathway. On the other hand, macrophage precursors are differentiated into osteoclasts that are transformed into osteocytes.Here, notch activation also plays a dual role by inciting it either through its Notch2/Dll1 complex or by providing Notch1/Notch1–3-deficient conditions in precursor cells, which affect RANKL/OPG and M-CSF signaling. However, it is inhibited by either knocking RANK or pumping M-CSF signaling or through its Notch1/Jagged1 complex. When activating Wnt/β-catenin pathway by suppressing SOST and DKK1, this transformation is inhibited.</p>
Full article ">Figure 4
<p>The figure depicts TGFβ/BMP signaling in whichSMAD-dependent pathway is initiated by binding of TGFβ ligand with TGFβ receptors R1 and R2 and BMP ligand with Type I/II receptor. This is followed by phosphorylation of R-Smad 2 and 3 and R-Smad 1,5 and 8 and the complexing of Co-Smad 4 with both molecules to form the activated quaternary complex, which modulates gene expression, affecting bone remodeling. Meanwhile, in the non-canonical pathway, the same ligand–receptor binding activates other signaling molecules, such as ERK-MAPK, PI3K, JNK, p38 and Cd42, by phosphorylating them. These activated molecules then alter expression of various bone-related genes.</p>
Full article ">
15 pages, 2594 KiB  
Article
Myoblast-Derived Galectin 3 Impairs the Early Phases of Osteogenesis Affecting Notch and Akt Activity
by Emanuela Amore, Vittoria Cenni, Manuela Piazzi, Michele Signore, Giulia Orlandi, Simona Neri, Stefano Biressi, Rosario Barone, Valentina Di Felice, Matilde Y. Follo, Jessika Bertacchini and Carla Palumbo
Biomolecules 2024, 14(10), 1243; https://doi.org/10.3390/biom14101243 - 30 Sep 2024
Viewed by 1326
Abstract
Galectin-3 (Gal-3) is a pleiotropic lectin produced by most cell types, which regulates multiple cellular processes in various tissues. In bone, depending on its cellular localization, Gal-3 has a dual and opposite role. If, on the one hand, intracellular Gal-3 promotes bone formation, [...] Read more.
Galectin-3 (Gal-3) is a pleiotropic lectin produced by most cell types, which regulates multiple cellular processes in various tissues. In bone, depending on its cellular localization, Gal-3 has a dual and opposite role. If, on the one hand, intracellular Gal-3 promotes bone formation, on the other, its circulating form affects bone remodeling, antagonizing osteoblast differentiation and increasing osteoclast activity. From an analysis of the secretome of cultured differentiating myoblasts, we interestingly found the presence of Gal-3. After that, we confirmed that Gal-3 was expressed and released in the extracellular environment from myoblast cells during their differentiation into myotubes, as well as after mechanical strain. An in vivo analysis revealed that Gal-3 was triggered by trained exercise and was specifically produced by fast muscle fibers. Speculating a role for this peptide in the muscle-to-bone cross talk, a direct co-culture in vitro system, simultaneously combining media that were obtained from differentiated myoblasts and osteoblast cells, confirmed that Gal-3 is a mediator of osteoblast differentiation. Molecular and proteomic analyses revealed that the secreted Gal-3 modulated the biochemical processes occurring in the early phases of bone formation, in particular impairing the activity of the STAT3 and PDK1/Akt signaling pathways and, at the same time, triggering that one of Notch. Circulating Gal-3 also affected the expression of the most common factors involved in osteogenetic processes, including BMP-2, -6, and -7. Intriguingly, Gal-3 was able to interfere with the ability of differentiating osteoblasts to interact with the components of the extracellular bone matrix, a crucial condition required for a proper osteoblast differentiation. All in all, our evidence lays the foundation for further studies to present this lectin as a novel myokine involved in muscle-to-bone crosstalk. Full article
(This article belongs to the Section Molecular Biology)
Show Figures

Figure 1

Figure 1
<p>Muscle differentiation and mechanical stimulation trigger Gal-3 expression and release. (<b>A</b>) The C2C12 cells were differentiated for the indicated days. At the end of each time point, the cells and culture media were collected. The cells were lysed and, together with the culture media, were resolved by SDS-PAGE and explored for the Gal-3 level of expression. β-Tubulin and MyoD were used as loading and differentiation markers, respectively. Original images can be found in <a href="#app1-biomolecules-14-01243" class="html-app">Supplementary File 1</a>. (<b>B</b>) The densitometric values of intracellular and secreted Gal-3 were graphed in the bar histogram, shown aside. The densitometric results were normalized on the values of corresponding β-tubulin. * <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. (<b>C</b>,<b>D</b>) The differentiated C2C12 cells were subjected to multiaxial stretching (st) for 6 and 24 h, or left untreated (nt). At the end of the stimuli, the cells were lysed and, together with corresponding media, were subjected to an immunoblot analysis to verify the Gal-3 expression. The bars are relative to three different experiments, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01. (<b>E</b>) The cells, as in (<b>C</b>), were lysed in Trizol and subjected to an RT-PCR analysis for monitoring the level of expression of Gal-3 mRNA. The bars show the relative amount of Gal-3 mRNA. (<span class="html-italic">n</span> = 3 *** <span class="html-italic">p</span> ≤ 0.001). Original images can be found in <a href="#app1-biomolecules-14-01243" class="html-app">Supplementary File 1</a>.</p>
Full article ">Figure 2
<p>Endurance training increases Gal-3 expression in muscle fibers of type IIB. (<b>A</b>,<b>B</b>) The tibialis anterior muscles from the sedentary (Sed) and trained (Tr) mice groups were analyzed to determine the intracellular level of Gal-3 expression. The bar graph represents the average OD of the samples of the two groups, normalized to the corresponding values of GAPDH. Original images can be found in <a href="#app1-biomolecules-14-01243" class="html-app">Supplementary File 1</a>. (<b>C</b>,<b>D</b>) A Gal-3 immunohistochemical analysis of the red (RTA) and white fibers (WTA) of the tibialis anterior muscles from the sedentary and trained mice groups with the images’ signal intensity quantification. (<b>E</b>,<b>F</b>) The Gal-3 immunohistochemical analysis of the red (RGA) and white fibers (WGA) of the gastrocnemius muscles from the sedentary and trained mice groups, with the images’ signal intensity quantification. Enlarged areas are present within the most significant panels. Bars 250 micron. A statistically significant increase in Gal-3 expression was seen in the white fibers of both muscle types in the trained mice group (<span class="html-italic">n</span> = 3, * <span class="html-italic">p</span> &lt; 0.05), when compared to the group of sedentary mice.</p>
Full article ">Figure 3
<p>Gal-3 secreted by C2C12 cells impairs the osteogenic potential of the MC-3T3 osteoblast progenitor cells as a recombinant and as the soluble form of Gal-3. (<b>A</b>,<b>B</b>) The MC-3T3 cells were differentiated in the presence of a recombinant soluble form of Gal-3 (rec-Gal-3, FC 3 microg/mL) or another vehicle (vehicle) (<span class="html-italic">n</span> = 3, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001) for three or seven days (D3 and D7, respectively). Where indicated, a combination of rec-Gal-3 and a specific antibody against Gal-3 was added (Gal-3 Ab, rabbit anti-Gal-3 1 µg/mL FC 3 µg/mL). As control, the MC-3T3 osteoblast progenitors were also collected at D0. The matrix bone mineralization was analyzed by alizarin red (AR) staining. After staining, the alizarin was diluted, quantified as described in the “Methods” Section, and the values were presented as a graph (<span class="html-italic">n</span> = 3, *** <span class="html-italic">p</span> &lt; 0.001). (<b>C</b>) The MC-3T3 osteoblast progenitors differentiated in the presence of rec-Gal-3 or another vehicle were harvested at D0, D3, and D7 and collected. One half of the cells’ pellet was then lysed and the total lysates were resolved and assayed for bone-alkaline phosphatase (ALP) and the Osterix level of expression. GAPDH was used as equal loading control. Original images can be found in <a href="#app1-biomolecules-14-01243" class="html-app">Supplementary File 1</a>. (<b>D</b>–<b>F</b>) The densitometric analysis of the level of expression of ALP and Osterix, normalized to GAPDH. (<b>E</b>) The other half of the cell pellets were lysed, and the ALP activity was monitored (<span class="html-italic">n</span> = 3, ** <span class="html-italic">p</span> ≤ 0.01). (<b>G</b>,<b>H</b>) The MC-3T3 osteoblast progenitors differentiated in the presence of rec-Gal-3 or another vehicle were harvested at D0, D3, and D7, lysed, and evaluated for Notch1 and cleaved Notch1 protein expression; a densitometric analysis was carried out, normalized on GAPDH levels, ** <span class="html-italic">p</span> ≤ 0.01. Original images can be found in <a href="#app1-biomolecules-14-01243" class="html-app">Supplementary File 1</a>. (<b>I</b>) The graph bar of the fold change of the level of expression of <span class="html-italic">Hey1</span> and <span class="html-italic">c-Myc</span> RNAs in D0- and D3-differentiated cells (with or without rec-Gal-3), relative to three different experiments (<span class="html-italic">n</span> = 3, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> ≤ 0.01). HPRT was used as a normalization housekeeping gene. (<b>J</b>,<b>K</b>) The MC-3T3 osteoblast precursor cells were maintained in the growth medium D0 and induced to differentiate in a canonical differentiation medium, D7, supplemented with the supernatant from the C2C12 cells that were overexpressing FLAG-Gal-3 (FLAG-Gal-3 OE, transfected for 48 h) or, as a control, from the C2C12 cells transfected with the empty vector (EV, transfected for 48 h). The differentiation medium (D7) was conditioned with the same amount of the C2C12 medium (EV and FLAG-Gal-3) in a medium ratio of 1:3. Samples were collected on the day of the change of the medium, D0, or at D7. The matrix bone mineralization was analyzed by alizarin red (AR) staining. The graph bars indicate the amount of AR staining (<span class="html-italic">n</span> = 3, *** <span class="html-italic">p</span> &lt; 0.001). (<b>L</b>) The MC-3T3 osteoblast progenitors were induced to differentiate in a canonical differentiation medium, D3, supplemented with the supernatant from the C2C12 cells that were subjected or not to a stretching protocol for 6 h. The differentiation medium (D3) was conditioned with one of the three C2C12 media (C2C12 media: not stretched, stretched for 6 h, or stretched for 6 h + neutralizing antibody). The total lysates were resolved and assayed for the bone-alkaline phosphatase (ALP) level of activity (<span class="html-italic">n</span> = 3, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> ≤ 0.01). (<b>M</b>) An RT-qPCR analysis of RUNX2 expression in the MC-3T3 cells differentiated as in M) (<span class="html-italic">n</span> = 3, ** <span class="html-italic">p</span> ≤ 0.01).</p>
Full article ">Figure 4
<p>Rec-Gal-3 modulates the expression of proteins with fundamental cellular functions during the differentiation of MC-3T3 osteoprogenitor cells. (<b>A</b>) A Venn diagram showing the proteins identified by an LC-MS analysis. We identified 117 proteins on D0, 366 on D3, and 353 on D3 + rec-Gal-3. (<b>B</b>) The list of proteins up- or down-regulated in the MC-3T3 cells after 3 days of differentiation with recombinant Gal-3, compared to D3. Only the proteins identified with a fold increase ≥ 4 were reported.</p>
Full article ">Figure 5
<p>Mechanistic insights into circulating Gal-3 functions during the differentiation of MC-3T3 osteoprogenitor cells. (<b>A</b>) Heat map showing the phosphorylation of the main signaling pathways detected by the RPPA analysis during the growing state (D0), the early (D3), and the middle (D7) phases of the differentiation of the MC-3T3 cells with or without rec-Gal-3. Squared phosphorylated proteins are the species most impaired by Gal-3 supplementation. The values are reported as the mean of signal intensities, with <span class="html-italic">n</span> = 3. (<b>B</b>) The level of expression of the proteins promoting osteogenesis during the differentiation of the MC-3T3 cells with or without rec-Gal-3 (+rec-Gal-3) through the “Bone Metabolism Antibody Array”. The cells were blocked on D0 and D3 (<span class="html-italic">n</span> = 3, * <span class="html-italic">p</span> ≤ 0.05; ** <span class="html-italic">p</span> ≤ 0.01).</p>
Full article ">Figure 6
<p>rec-Gal-3 affects the ability of differentiated osteoblasts to interact with ECM components. The evaluation of the ability of the MC-3T3 cells maintained in a growth medium (D0), differentiated for 3 days (D3) with or without recombinant Gal-3 (D3 + rec-Gal-3) to interact with the typical components of the extracellular bone matrix (Collagen I, Fibronectin, Laminin) through the “ECM adhesion assay” (<span class="html-italic">n</span> = 2, * <span class="html-italic">p</span> ≤ 0.05; ** <span class="html-italic">p</span> ≤ 0.01).</p>
Full article ">
Back to TopTop