[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (1,173)

Search Parameters:
Keywords = YAP

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
10 pages, 374 KiB  
Article
Concordance of HER2 Expression in Paired Primary and Metastatic Sites of Endometrial Serous Carcinoma and the Effect of Intratumoral Heterogeneity
by Francis Hong Xin Yap, Yancey Wilson, Joanne Peverall, Benhur Amanuel, Ben Allanson and Sukeerat Ruba
J. Mol. Pathol. 2024, 5(3), 405-414; https://doi.org/10.3390/jmp5030027 (registering DOI) - 14 Sep 2024
Viewed by 194
Abstract
Primary endometrial serous carcinoma, known for its aggressive nature and poor prognosis, shares similarities with breast and gastric cancers in terms of potential HER2 overexpression as a therapeutic target. Assessing HER expression is complicated by tumor heterogeneity and discrepancies between primary and metastatic [...] Read more.
Primary endometrial serous carcinoma, known for its aggressive nature and poor prognosis, shares similarities with breast and gastric cancers in terms of potential HER2 overexpression as a therapeutic target. Assessing HER expression is complicated by tumor heterogeneity and discrepancies between primary and metastatic sites. In this study, we retrospectively analyzed HER amplification and expression in 16 pairs of primary endometrial serous carcinoma resections and corresponding metastases. HER2 status was determined using immunohistochemistry (IHC), with criteria based on the percentage and intensity of tumor cell staining. Confirmatory techniques, such as dual in situ hybridization (DISH) and fluorescence in situ hybridization (FISH), were also employed. This study reports on the concordance rates and the presence and pattern of HER2 heterogeneity. Our results showed an 87.5% concordance rate in HER2 amplification status between primary and metastatic sites, with 33% of cases scored as 2+ being amplified. Heterogeneity was observed in 100% of amplified cases and 95% of non-amplified cases on in situ testing, with variations in heterogeneity patterns between techniques. In conclusion, our findings emphasize the importance of testing both primary and metastatic sites or recurrences, with a concordance rate of 87.5%. In addition, a review of the literature and combining the results showed a concordance rate of up to 68%. The presence and pattern of heterogeneity, particularly in cases of mosaic or clustered heterogeneity in the primary tumor, may serve as reliable indicators of concordance, predicting a non-amplified HER2 status in corresponding metastases. Full article
Show Figures

Figure 1

Figure 1
<p>HER2 testing algorithm in endometrial serous carcinoma.</p>
Full article ">
9 pages, 204 KiB  
Article
Health-Related Quality of Life with Six Domains: A Comparison of Healthcare Providers without Chronic Diseases and Participants with Chronic Diseases
by Mohamad Adam Bujang, Yoon Khee Hon, Wei Hong Lai, Eileen Pin Pin Yap, Xun Ting Tiong, Selvasingam Ratnasingam, Alex Ren Jye Kim, Masliyana Husin, Yvonne Yih Huan Jee, Nurul Fatma Diyana Ahmad, Cheng Hoon Chew, Clare Hui Hong Tan, Sing Yee Khoo, Fazalena Johari and Alan Yean Yip Fong
J. Clin. Med. 2024, 13(18), 5398; https://doi.org/10.3390/jcm13185398 - 12 Sep 2024
Viewed by 272
Abstract
Background/Objectives: This study aims to compare the health-related quality of life (HRQOL) between healthcare providers without chronic diseases and participants with chronic diseases presenting with one of the four different primary diagnoses on the health-related quality of life with six domains (HRQ-6D) scale. [...] Read more.
Background/Objectives: This study aims to compare the health-related quality of life (HRQOL) between healthcare providers without chronic diseases and participants with chronic diseases presenting with one of the four different primary diagnoses on the health-related quality of life with six domains (HRQ-6D) scale. Methods: This is a cross-sectional study to compare the HRQOL between healthcare providers without chronic diseases and participants with chronic diseases. Data collection was performed from May 2022 to May 2023. Data for the comparison group were taken from healthcare providers without chronic diseases, and for the participant group with chronic diseases, the data were collected from actual patients with one of four types of primary diagnoses who were recruited from specialist cardiology, oncology, psychiatry, and nephrology clinics. All the participants of this study filled in the HRQ-6D. Results: There were 238 (58.6%) healthcare providers without chronic diseases who participated in this study, as well as 41 (10.1%) patients with end-stage renal disease (ESRD), 48 (11.8%) patients with cancer, and 40 (9.9%) patients who were depressed, and the remaining patients had heart disease. The means (SD) of HRQ-6D scores among healthcare providers without chronic diseases for pain, physical strength, emotion, mobility, self-care, perception of future health, and overall HRQ-6D score were 75.3% (19.8), 74.5% (21.1), 85.6% (18.4%), 93.0% (12.3), 91.6% (13.9), 74.2% (23.3), and 82.4% (13.6), respectively. In comparisons between healthcare providers without chronic diseases and participants with chronic diseases, all mean differences of the overall HRQ-6D score and its domains and dimensions were statistically significant (p < 0.001). Conclusions: The overall score of the HRQ-6D, as well as its domains and dimensions are sensitive in detecting the study participants with chronic diseases from among those without chronic diseases. Therefore, the HRQ-6D is a reliable and valid scale to measure HRQOL. Future studies may use this scale for interventional, observational, and cost-effectiveness studies. Full article
16 pages, 1972 KiB  
Perspective
Navigating the Healthcare Metaverse: Immersive Technologies and Future Perspectives
by Kevin Yi-Lwern Yap
Virtual Worlds 2024, 3(3), 368-383; https://doi.org/10.3390/virtualworlds3030020 - 11 Sep 2024
Viewed by 1032
Abstract
The year is 2030. The internet has evolved into the metaverse. People navigate through advanced avatars, shop in digital marketplaces, and connect with others through extended reality social media platforms. Three-dimensional patient scans, multidisciplinary tele-collaborations, digital twins and metaverse health records are part [...] Read more.
The year is 2030. The internet has evolved into the metaverse. People navigate through advanced avatars, shop in digital marketplaces, and connect with others through extended reality social media platforms. Three-dimensional patient scans, multidisciplinary tele-collaborations, digital twins and metaverse health records are part of clinical practices. Younger generations regularly immerse themselves in virtual worlds, playing games and attending social events in the metaverse. This sounds like a sci-fi movie, but as the world embraces immersive technologies post-COVID-19, this future is not too far off. This article aims to provide a foundational background to immersive technologies and their applications and discuss their potential for transforming healthcare and education. Moreover, this article will introduce the metaverse ecosystem and characteristics, and its potential for health prevention, treatment, education, and research. Finally, this article will explore the synergy between generative artificial intelligence and the metaverse. As younger generations of healthcare professionals embrace this digital frontier, the metaverse’s potential in healthcare is definitely attractive. Mainstream adoption may take time, but it is imperative that healthcare professionals be equipped with interdisciplinary skills to navigate the plethora of immersive technologies in the future of healthcare. Full article
(This article belongs to the Special Issue Serious Games and Extended Reality in Healthcare and/or Education)
Show Figures

Figure 1

Figure 1
<p>The reality–virtuality (RV) continuum.</p>
Full article ">Figure 2
<p>Differences between non-immersive and immersive virtual reality (VR).</p>
Full article ">Figure 3
<p>Types of augmented reality (AR) technologies.</p>
Full article ">Figure 4
<p>The four types of metaverse characteristics.</p>
Full article ">
14 pages, 2899 KiB  
Review
YAP/TAZ Signaling in the Pathobiology of Pulmonary Fibrosis
by Kostas A. Papavassiliou, Amalia A. Sofianidi, Fotios G. Spiliopoulos, Vassiliki A. Gogou, Antonios N. Gargalionis and Athanasios G. Papavassiliou
Cells 2024, 13(18), 1519; https://doi.org/10.3390/cells13181519 - 10 Sep 2024
Viewed by 205
Abstract
Pulmonary fibrosis (PF) is a severe, irreversible lung disease characterized by progressive scarring, with idiopathic pulmonary fibrosis (IPF) being the most prevalent form. IPF’s pathogenesis involves repetitive lung epithelial injury leading to fibroblast activation and excessive extracellular matrix (ECM) deposition. The prognosis for [...] Read more.
Pulmonary fibrosis (PF) is a severe, irreversible lung disease characterized by progressive scarring, with idiopathic pulmonary fibrosis (IPF) being the most prevalent form. IPF’s pathogenesis involves repetitive lung epithelial injury leading to fibroblast activation and excessive extracellular matrix (ECM) deposition. The prognosis for IPF is poor, with limited therapeutic options like nintedanib and pirfenidone offering only modest benefits. Emerging research highlights the dysregulation of the yes-associated protein (YAP)/transcriptional coactivator with PDZ-binding motif (TAZ) signaling pathway as a critical factor in PF. YAP and TAZ, components of the Hippo pathway, play significant roles in cell proliferation, differentiation, and fibrosis by modulating gene expression through interactions with TEA domain (TEAD) transcription factors. The aberrant activation of YAP/TAZ in lung tissue promotes fibroblast activation and ECM accumulation. Targeting the YAP/TAZ pathway offers a promising therapeutic avenue. Preclinical studies have identified potential treatments, such as trigonelline, dopamine receptor D1 (DRD1) agonists, and statins, which inhibit YAP/TAZ activity and demonstrate antifibrotic effects. These findings underscore the importance of YAP/TAZ in PF pathogenesis and the potential of novel therapies aimed at this pathway, suggesting a new direction for improving IPF treatment outcomes. Further research is needed to validate these approaches and translate them into clinical practice. Full article
(This article belongs to the Special Issue Cellular Signaling and Therapeutic Approaches of Pulmonary Fibrosis)
Show Figures

Figure 1

Figure 1
<p>YAP/TAZ signaling pathway in the pathogenesis of PF. YAP and TAZ are highly expressed in fibrotic lung tissue, with TAZ showing prominent nuclear expression in spindle-shaped fibroblastic cells. At a molecular level, TBK1 and SPHK1 promote the fibrotic effects of YAP/TAZ, which are regulated through interactions with PAI-1 and Twist1. TGF-β signaling augments the expression of PAI-1. (<b>A</b>) The role of YAP/TAZ in the TGF-β cascade. GPCR ligands, such as LPA, S1P, and thrombin, facilitate the accumulation of YAP in the nucleus through the mediation of Rho. TGF-β activates Smad2/3/4 complexes, leading to their translocation into the nucleus, where they utilize YAP as a coactivator to drive the transcription of fibrogenic YAP/Smad target genes. (<b>B</b>) In the pathogenesis of IPF, YAP/TAZ interacts with several other signaling pathways, such as EGFR, Wnt, and Notch, to promote lung fibroblast activation and pulmonary fibrosis. This figure was created based on the tools provided by Biorender.com (<a href="https://biorender.com/" target="_blank">https://biorender.com/</a>; accessed 5 August 2024). EGFR: epidermal growth factor receptor, GPCR: G protein-coupled receptor, IPF: idiopathic pulmonary fibrosis, LPA: lysophosphatidic acid, PAI-1: plasminogen activator inhibitor-1, Rho: ras homolog family member, S1P: sphingosine-1-phosphate, Smad: mothers against decapentaplegic homolog, SPHK1: sphingosine kinase 1, TAZ: transcriptional coactivator with PDZ-binding motif, TBK1: TANK binding kinase 1, TGF-β: transforming growth factor beta, Twist1: Twist Family BHLH transcription factor 1, Wnt: wingless-related integration site, YAP: yes-associated protein.</p>
Full article ">Figure 2
<p>The role of YAP/TAZ in regulating glycolytic reprogramming in lung fibrosis. YAP/TAZ enhances the expression of glycolytic enzymes, boosting glucose metabolism and lactate production. Integrins, through signaling pathways like PI3K/Akt and MAPK/ERK, further regulate glycolysis, impacting glucose utilization and energy metabolism. Integrins/FAK regulate glycolysis via the YAP/TAZ axis. This axis facilitates the nuclear translocation of YAP/TAZ, enhancing their fibrotic effects. Additionally, HK2, a key glycolytic enzyme, is crucial for YAP/TAZ nuclear translocation. The activation of lung fibroblasts is often marked by upregulation of glycolytic enzymes, such as PFK1, which interacts with YAP/TAZ coactivator TEADs, promoting fibroblast differentiation. This figure was created based on the tools provided by Biorender.com (<a href="https://biorender.com/" target="_blank">https://biorender.com/</a>; accessed 5 August 2024). Akt: protein kinase B, ERK: extracellular signal-regulated kinase, FAK: focal adhesion kinase, HK2: hexokinase 2, MAPK: mitogen-activated protein kinase, PI3K: phosphoinositide 3-kinase, PFK1: phosphofructokinase-1, TEADs: TEA domain transcription factors, YAP: yes-associated protein, TAZ: transcriptional coactivator with PDZ-binding motif.</p>
Full article ">
13 pages, 5475 KiB  
Article
Taxonomic Exploration of Rare Amphipods: A New Genus and Two New Species (Amphipoda, Iphimedioidea, Laphystiopsidae) Described from Seamounts in the Western Pacific
by Yanrong Wang, Zhongli Sha and Xianqiu Ren
Diversity 2024, 16(9), 564; https://doi.org/10.3390/d16090564 - 10 Sep 2024
Viewed by 199
Abstract
During two expeditions to the seamounts in the Yap-Caroline area of the Western Pacific, a new genus, Phoxirostus gen. nov., in the family Laphystiopsidae Stebbing, 1899, is erected for two new species, P. longicarpus sp. nov. (type species) and P. yapensis sp. nov. [...] Read more.
During two expeditions to the seamounts in the Yap-Caroline area of the Western Pacific, a new genus, Phoxirostus gen. nov., in the family Laphystiopsidae Stebbing, 1899, is erected for two new species, P. longicarpus sp. nov. (type species) and P. yapensis sp. nov. The new genus can be distinguished from the other three laphystiopsid genera by the acute rostrum not overreaching the distal end of the first peduncular article of antenna 1, the outer plate of maxilla 1 bearing 10–11 spines, and the elongated carpus of pereopods 3–7 being distinctly longer than half the length of the propodus. Phoxirostus longicarpus sp. nov. differs from P. yapensis sp. nov. by the shape of the eyes and coxa 4, the presence of posterodistal protrusions on pleonite 1, and the number of posterodistal protrusions on pleonite 2. Generic analysis of one mitochondrial (COI) and one nuclear (H3) gene using maximum likelihood and Bayesian inference clarified the phylogenetic position of the Laphystiopsidae within the superfamily Iphimedioidea Boeck, 1871. Full article
(This article belongs to the Special Issue Diversity and Evolution within the Amphipoda)
Show Figures

Figure 1

Figure 1
<p><span class="html-italic">Phoxirostus longicarpus</span> <b>sp. nov.</b>, MBM 286818, holotype, female (6.0 mm): showing that it is associated with the sponge and photographed after being fixed in 95% ethanol.</p>
Full article ">Figure 2
<p><span class="html-italic">Phoxirostus longicarpus</span> <b>sp. nov.</b>, MBM 286818, holotype, female (6.0 mm): A1, antenna 1; A2, antenna 2; G1 L, left gnathopod 1; G2 L, left gnathopod 2; H, head, U1 R, right uropod1; U2 L, left uropod 2; U3 R, right uropod 3, and the arrow points to the ventral view of inner ramus; T, telson.</p>
Full article ">Figure 3
<p><span class="html-italic">Phoxirostus longicarpus</span> <b>sp. nov.</b>, MBM 286818, holotype, female (6.0 mm): UL, upper lip; LL, lower lip; Md L, left mandible; Mx1, maxilla 1; Mx2, maxilla 2; Mxp, maxilliped.</p>
Full article ">Figure 4
<p><span class="html-italic">Phoxirostus longicarpus</span> <b>sp. nov.</b>, MBM 286818, holotype, female (6.0 mm): P3 L, left pereopod 3; P4 L, left pereopod 4; P5 L, left pereopod 5; P6 R, right pereopod 6; P7 R, right pereopod 7.</p>
Full article ">Figure 5
<p><span class="html-italic">Phoxirostus yapensis</span> <b>sp. nov.</b>, MBM 286617, holotype, female (7.1 mm): photographed immediately after being collected by Wei Jiang.</p>
Full article ">Figure 6
<p><span class="html-italic">Phoxirostus yapensis</span> <b>sp. nov.</b>, MBM 286617, holotype, female (7.1 mm): UL, upper lip; LL, lower lip; Md L, left mandible, and the arrow points to details of two distal articles of palp; Md R, only shows the incisor and accessory spines; Mx1 R, right maxilla 1; Mx2, maxilla 2; Mxp, maxilliped; A1, antenna 1; A2, antenna 2.</p>
Full article ">Figure 7
<p><span class="html-italic">Phoxirostus yapensis</span> <b>sp. nov.</b>, MBM 286617, holotype, female (7.1 mm): G1 R, right gnathopod 1; G2 R, right gnathopod 2; P4 R, right pereopod 4; P5 R, right pereopod 5; P6 R, right pereopod 6; P7 R, right pereopod 7; H, head, arrow points acute rostrum; T, telson.</p>
Full article ">Figure 8
<p><span class="html-italic">Phoxirostus yapensis</span> <b>sp. nov.</b>, MBM 286617, paratype, male (5.3 mm): G1 R, right gnathopod 1; G2 R, right gnathopod 2; P3 L, left pereopod 3; P4 L, left pereopod 4; P5 L, left pereopod 5; P6 L, left pereopod 6; P7 L, left pereopod 7; E1–3, epimeron plates 1–3.</p>
Full article ">Figure 9
<p>Phylogenetic tree of the superfamily Iphimedioidea Boeck, 1871, taxa resolved based on the combined dataset of four genes (COI and H3): (<b>A</b>) Bayesian inference (BI) tree; (<b>B</b>) maximum likelihood tree.</p>
Full article ">Figure 10
<p>The distribution of laphystiopsid species: the location of the sampling site of two new species (red and yellow rhombus); distribution of <span class="html-italic">Laphystiopsis</span> species (green, pink, red, yellow, and blue square); distribution of <span class="html-italic">Prolaphystius</span> species (pink triangle) and Prolaphystiopsis species (red and yellow circle).</p>
Full article ">
19 pages, 6046 KiB  
Article
Activation of Yes-Associated Protein Is Indispensable for Transformation of Kidney Fibroblasts into Myofibroblasts during Repeated Administration of Cisplatin
by Jia-Bin Yu, Babu J. Padanilam and Jinu Kim
Cells 2024, 13(17), 1475; https://doi.org/10.3390/cells13171475 - 2 Sep 2024
Viewed by 426
Abstract
Cisplatin is a potent chemotherapy medication that is used to treat various types of cancer. However, it can cause nephrotoxic side effects, which lead to acute kidney injury (AKI) and subsequent chronic kidney disease (CKD). Although a clinically relevant in vitro model of [...] Read more.
Cisplatin is a potent chemotherapy medication that is used to treat various types of cancer. However, it can cause nephrotoxic side effects, which lead to acute kidney injury (AKI) and subsequent chronic kidney disease (CKD). Although a clinically relevant in vitro model of CKD induced by repeated administration of low-dose cisplatin (RAC) has been established, its underlying mechanisms remain poorly understood. Here, we compared single administration of high-dose cisplatin (SAC) to repeated administration of low-dose cisplatin (RAC) in myofibroblast transformation and cellular morphology in a normal rat kidney fibroblast NRK-49F cell line. RAC instead of SAC transformed the fibroblasts into myofibroblasts as determined by α-smooth muscle actin, enlarged cell size as represented by F-actin staining, and increased cell flattening as expressed by the semidiameter ratio of attached cells to floated cells. Those phenomena, as well as cellular senescence, were significantly detected from the time right before the second administration of cisplatin. Interestingly, inhibition of the interaction between Yes-associated protein (YAP) and the transcriptional enhanced associated domain (TEAD) using Verteporfin remarkedly reduced cell size, cellular senescence, and myofibroblast transformation during RAC. These findings collectively suggest that YAP activation is indispensable for cellular hypertrophy, senescence, and myofibroblast transformation during RAC in kidney fibroblasts. Full article
Show Figures

Figure 1

Figure 1
<p>RAC transforms kidney fibroblasts into myofibroblasts, but SAC does not. (<b>A</b>) Experimental timetable for SAC and RAC. (<b>B</b>) Cellular viability after SAC and RAC in a dose-dependent manner (<span class="html-italic">n</span> = 4 experiments, triplicate wells per experiment). <span class="html-italic">F</span><sub>5,18</sub> = 96.877, <span class="html-italic">p</span> &lt; 0.001 and <span class="html-italic">F</span><sub>5,18</sub> = 389.137, <span class="html-italic">p</span> &lt; 0.001 for 1-way ANOVA, respectively. (<b>C</b>) Illustrative Western blots depicting the expressions of α-SMA, vimentin, and fibronectin are presented. The anti-β-actin antibody served as a control for loading. (<b>D</b>–<b>F</b>) Quantification of α-SMA, vimentin, and fibronectin expressions (<span class="html-italic">n</span> = 4 experiments). <span class="html-italic">F</span><sub>2,9</sub> = 13.026, <span class="html-italic">p</span> = 0.002 and <span class="html-italic">F</span><sub>2,9</sub> = 18.540, <span class="html-italic">p</span> &lt; 0.001 for 1-way ANOVA on α-SMA expression during SAC and RAC, respectively. <span class="html-italic">F</span><sub>2,9</sub> = 26.880, <span class="html-italic">p</span> &lt; 0.001 and <span class="html-italic">F</span><sub>2,9</sub> = 18.560, <span class="html-italic">p</span> &lt; 0.001 for 1-way ANOVA on vimentin expression during SAC and RAC, respectively. <span class="html-italic">F</span><sub>2,9</sub> = 53.583, <span class="html-italic">p</span> &lt; 0.001 and <span class="html-italic">F</span><sub>2,9</sub> = 27.221, <span class="html-italic">p</span> &lt; 0.001 for 1-way ANOVA on fibronectin expression during SAC and RAC, respectively. * <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 versus 0 μM.</p>
Full article ">Figure 2
<p>RAC induces cellular hypertrophy and flattening of kidney fibroblasts, but SAC does not. (<b>A</b>) Representative images of F-actin- and DAPI-stained cells. Scale bar = 50 μm. (<b>B</b>,<b>C</b>) Quantifications of cellular and nuclear attachment area from F-actin and DAPI staining, respectively (<span class="html-italic">n</span> = 4 experiments). <span class="html-italic">F</span><sub>2,9</sub> = 28.761, <span class="html-italic">p</span> &lt; 0.001 and <span class="html-italic">F</span><sub>2,9</sub> = 47.085, <span class="html-italic">p</span> &lt; 0.001 for 1-way ANOVA on cellular attachment areas during SAC and RAC, respectively. <span class="html-italic">F</span><sub>2,9</sub> = 15.796, <span class="html-italic">p</span> = 0.001 and <span class="html-italic">F</span><sub>2,9</sub> = 928.345, <span class="html-italic">p</span> &lt; 0.001 for 1-way ANOVA on nuclear attachment areas during SAC and RAC, respectively. (<b>D</b>,<b>E</b>) Cellular and nuclear volumes were defined based on mean values of FSC-A and PE-A, respectively (<span class="html-italic">n</span> = 4 experiments, 10,000 events per experiment). <span class="html-italic">H</span> = 3.512, <span class="html-italic">N</span><sub>3</sub> = 4, <span class="html-italic">p</span> = 0.197 for Kruskal-Wallis <span class="html-italic">H</span> test on cellular volume during SAC. <span class="html-italic">F</span><sub>2,9</sub> = 23.179, <span class="html-italic">p</span> &lt; 0.001 for 1-way ANOVA on cellular volume during RAC. <span class="html-italic">F</span><sub>2,9</sub> = 22.314, <span class="html-italic">p</span> &lt; 0.001 and <span class="html-italic">F</span><sub>2,9</sub> = 19.585, <span class="html-italic">p</span> &lt; 0.001 for 1-way ANOVA on nuclear volumes during SAC and RAC, respectively. (<b>F</b>,<b>G</b>) Cellular and nuclear flattening ratios were defined using relative semidiameters of attached and floated cells (<span class="html-italic">n</span> = 16, 4 times of the experiment with attached cells × 4 times of the experiment with floated cells). <span class="html-italic">H</span> = 27.467, <span class="html-italic">N</span><sub>3</sub> = 16, <span class="html-italic">p</span> ≤ 0.001 for the Kruskal–Wallis <span class="html-italic">H</span> test on the flattening ratio of cells during SAC. <span class="html-italic">F</span><sub>2,45</sub> = 406.263, <span class="html-italic">p</span> &lt; 0.001 for 1-way ANOVA on the flattening ratio of cells during RAC. <span class="html-italic">F</span><sub>2,45</sub> = 16.427, <span class="html-italic">p</span> &lt; 0.001 and <span class="html-italic">F</span><sub>2,45</sub> = 2881.397, <span class="html-italic">p</span> &lt; 0.001 for 1-way ANOVA on the flattening ratios of the nucleus during SAC and RAC, respectively. * <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 versus 0 μM.</p>
Full article ">Figure 3
<p>RAC induces cellular senescence in kidney fibroblasts, whereas SAC does not. (<b>A</b>) Representative images of SA-β-gal-stained cells. Scale bar = 50 μm. (<b>B</b>) Western blots showing p21 expression, with β-actin serving as the loading control. (<b>C</b>) Quantification of p21 expression (<span class="html-italic">n</span> = 4 experiments). One-way ANOVA showed <span class="html-italic">F</span><sub>2,9</sub> = 9.411, <span class="html-italic">p</span> = 0.006 for SAC, and <span class="html-italic">F</span><sub>2,9</sub> = 14.512, <span class="html-italic">p</span> = 0.002 for RAC. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 versus 0 μM. (<b>D</b>) Mean values of FSC-A and α-SMA expression were used to assess cellular volumes and myofibroblast transformation in RAC-exposed NRK-49F cells, respectively (10,000 events per experiment).</p>
Full article ">Figure 4
<p>Cellular hypertrophy and flattening induced by RAC in a time-dependent manner. NRK-49F cells were repeatedly treated with 10 μM cisplatin for 6 h and normal culture media for 18 h. (<b>A</b>) Cellular viability during RAC (<span class="html-italic">n</span> = 3 experiments, triplicate wells per experiment). <span class="html-italic">F</span><sub>7,16</sub> = 11.698, <span class="html-italic">p</span> &lt; 0.001 for 1-way ANOVA. (<b>B</b>) Representative images of F-actin- and DAPI-stained cells. Scale bar, 50 μm. (<b>C</b>,<b>D</b>) Quantifications of cellular and nuclear attachment surface area from F-actin and DAPI staining, respectively (<span class="html-italic">n</span> = 4 experiments). <span class="html-italic">F</span><sub>4,15</sub> = 147.484, <span class="html-italic">p</span> &lt; 0.001 and <span class="html-italic">F</span><sub>4,15</sub> = 76.959, <span class="html-italic">p</span> &lt; 0.001 for 1-way ANOVA, respectively. (<b>E</b>,<b>F</b>) Cellular and nuclear volume were defined using mean values of FSC-A and PE-A, respectively (<span class="html-italic">n</span> = 4 experiments, 10,000 events per experiment). <span class="html-italic">F</span><sub>4,15</sub> = 171.361, <span class="html-italic">p</span> &lt; 0.001 and <span class="html-italic">F</span><sub>4,15</sub> = 72.237, <span class="html-italic">p</span> &lt; 0.001 for 1-way ANOVA, respectively. (<b>G</b>,<b>H</b>) Cellular and nuclear flattening ratios were defined using relative semidiameters of attached and floated cells (<span class="html-italic">n</span> = 4 experiments). <span class="html-italic">F</span><sub>4,75</sub> = 506.240, <span class="html-italic">p</span> &lt; 0.001 and <span class="html-italic">F</span><sub>4,75</sub> = 76.545, <span class="html-italic">p</span> &lt; 0.001 for 1-way ANOVA, respectively. * <span class="html-italic">p</span> &lt; 0.1, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 versus 0 h; †† <span class="html-italic">p</span> &lt; 0.01, ††† <span class="html-italic">p</span> &lt; 0.001 versus 6 h; § <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 versus 24 h; ‡‡ <span class="html-italic">p</span> &lt; 0.01, ‡‡‡ <span class="html-italic">p</span> &lt; 0.001 versus 30 h.</p>
Full article ">Figure 5
<p>Cellular senescence and transformation of kidney fibroblasts into myofibroblasts induced by RAC in a time-dependent manner. NRK-49F cells were repeatedly treated with 10 μM cisplatin for 6 h and normal culture media for 18 h (<span class="html-italic">n</span> = 4 experiments). (<b>A</b>) Representative images of SA-β-gal-stained cells. Scale bar, 50 μm. (<b>B</b>) The percentage of SA-β-gal-positive cells. <span class="html-italic">F</span><sub>4,15</sub> = 144.864, <span class="html-italic">p</span> &lt; 0.001 for 1-way ANOVA. (<b>C</b>) Representative images of α-SMA-stained cells. DNA was counterstained with DAPI. Scale bar, 50 μm. (<b>D</b>) The percentage of α-SMA-positive cells. <span class="html-italic">F</span><sub>4,15</sub> = 309.888, <span class="html-italic">p</span> &lt; 0.001 for 1-way ANOVA. * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001 versus 0 h; † <span class="html-italic">p</span> &lt; 0.05, ††† <span class="html-italic">p</span> &lt; 0.001 versus 6 h; §§§ <span class="html-italic">p</span> &lt; 0.001 versus 24 h; ‡‡‡ <span class="html-italic">p</span> &lt; 0.001 versus 30 h.</p>
Full article ">Figure 6
<p>Cisplatin-induced JNK activation contributes to cellular hypertrophy but is not linked to cellular senescence and transformation into myofibroblasts in kidney fibroblast cells. (<b>A</b>,<b>B</b>) NRK-49F cells were treated with 10 μM cisplatin for 6 h (<span class="html-italic">n</span> = 4 experiments). (<b>A</b>) Representative Western blots of p-JNK and t-JNK expression. Anti-β-actin antibody was used as a loading control. (<b>B</b>) Quantifications of p-JNK expression. <span class="html-italic">t</span><sub>6</sub> = –6.185 for 2-tailed unpaired Student’s <span class="html-italic">t</span>-test. (<b>C</b>–<b>J</b>) NRK-49F cells were treated with 10 μM cisplatin plus 3 μM SP600125 (SP) as a pan-JNK inhibitor in a 0.03% DMSO vehicle (veh) for 6 h, changed into normal culture media, and incubated for 18 h. This cycle was performed once (<b>C</b>–<b>E</b>) or repeated 2 times (<b>F</b>–<b>J</b>). (<b>C</b>) Representative images of F-actin- and DAPI-stained cells. Scale bar, 50 μm. (<b>D</b>,<b>E</b>) Quantifications of cellular and nuclear attachment surface area from F-actin and DAPI staining, respectively (<span class="html-italic">n</span> = 4 experiments). <span class="html-italic">t</span><sub>3.432</sub> = 10.289 for 2-tailed unpaired Welch’s <span class="html-italic">t</span>-test on the cellular attachment area. <span class="html-italic">t</span><sub>6</sub> = 6.683 for 2-tailed unpaired Student’s <span class="html-italic">t</span>-test on the nuclear attachment area. (<b>F</b>) Representative images of SA-β-gal-stained cells. Scale bar, 50 μm. (<b>G</b>) The percentage of SA-β-gal-positive cells (<span class="html-italic">n</span> = 4 experiments). <span class="html-italic">t</span><sub>6</sub> = –0.631 for 2-tailed unpaired Student’s <span class="html-italic">t</span>-test. (<b>H</b>) Representative Western blots of α-SMA and fibronectin expression. Anti-β-actin antibody was used as a loading control. (<b>I</b>,<b>J</b>) Quantifications of α-SMA and fibronectin expression (<span class="html-italic">n</span> = 6 experiments). <span class="html-italic">t</span><sub>6</sub> = –0.122 and 1.948 for 2-tailed unpaired Student’s <span class="html-italic">t</span>-test on α-SMA and fibronectin, respectively. *** <span class="html-italic">p</span> &lt; 0.001 versus 0 h; ### <span class="html-italic">p</span> &lt; 0.001 versus vehicle.</p>
Full article ">Figure 7
<p>Cisplatin induces Hippo-independent YAP inactivation in kidney fibroblast cells. (<b>A</b>–<b>E</b>) NRK-49F cells were treated with 10 μM cisplatin for 0, 1, 3, or 6 h. (<b>A</b>) Representative images of YAP-stained cells. DNA was counterstained with DAPI. Scale bar, 50 μm. (<b>B</b>) Representative Western blots of p-YAP, t-YAP, p-MOB1, and p-MST1/2 expression. Anti-β-actin antibody was used as a loading control. (<b>C</b>–<b>E</b>) Quantifications of YAP activation (ratio of t-YAP to p-YAP), MOB1 phosphorylation (p-MOB1), and MST1/2 phosphorylation (p-MST1/2) (<span class="html-italic">n</span> = 4 experiments in (<b>C</b>,<b>E</b>); <span class="html-italic">n</span> = 6 experiments in (<b>D</b>). <span class="html-italic">F</span><sub>3,12</sub> = 104.106, <span class="html-italic">p</span> &lt; 0.001 and <span class="html-italic">F</span><sub>3,12</sub> = 31.203, <span class="html-italic">p</span> &lt; 0.001 for 1-way ANOVA on YAP activation and MST1/2 phosphorylation, respectively. <span class="html-italic">H</span> = 13.467, <span class="html-italic">N</span><sub>4</sub> = 6, <span class="html-italic">p</span> &lt; 0.05 for the Kruskal–Wallis <span class="html-italic">H</span> test on MOB1 phosphorylation. (<b>F</b>–<b>K</b>) NRK-49F cells were treated with 3 μM Verteporfin (VP) as a YAP inhibitor in a 0.1% DMSO vehicle (veh) for 24 h (<span class="html-italic">n</span> = 4 experiments). (<b>F</b>) Representative images of F-actin- and DAPI-stained cells. Scale bar, 50 μm. (<b>G</b>,<b>H</b>) Quantifications of cellular and nuclear attachment surface areas from F-actin and DAPI staining, respectively. <span class="html-italic">t</span><sub>6</sub> = 0.641 and –1.342 for 2-tailed unpaired Student’s <span class="html-italic">t</span>-test, respectively. (<b>I</b>) Representative Western blots of α-SMA and fibronectin expression. Anti-β-actin antibody was used as a loading control. (<b>J</b>,<b>K</b>) Quantifications of α-SMA and fibronectin expression. <span class="html-italic">t</span><sub>6</sub> = 4.369 and 15.560 for 2-tailed unpaired Student’s <span class="html-italic">t</span>-test, respectively. * <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 versus 0 h; ## <span class="html-italic">p</span> &lt; 0.01, ### <span class="html-italic">p</span> &lt; 0.001 versus vehicle.</p>
Full article ">Figure 8
<p>YAP inhibition attenuates cellular hypertrophy, senescence, and transformation into myofibroblasts induced by RAC in kidney fibroblast cells. NRK-49F cells were treated with 10 μM cisplatin plus 3 μM Verteporfin (VP) as a YAP inhibitor in a 0.1% DMSO vehicle (veh) for 6 h, changed into normal culture media, and incubated for 18 h. The treatment was repeated 2 times (<span class="html-italic">n</span> = 4 experiments). (<b>A</b>) Representative images of F-actin- and DAPI-stained cells. Scale bar, 50 μm. (<b>B</b>,<b>C</b>) Quantification of cellular and nuclear attachment surface areas from F-actin and DAPI staining, respectively. <span class="html-italic">t</span><sub>6</sub> = 13.857 and 9.954 for 2-tailed unpaired Student’s <span class="html-italic">t</span>-test, respectively. (<b>D</b>) Representative images of SA-β-gal-stained cells. Scale bar, 50 μm. (<b>E</b>) The percentage of SA-β-gal-positive cells. <span class="html-italic">t</span><sub>6</sub> = 11.273 for 2-tailed unpaired Student’s <span class="html-italic">t</span>-test. (<b>F</b>) Representative Western blot of α-SMA and fibronectin expression. Anti-β-actin antibody was used as a loading control. (<b>G</b>–<b>K</b>) Quantification of α-SMA, vimentin, fibronectin, CTGF, and p21 expression. Two-tailed unpaired Student’s <span class="html-italic">t</span>-test showed <span class="html-italic">t</span><sub>6</sub> = 3.545 (<b>G</b>), <span class="html-italic">t</span><sub>6</sub> = 2.822 (<b>H</b>), and <span class="html-italic">t</span><sub>6</sub> = 3.140 (<b>I</b>). Two-tailed unpaired Welch’s <span class="html-italic">t</span>-test showed <span class="html-italic">t</span><sub>6</sub> = 2.714 (<b>J</b>) and <span class="html-italic">t</span><sub>6</sub> = 5.187 (<b>K</b>). # <span class="html-italic">p</span> &lt; 0.05, ### <span class="html-italic">p</span> &lt; 0.001 versus vehicle.</p>
Full article ">
15 pages, 1924 KiB  
Review
Regulation of Tumor Microenvironment through YAP/TAZ under Tumor Hypoxia
by Sung Hoon Choi and Do Young Kim
Cancers 2024, 16(17), 3030; https://doi.org/10.3390/cancers16173030 - 30 Aug 2024
Viewed by 360
Abstract
In solid tumors such as hepatocellular carcinoma (HCC), hypoxia is one of the important mechanisms of cancer development that closely influences cancer development, survival, and metastasis. The development of treatments for cancer was temporarily revolutionized by immunotherapy but continues to be constrained by [...] Read more.
In solid tumors such as hepatocellular carcinoma (HCC), hypoxia is one of the important mechanisms of cancer development that closely influences cancer development, survival, and metastasis. The development of treatments for cancer was temporarily revolutionized by immunotherapy but continues to be constrained by limited response rates and the resistance and high costs required for the development of new and innovative strategies. In particular, solid tumors, including HCC, a multi-vascular tumor type, are sensitive to hypoxia and generate many blood vessels for metastasis and development, making it difficult to treat HCC, not only with immunotherapy but also with drugs targeting blood vessels. Therefore, in order to develop a treatment strategy for hypoxic tumors, various mechanisms must be explored and analyzed to treat these impregnable solid tumors. To date, tumor growth mechanisms linked to hypoxia are known to be complex and coexist with various signal pathways, but recently, mechanisms related to the Hippo signal pathway are emerging. Interestingly, Hippo YAP/TAZ, which appear during early tumor and normal tumor growth, and YAP/TAZ, which appear during hypoxia, help tumor growth and proliferation in different directions. Peculiarly, YAP/TAZ, which have different phosphorylation directions in the hypoxic environment of tumors, are involved in cancer proliferation and metastasis in various carcinomas, including HCC. Analyzing the mechanisms that regulate the function and expression of YAP in addition to HIF in the complex hypoxic environment of tumors may lead to a variety of anti-cancer strategies and combining HIF and YAP/TAZ may develop the potential to change the landscape of cancer treatment. Full article
(This article belongs to the Section Tumor Microenvironment)
Show Figures

Figure 1

Figure 1
<p>YAP/TAZ mechanisms in solid tumors, particularly HCCs, in hypoxic environments. In early tumor development, YAP/TAZ act as important cell signals, not only for tumor apoptosis, but also for tumor proliferation. However, when tumor growth develops excessively and a hypoxic environment is created, YAP separates from TAZ and enters blood vessels through HIF1a. By generating and activating EMT, it maintains the microenvironment around the tumor so that the tumor can develop better, and it also develops an environment for the tumor to metastasize to other organs or tissues. In early-stage tumors, YAP/TAZ bind to TEAD and act as a transcriptional activator through the Hippo off signaling pathway, following the G protein-coupled receptor (GPCR) or glucagon receptor (Gcgr) signaling system to induce the development of early-stage tumors. When a tumor grows beyond a certain size and is unable to obtain the oxygen and nutrients it needs to proliferate, the angiogenic switch is triggered and through the hypoxia signaling system, overexpressed HIF and YAP/TAZ combine with TEAD to activate the expression of various genes to overcome hypoxia. Specifically, when localized to the nucleus, YAP is recruited together with hypoxia-inducible factor 1α (HIF-1α) for PKM2 transcription at the pyruvate kinase M2 (<span class="html-italic">PKM2</span>) gene promoter, contributing to tumor formation and metastasis.</p>
Full article ">Figure 2
<p>Tumor development through YAP/TAZ under hypoxic environment in various cancers. Changes in YAP/TAZ under hypoxia conditions are not unique to HCC. In various cancers, tumors develop through mechanisms that increase tumor proliferation and evade tumor cell death through YAP/TAZ. When a certain size is reached, the hypoxic stimulation and angiogenesis switch are activated in the tumor, and changes occur in the tumor microenvironment. YAP/TAZ also avoid cell death due to hypoxia through different phosphorylation mechanisms, leading to the development of metastasis and proliferation through increased EMT and resistance to chemotherapy. This is one of the main mechanisms of development in intractable tumors such as liver cancer, breast cancer, lung cancer, colon cancer, and ovarian cancer.</p>
Full article ">Figure 3
<p>Phosphorylation of YAP/TAZ under hypoxic conditions. Under normoxia environments, phosphorylation of YAP/TAZ increases the expression of genes related to tumor growth through the combination of transcription activators; YAP/TAZ bind to TEAD and acts as a transcriptional activator through the Hippo-Off signaling pathway, mediating the translocation of the YAP/TEZ complex into the nucleus by inducing phosphorylation of Rho, MST1/2, and LAST1/2, in sequence, depending on the GPCR or GCGR signaling system. Once translocated into the nucleus, YAP/TAZ bind to TEAD and activate cell proliferation and the cell cycle, and changes in cellular structure, including hypertrophy of heterogeneous cells, leading to tumor development. However, under hypoxic environments, YAP and TAZ bind to HIF differently through a variant phosphorylation process. Hypoxia-mediated HIF-1a translocates from the cytosol to the nucleus, where it binds to unphosphorylated YAP and binds to TEAD. YAP/HIF binds to DNA, inhibits DNA damage and apoptosis, and induces the expression of a variety of genes involved in cell proliferation, angiogenesis, and metastasis. In addition, interaction of HIF-1α with TAZ also stimulates TAZ/TEAD transcriptional activity. TAZ and HIF-1α interact and function as mutual transcriptional cofactors. HIF-1α acts as a cofactor of the TAZ/TEADs complex for the transcription of target genes, and TAZ acts as a cofactor of HIF-1α for the transcription of target genes such as PAI1, BIRC5, CTGF, PDK1, and LDHA in the hypoxic tumor microenvironment. Moreover, TAZ also regulates the mechanism by which HIF binds to and regulates TAZ expression. These different hypoxic mechanisms of YAP/TAZ affect tumor formation and patient mortality and are involved in sensitivity to anticancer drugs and development of tumor metastasis.</p>
Full article ">
18 pages, 4133 KiB  
Article
Context-Dependent Distinct Roles of SOX9 in Combined Hepatocellular Carcinoma–Cholangiocarcinoma
by Yoojeong Park, Shikai Hu, Minwook Kim, Michael Oertel, Aatur Singhi, Satdarshan P. Monga, Silvia Liu and Sungjin Ko
Cells 2024, 13(17), 1451; https://doi.org/10.3390/cells13171451 - 29 Aug 2024
Viewed by 423
Abstract
Combined hepatocellular carcinoma–cholangiocarcinoma (cHCC-CCA) is a challenging primary liver cancer subtype with limited treatment options and a devastating prognosis. Recent studies have underscored the context-dependent roles of SOX9 in liver cancer formation in a preventive manner. Here, we revealed that liver-specific developmental Sox9 [...] Read more.
Combined hepatocellular carcinoma–cholangiocarcinoma (cHCC-CCA) is a challenging primary liver cancer subtype with limited treatment options and a devastating prognosis. Recent studies have underscored the context-dependent roles of SOX9 in liver cancer formation in a preventive manner. Here, we revealed that liver-specific developmental Sox9 elimination using Alb-Cre;Sox9(flox/flox) (LKO) and CRISPR/Cas9-based tumor-specific acute Sox9 elimination (CKO) in SB-HDTVI-based Akt-YAP1 (AY) and Akt-NRAS (AN) cHCC-CCA models showed contrasting responses. LKO abrogates the AY CCA region while stimulating poorly differentiated HCC proliferation, whereas CKO prevents AY and AN cHCC-CCA development irrespective of tumor cell fate. Additionally, AN, but not AY, tumor formation partially depends on the Sox9-Dnmt1 cascade. SOX9 is dispensable for AY-mediated, HC-derived, LPC-like immature CCA formation but is required for their maintenance and transformation into mature CCA. Therapeutic Sox9 elimination using the OPN-CreERT2 strain combined with inducible Sox9 iKO specifically reduces AY but not AN cHCC-CCA tumors. This necessitates the careful consideration of genetic liver cancer studies using developmental Cre and somatic mutants, particularly for genes involved in liver development. Our findings suggest that SOX9 elimination may hold promise as a therapeutic approach for a subset of cHCC-CCA and highlight the need for further investigation to translate these preclinical insights into personalized clinical applications. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Chronic developmental deletion of <span class="html-italic">Sox9</span> switches the fate of <span class="html-italic">Akt-YAP1</span>-driven cHCC-CCA to aggressive HCC at the expense of CCA. (<b>A</b>) Experimental design illustrating plasmids used for HDTVI, mice used in study, and time-points analyzed. (<b>B</b>) Representative gross images from <span class="html-italic">Akt-YAP1</span>-injected <span class="html-italic">Sox9</span>-floxed mice (LWT) and <span class="html-italic">Akt-YAP1</span>-injected <span class="html-italic">Alb-Cre;Sox9<sup>(f/f)</sup></span> liver-specific <span class="html-italic">Sox9</span> knockout or <span class="html-italic">Sox9</span> LKO mice showing tumor-laden enlarged livers in both cases. (<b>C</b>) LW/BW ratio depicts a significantly lower tumor burden in <span class="html-italic">Akt-YAP1 Sox9</span> LKO as compared to LWT at 5 weeks but not earlier than the 2-week time-point. (<b>D</b>) Kaplan–Meier survival curve showing a significant decrease in the survival of <span class="html-italic">Akt-YAP1</span> mice that were <span class="html-italic">Sox9</span> LKO as compared to LWT. (<b>E</b>) Representative tiled image of tumor-bearing livers at 5 weeks in <span class="html-italic">Akt-YAP1</span> LWT stained for CK19 IHC showing the CCA component of the positive cHCC-CCA staining. <span class="html-italic">Sox9 LKO</span> livers were full of circumscribed tumors that were negative for CK19 at the same time-point. (<b>F</b>) Quantification of CK19 IHC verifies significantly reduced staining in <span class="html-italic">Akt-YAP1 Sox9</span> LKO as compared to <span class="html-italic">Akt-YAP1</span> LWT at 5 weeks, as shown in E. Error bar: standard error of the mean; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.05; **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 2
<p>Absence of <span class="html-italic">Sox9</span> induces <span class="html-italic">Akt-YAP1</span>-mediated HC-derived panCK and HNF4α-positive HCC associated with liver progenitor cell characteristics. (<b>A</b>) Experimental design illustrating plasmids used for HDTVI, mice used in study, and time-points analyzed. (<b>B</b>) Representative serial sections of IHC images of 5-week <span class="html-italic">Akt-YAP1</span> LWT show CCA component to be positive for SOX9, YAP1, and panCK and HCC to be strongly positive for HNF4α, while in <span class="html-italic">Akt-YAP1 Sox9</span> LKO, no CCA was seen and HCC was negative for SOX9 but positive for HNF4α with panCK- and YAP1-positive cells interspersed in the tumor parenchyma. (<b>C</b>,<b>D</b>) qPCR data showing significantly decreased mRNA expression of <span class="html-italic">Sox9</span>, <span class="html-italic">CK19</span>, and <span class="html-italic">EpCAM</span> and significantly increased expression of <span class="html-italic">Tdo2</span> when comparing tumor-bearing livers in <span class="html-italic">Akt-YAP1</span> LWT and <span class="html-italic">Sox9</span> LKO models at 5 weeks. (<b>E</b>) Representative IHC staining for SOX9 and YAP1 depicting SOX9-low and nuclear YAP1-high or SOX9-high and YAP1-high cHCC-CCA from TMA (32 patients). TMA sections were enlarged for better view of nuclear expression of SOX9 and YAP1. Red arrows point to nuclear SOX9-high cells; black arrows point to nuclear YAP1-high cells; red empty arrows point to nuclear SOX9-negative cells; and black empty arrows point to nuclear YAP1-high cells. Percentage of patients positive for each combination are indicated. Scale bars: 100 µm; Error bar: standard error of the mean; * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 3
<p>RNA-seq analysis of mouse models and comparison with human liver cancer studies. (<b>A</b>) Heatmap for the differentially expressed genes comparing LWT and the <span class="html-italic">Akt-YAP1 Sox9</span> LKO model. (<b>B</b>) Common top enriched pathways between mouse (<span class="html-italic">Akt-YAP1 Sox9</span> LKO) and human (TCGA, LIHC) study. (<b>C</b>) Heatmap of gene signatures in the TCGA LIHC that are selected by the mouse model (LWT vs. <span class="html-italic">Akt-YAP1 Sox9</span> LKO). (<b>D</b>) Nearest Template Prediction (NTP) analysis of the TCGA LIHC whole-tumor gene expression dataset using the <span class="html-italic">Akt-YAP1 Sox9</span> LKO signature (AYSn signature) captured 12% of HCC. This subgroup of patients is enriched in the S1 class [<a href="#B22-cells-13-01451" class="html-bibr">22</a>] (<span class="html-italic">p</span> = 0.0015) and has an ICC-like signature [<a href="#B23-cells-13-01451" class="html-bibr">23</a>] (<span class="html-italic">p</span> = 0.03).</p>
Full article ">Figure 4
<p>SOX9 is required for <span class="html-italic">Akt-YAP1</span>-mediated HC reprogramming into CK19<sup>+</sup> mature CCA. (<b>A</b>) Experimental design illustrating plasmids used for HDTVI, mice used in study, and time-points analyzed. (<b>B</b>) Representative serial section IHC images of both <span class="html-italic">Akt-YAP1</span> LWT and <span class="html-italic">Akt-YAP1 Sox9</span> LKO show CCA-like components to be positive for SOX9, YAP1, panCK, and weak positive HNF4α with liver progenitor cell morphology (black dash lined) and HCC to be strongly positive for HNF4α at 2 weeks post-HDTVI. (<b>C</b>) Confocal images of immunofluorescence staining of LWT or <span class="html-italic">Sox9</span>-LKO livers at 2 and 5 weeks verify the essential roles for <span class="html-italic">Sox9</span> in CCA maturation. White dashed line points to panCK<sup>+</sup>;SOX9<sup>+</sup>;CK19<sup>+</sup> CCA cells and yellow dashed line to panCK<sup>+</sup>;SOX9<sup>−</sup>;CK19<sup>−</sup> LPC-like immature CCA cells. Scale bars: 100 µm.</p>
Full article ">Figure 5
<p>Developmental removal of <span class="html-italic">Sox9</span> promotes proliferation of <span class="html-italic">Akt-YAP1</span>-mediated liver cancer. (<b>A</b>) Experimental design illustrating plasmids used for HDTVI, mice used in study, and time-points analyzed. (<b>B</b>) Representative IF for Ki-67 (red), YAP1 (green), HA-tag (gray), and DAPI (blue) in liver sections from 5 weeks for <span class="html-italic">Akt-YAP1</span> LWT or <span class="html-italic">Sox9</span> LKO. Yellow arrows indicate Ki-67<sup>+</sup>;HA<sup>−</sup>tag+;YAP<sup>+</sup> proliferating liver cancer cells and Yellow empty arrows point to Ki-67<sup>−</sup>;HA-tag<sup>+</sup>;YAP<sup>+</sup> non-proliferating liver cancer cells. (<b>C</b>) The percentage of Ki-67-positive nuclei normalized to HA-tag-positive total tumor cell nuclei in representative images shown in B demonstrate significant increase in proliferation of transduced tumor cells in <span class="html-italic">Sox9</span> LKO as compared to LWT. (<b>D</b>) IHC for TUNEL to detect non-viable tumor cells shows comparable cell death was evident between <span class="html-italic">Sox9</span> LKO and LWT in <span class="html-italic">Akt-YAP1</span> model at 5-week time-point. Black arrows indicate nuclear TUNEL-positive apoptotic cells. (<b>E</b>) The number of TUNEL-positive nuclei normalized to field in representative images shown in D demonstrates comparable cell death in <span class="html-italic">Sox9</span> LKO as compared to WT. Scale bars: 100 µm; error bar: standard error of the mean; **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 6
<p>Acute <span class="html-italic">Sox9</span> deletion prevents the formation of combined HCC-CCA mediated by <span class="html-italic">Akt-YAP1</span> or <span class="html-italic">Akt-NRAS</span>, although <span class="html-italic">Akt-NRAS</span> tumor exhibit partial dependence on <span class="html-italic">Dnmt1</span>. (<b>A</b>) Experimental design illustrating plasmids used for HDTVI, mice used in study, and time-points analyzed. (<b>B</b>) A model illustrating the experimental design utilizing sleeping beauty transposon/transposase-CRISPR/Cas9-based inducible <span class="html-italic">Sox9</span> knockout plasmid. Representative gross images from <span class="html-italic">Akt-NRAS</span> (<b>C</b>) or <span class="html-italic">Akt-YAP1</span> (<b>D</b>)-injected WT (CWT), acute <span class="html-italic">Sox9</span>-knockout (CKO), and <span class="html-italic">Dnmt1</span>-injected Sox9-CKO (<span class="html-italic">Sox9</span> CKO-<span class="html-italic">Dnmt1</span>) livers showing tumor burden. (<b>E</b>) LW/BW ratio depicts significantly lower tumor burden in <span class="html-italic">Akt-NRAS</span> Sox9 CKO and <span class="html-italic">Akt-NRAS Sox9</span> CKO-<span class="html-italic">Dnmt1</span> mice as compared to CWT at 5 weeks. (<b>F</b>) LW/BW ratio depicts significantly lower tumor burden in <span class="html-italic">Akt-YAP1 Sox9</span> CKO and <span class="html-italic">Akt-YAP1 Sox9</span> CKO-<span class="html-italic">Dnmt1</span> mice as compared to CWT at 5 weeks. (<b>G</b>,<b>H</b>) Representative IHC images of tumor-bearing livers at 5 weeks in <span class="html-italic">Akt-NRAS</span> CWT, <span class="html-italic">Akt-NRAS Sox9</span> CKO, and <span class="html-italic">Akt-NRAS Sox9</span> CKO-<span class="html-italic">Dnmt1</span> liver stained for HA-tag and SOX9 showing cHCC-CCA component. HA-tag<sup>+</sup> <span class="html-italic">Akt-NRAS</span> cHCC/CCA tumor burden was robustly abrogated in <span class="html-italic">Sox9</span> CKO livers while being slightly but significantly restored in <span class="html-italic">Dnmt1</span>-injected <span class="html-italic">Sox9</span>-CKO livers. (<b>I</b>,<b>J</b>) Representative IHC images of tumor-bearing livers at 5 weeks in <span class="html-italic">Akt-YAP1</span> CWT, <span class="html-italic">Akt-YAP1 Sox9</span> CKO, and <span class="html-italic">Akt-YAP1 Sox9</span> CKO-<span class="html-italic">Dnmt1</span> liver stained for HA-tag and SOX9 showing cHCC-CCA component. HA-tag<sup>+</sup> <span class="html-italic">Akt-YAP1</span> cHCC/CCA tumor burden was robustly abrogated in both <span class="html-italic">Sox9</span> CKO and <span class="html-italic">Sox9</span> CKO-<span class="html-italic">Dnmt1</span> livers. Each dot in the graphs represent an individual mouse. Scale bars: 100 µm. Error bar: standard error of the mean; * <span class="html-italic">p</span> &lt; 0.05; **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 7
<p>Therapeutic <span class="html-italic">Sox9</span> elimination reduces <span class="html-italic">Akt-YAP1-dependent</span> advanced combined HCC-CCA. (<b>A</b>) Experimental design illustrating plasmids used for HDTVI, Tamoxifen treatment, mouse strain used in study, and time-points analyzed. (<b>B</b>,<b>D</b>) Representative gross images from <span class="html-italic">Akt-YAP1</span> (<b>B</b>) or <span class="html-italic">Akt-NRAS</span> (<b>D</b>)-injected <span class="html-italic">OPN-CreERT2</span> mice along with <span class="html-italic">Stop<sup>(f/f)−</sup>Cas9-sg-empty</span> (<span class="html-italic">Sox9</span>-iWT) and <span class="html-italic">Stop<sup>(f/f)</sup>-Cas9-sg-Sox9</span> (<span class="html-italic">Sox9</span>-iKO) treated with Tamoxifen (100 mg/kg triple) displaying gross tumor burden. (<b>C</b>,<b>E</b>) LW/BW ratio depicts significantly lower tumor burden in <span class="html-italic">Akt-YAP1</span> (<b>C</b>) but not in <span class="html-italic">Akt-NRAS</span> (E) <span class="html-italic">Sox9</span> iKO as compared to iWT at 6 weeks. (<b>F</b>) Representative IHC image of tumor-bearing livers at 6 weeks in <span class="html-italic">Akt-YAP1</span> iWT stained for HA-tag IHC showing component of the cHCC-CCA staining. <span class="html-italic">Sox9</span> iKO livers significantly reduced HA-tag<sup>+</sup> tumor burden. (<b>G</b>) Quantification of HA-tag IHC verifies significantly reduced staining in <span class="html-italic">Akt-YAP1 Sox9</span> iKO as compared to <span class="html-italic">Akt-YAP1 Sox9</span> iWT at 6 weeks, as shown in (<b>B</b>). Each dot in the graphs represent an individual mouse. Error bar: standard error of the mean; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">
15 pages, 790 KiB  
Review
Targeting the Hippo- Yes-Associated Protein/Transcriptional Coactivator with PDZ-Binding Motif Signaling Pathway in Primary Liver Cancer Therapy
by Yina Wang and Liangyou Rui
Onco 2024, 4(3), 217-231; https://doi.org/10.3390/onco4030016 - 22 Aug 2024
Viewed by 406
Abstract
Liver cancer imposes a pervasive global health challenge, ranking among the most prevalent cancers worldwide. Its prevalence and mortality rates are on a concerning upward trajectory and exacerbated by the dearth of efficacious treatment options. The Hippo signaling pathway, originally discovered in Drosophila, [...] Read more.
Liver cancer imposes a pervasive global health challenge, ranking among the most prevalent cancers worldwide. Its prevalence and mortality rates are on a concerning upward trajectory and exacerbated by the dearth of efficacious treatment options. The Hippo signaling pathway, originally discovered in Drosophila, comprises the following four core components: MST1/2, WW45, MOB1A/B, and LATS1/2. This pathway regulates the cellular localization of the transcriptional coactivator Yes-associated protein/transcriptional coactivator with PDZ-binding motif (YAP/TAZ) through a series of enzymatic reactions. The Hippo-YAP/TAZ pathway maintains a balance between cell proliferation and apoptosis, regulates tissue and organ sizes, and stabilizes the internal environment. Abnormalities of any genes within the Hippo signaling pathway, such as deletion or mutation, disturb the delicate balance between cell proliferation and apoptosis, creating a favorable condition for tumor initiation and progression. Mutations or epigenetic alterations in the Hippo signaling pathway components can lead to its inactivation. Consequently, YAP/TAZ becomes overexpressed and activated, promoting excessive cell proliferation and inhibiting apoptosis. This dysregulation is closely associated with the development of liver cancer. This review discusses the pivotal role of the Hippo signaling pathway in the pathogenesis and progression of liver cancer. By elucidating its mechanisms, we aim to offer new insights into potential therapeutic targets for effectively combating liver cancer. Full article
Show Figures

Figure 1

Figure 1
<p>Role of the Hippo-YAP/TAZ signaling pathway in primary liver cancer. YAP/TAZ: Yes-associated protein/transcriptional coactivator with PDZ-binding motif; MST1/2: Ste20-like kinases 1/2; LATS1/2: large tumor suppressor 1/2; CTGF: connective tissue growth factor; CYR61: cysteine-rich 61. This Figure was created with <a href="http://BioRender.com" target="_blank">BioRender.com</a>.</p>
Full article ">
Article
TLK1>Nek1 Axis Promotes Nuclear Retention and Activation of YAP with Implications for Castration-Resistant Prostate Cancer
by Damilola Olatunde and Arrigo De Benedetti
Cancers 2024, 16(16), 2918; https://doi.org/10.3390/cancers16162918 - 22 Aug 2024
Viewed by 631
Abstract
Despite some advances in controlling the progression of prostate cancer (PCa) that is refractory to the use of ADT/ARSI, most patients eventually succumb to the disease, and there is a pressing need to understand the mechanisms that lead to the development of CRPC. [...] Read more.
Despite some advances in controlling the progression of prostate cancer (PCa) that is refractory to the use of ADT/ARSI, most patients eventually succumb to the disease, and there is a pressing need to understand the mechanisms that lead to the development of CRPC. A common mechanism is the ability to integrate AR signals from vanishing levels of testosterone, with the frequent participation of YAP as a co-activator, and pointing to the deregulation of the Hippo pathway as a major determinant. We have recently shown that YAP is post-transcriptionally activated via the TLK1>NEK1 axis by stabilizing phosphorylation at Y407. We are now solidifying this work by showing the following: (1) The phosphorylation of Y407 is critical for YAP retention/partition in the nuclei, and J54 (TLK1i) reverses this along with YAP-Y407 dephosphorylation. (2) The enhanced degradation of (cytoplasmic) YAP is increased by J54 counteracting its Enzalutamide-induced accumulation. (3) The basis for all these effects, including YAP nuclear retention, can be explained by the stronger association of pYAP-Y407 with its transcriptional co-activators, AR and TEAD1. (4) We demonstrate that ChIP for GFP-YAP-wt, but hardly for the GFP-YAP-Y407F mutant, at the promoters of typical ARE- and TEAD1-driven genes is readily detected but becomes displaced after treatment with J54. (5) While xenografts of LNCaP cells show rapid regression following treatment with ARSI+J54, in the VCaP model, driven by the TMPRSS2-ERG oncogenic translocation, tumors initially respond well to the combination but subsequently recur, despite the continuous suppression of pNek1-T141 and pYAP-Y407. This suggests an alternative parallel pathway for CRPC progression for VCaP tumors in the long term, which may be separate from the observed ENZ-driven YAP deregulation, although clearly some YAP gene targets like PD-L1, that are found to accumulate following prolonged ENZ treatment, are still suppressed by the concomitant addition of J54. Full article
(This article belongs to the Section Cancer Therapy)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>YAP-increased expression in LNCaP treated with ENZ is suppressed with J54. (<b>A</b>) LNCaP cells grown in 6-well plates were treated with ENZ+/−J54 (1 µM each) for indicated times. Cell lysates (20 µg) were processed for WB for YAP and mRNA expression (<b>B</b>). (<b>C</b>) VCaP cells grown in 6-well plates were treated with ENZ+/−J54 (1 µM each) for 4 h, and cell lysates were thereafter processed for WB. The uncropped bolts are shown in <a href="#app1-cancers-16-02918" class="html-app">Supplementary Materials</a>.</p>
Full article ">Figure 2
<p>J54 elicits rapid GFP-YAP-Y407 dephosphorylation and nuclear export prior to cytoplasmic degradation—the default of an active Hippo pathway (LATS1-mediated pS396). Microscopic and WB depiction of the process and a graphical illustration. The uncropped bolts are shown in <a href="#app1-cancers-16-02918" class="html-app">Supplementary Materials</a>.</p>
Full article ">Figure 3
<p>Cell fractionation reveals nuclear localization of GFP-YAP-wt, and it is predominantly cytoplasmic for the Y407F mutant. (<b>A</b>) shows the subcellular localization of GFP YAP when the cells were probed with anti-GFP while (<b>B</b>) depicts the localization of the pYAP Y407 in the respective cells. (<b>C</b>) Subcellular redistribution YAP and pYAP Y407 upon treatment with J54 (a TLKi). Actin was used as a marker for the cytoplasmic fraction and was absent in the nuclei. Orc2 was used as a marker for the nuclei and was not present in the cytoplasm even when the blot was overexposed to reveal some cross-reacting bands known to be detected with this SL-Bio antiserum (see <a href="#app1-cancers-16-02918" class="html-app">Supplementary Materials</a>).</p>
Full article ">Figure 4
<p>Stronger association of GFP-YAP-wt with its transcriptional co-activators. (<b>A</b>) IPs were carried out with GFP antiserum, and WBs were probed for GFP, TEAD1, or AR. Inputs are also shown in the right panel. (<b>B</b>) A luciferase reporter assay showing the stronger association of the YAP-WT and its reversal with J54 treatment. (<b>C</b>) The Matrigel invasion assay reveals the invasive property of the respective cells and the effect of J54 treatment on YAP-WT’s invasive potential. (<b>D</b>) The immunoblot for MMPs ascertains the involvement of MMP9 and MMP10 for basement invasion. The uncropped bolts are shown in <a href="#app1-cancers-16-02918" class="html-app">Supplementary Materials</a>. * Significant as <span class="html-italic">p</span> &gt; 0.01.</p>
Full article ">Figure 5
<p>ChIP of GFP-YAP-wt vs. Y407F mutant at promoters of canonical CRE and ARE target genes reveals significantly different occupancy. PIS is pre-immune serum vs. GFP antiserum. GFP-YAP-WT increasingly occupied promoters of (<b>A</b>) FKBP5, (<b>B</b>) PSA, (<b>C</b>) SOX4, (<b>D</b>) SNX25, (<b>E</b>) CTGF, and (<b>F</b>) CYR61 genes compared to Y407F mutant. * Significant as <span class="html-italic">p</span> &gt; 0.01.</p>
Full article ">Figure 6
<p>Treatment of mice harboring VCaP subcutaneous flank tumors. (<b>A</b>,<b>B</b>) After inoculation of 10<sup>6</sup> cells in each flank of NOD-SCID mice, treatment started when the tumors reached 120 mm<sup>3</sup>, and resulted in a brief growth suppression with ENZ alone and was more sustained in combination with J54, but after ~2 months ((<b>A</b>)–end-point), most tumors relapsed and were processed for multi-panel WBs (<b>C</b>–<b>E</b>). Note that pNek1-T141 and pYAP-Y407 (<b>C</b>) were generally increased in animals treated with ENZ but suppressed when concomitantly treated with J54. Total YAP was slightly decreased with J54. The uncropped bolts are shown in <a href="#app1-cancers-16-02918" class="html-app">Supplementary Materials</a>.</p>
Full article ">Figure 6 Cont.
<p>Treatment of mice harboring VCaP subcutaneous flank tumors. (<b>A</b>,<b>B</b>) After inoculation of 10<sup>6</sup> cells in each flank of NOD-SCID mice, treatment started when the tumors reached 120 mm<sup>3</sup>, and resulted in a brief growth suppression with ENZ alone and was more sustained in combination with J54, but after ~2 months ((<b>A</b>)–end-point), most tumors relapsed and were processed for multi-panel WBs (<b>C</b>–<b>E</b>). Note that pNek1-T141 and pYAP-Y407 (<b>C</b>) were generally increased in animals treated with ENZ but suppressed when concomitantly treated with J54. Total YAP was slightly decreased with J54. The uncropped bolts are shown in <a href="#app1-cancers-16-02918" class="html-app">Supplementary Materials</a>.</p>
Full article ">
15 pages, 1934 KiB  
Article
Higher Circulating Neutrophil Counts Is Associated with Increased Risk of All-Cause Mortality and Cardiovascular Disease in Patients with Diabetic Kidney Disease
by Ruiyan Xie, David M. Bishai, David T. W. Lui, Paul C. H. Lee and Desmond Y. H. Yap
Biomedicines 2024, 12(8), 1907; https://doi.org/10.3390/biomedicines12081907 - 20 Aug 2024
Viewed by 573
Abstract
Background: Accumulating evidence has suggested the pathogenic roles of chronic inflammation and neutrophils in diabetic kidney disease (DKD). This study investigated the relationship between neutrophils, all-cause, and cardiovascular disease (CVD) mortality in type 2 diabetes mellitus (T2DM) patients with DKD. Methods: We used [...] Read more.
Background: Accumulating evidence has suggested the pathogenic roles of chronic inflammation and neutrophils in diabetic kidney disease (DKD). This study investigated the relationship between neutrophils, all-cause, and cardiovascular disease (CVD) mortality in type 2 diabetes mellitus (T2DM) patients with DKD. Methods: We used data from the National Health and Nutrition Examination Surveys (NHANES) from 2005 to 2020 to investigate the relationship between circulating neutrophils counts, kidney function indices, all-cause, and CVD mortality in adult T2DM patients with DKD. Clinical predictive models and risk scores for long-term mortality were constructed. Results: 44,332 patients [8034 with T2DM and 36,323 without T2DM] were included. Two thousand two hundred twenty patients had DKD, and 775 died (31.5% related to CVD) during a follow-up of 6.18 (range: 5.94–6.42) years. Higher neutrophil counts (Quartile 4, Q4) were associated with increased all-cause and CVD mortality [HR 1.73 (95% CI 1.34–2.25) and 1.81 (95% CI 1.14–2.89), respectively, p < 0.0001 and 0.01]. Neutrophil counts in Q4 showed a positive correlation with urine albumin-creatinine ratio (UACR) but a negative association with eGFR (p < 0.01 for all). Clinical predictive models incorporating neutrophil counts showed satisfactory performance in forecasting 5-year and 10-year CVD mortality-free survival (ROC AUC 0.824 and 0.842, respectively), and the nomogram-predicted survival demonstrated good concordance with observed survival. Conclusions: Higher levels of circulating neutrophil counts show a significant correlation with renal abnormalities and higher all-cause and CVD mortality in T2DM patients with DKD. The novel clinical predictive models and risk scores incorporating neutrophil counts may facilitate stratification and, hence, risk factor management in DKD patients. Full article
(This article belongs to the Section Endocrinology and Metabolism Research)
Show Figures

Figure 1

Figure 1
<p>The flow chart of this study.</p>
Full article ">Figure 2
<p>Dose-response relationship between circulating neutrophil counts of DKD patients and (<b>A</b>) CVD mortality and (<b>B</b>) all-cause mortality. Adjusted for age, sex, race, smoking, alcohol use, hypertension, hyperlipidemia, eGFR, and UACR in a logistic regression with the RCS model. The shaded area represents the estimated relative risk and the 95% CI. The vertical line represents cut-off value, and the horizontal dashed line represents reference line of no association is indicated at a hazard ration of 1.0. CI, confidence interval. <span class="html-italic">p</span> non-linearity = 0.0029 (CVD mortality) and 0.0000 (all-cause mortality), respectively.</p>
Full article ">Figure 3
<p>Clinical predictive models for long-term CVD mortality-free survival among individuals with DKD. (<b>A</b>) Nomogram for predicting 5- and 10-year CVD mortality-free survival between DKD patients in the development cohort. ROC curves of the predictive nomogram in (<b>B</b>) development and (<b>C</b>) validation cohorts. Q1, Quartile 1; Q2, Quartile 2; Q3, Quartile 3; Q4, Quartile 4.</p>
Full article ">Figure 4
<p>CVD-free mortality in patients with diabetic kidney disease according to risk scores and baseline neutrophils counts. (<b>A</b>,<b>B</b>) Development cohort and (<b>C</b>,<b>D</b>) Validation cohort.</p>
Full article ">Figure 5
<p>Survival is free of (<b>A</b>) cardiovascular disease mortality and (<b>B</b>) all-cause mortality in diabetic kidney disease patients with different quartiles of neutrophil counts.</p>
Full article ">
11 pages, 4595 KiB  
Communication
Nonlinear Optical Response of Au/CsPbI3 Quantum Dots and Its Laser Modulation Characteristics at 2.7 μm
by Mengqi Lv, Jin Zhao, Leilei Guo, Yanxu Zhang, Qiuling Zhao, Lihua Teng, Maorong Wang, Shuaiyi Zhang and Xia Wang
Micromachines 2024, 15(8), 1043; https://doi.org/10.3390/mi15081043 - 18 Aug 2024
Viewed by 544
Abstract
A passively Q-switched Er:YAP laser of 2.7 µm, utilizing Au-doped CsPbI3 quantum dots (QDs) as a saturable absorber (SA), was realized. It was operated stably with a minimum pulse width of 185 ns and a maximum repetition rate of 480 kHz. The [...] Read more.
A passively Q-switched Er:YAP laser of 2.7 µm, utilizing Au-doped CsPbI3 quantum dots (QDs) as a saturable absorber (SA), was realized. It was operated stably with a minimum pulse width of 185 ns and a maximum repetition rate of 480 kHz. The maximum pulse energy and the maximum peak power were 0.6 μJ and 2.9 W, respectively, in the Q-switched operation. The results show that the CsPbI3 QDs SA exhibits remarkable laser modulation properties at ~3 μm. Full article
(This article belongs to the Special Issue Optical and Laser Material Processing)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Normalized photoluminescence spectrum (PL) (red line) and absorption spectrum (blue line) for the CsPbI<sub>3</sub> perovskite QDs’ dispersion; (<b>b</b>) CsPbI<sub>3</sub>-Au perovskite QDs.</p>
Full article ">Figure 2
<p>(<b>a</b>) Transmission electron microscopy (TEM) image of CsPbI<sub>3</sub> perovskite QDs and (<b>d</b>) CsPbI<sub>3</sub>-Au perovskite QDs; elemental mapping of (<b>b</b>) I, (<b>c</b>) Pb, (<b>e</b>) Cs, and (<b>f</b>) Au.</p>
Full article ">Figure 3
<p>(<b>a</b>,<b>c</b>) Open-aperture Z-scan experimental results of CsPbI<sub>3</sub> perovskite QDs and CsPbI<sub>3</sub>-Au perovskite QDs, respectively, (<b>b</b>,<b>d</b>) and nonlinear transmission versus intensity of CsPbI<sub>3</sub> and CsPbI<sub>3</sub>-Au, respectively.</p>
Full article ">Figure 4
<p>Transmittance of CsPbI<sub>3</sub> QDs SA and Au-doped CsPbI<sub>3</sub> SA at 2.7 μm; inset: magnified view of transmittance of CsPbI<sub>3</sub> QDs SA and Au-doped CsPbI<sub>3</sub> SA in the range of 2400–3000 nm.</p>
Full article ">Figure 5
<p>Experimental scheme of the passively Q-switched Er:YAP laser based on the CsPbI<sub>3</sub> Au-doped QDs SA.</p>
Full article ">Figure 6
<p>Thermal focal length of Er:YAP crystal versus pump power.</p>
Full article ">Figure 7
<p>(<b>a</b>) Average output power of the Er:YAP laser versus various pump power for continuous wave (CW) operation using <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>o</mi> <mi>c</mi> </mrow> </msub> </mrow> </semantics></math> = 1% and 4%; (<b>b</b>) average output power of the Q-switched operation versus diverse incident power using <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>o</mi> <mi>c</mi> </mrow> </msub> </mrow> </semantics></math> = 1% and 4%; inset of (<b>a</b>,<b>b</b>) the laser spectrum at wavelengths of 2731.0 nm and 2730.8 nm, respectively; the passively Q-switched Er:YAP laser related parameters versus the absorbed pump power, (<b>c</b>) pulse width, (<b>d</b>) repetition rate, (<b>e</b>) peak power, and (<b>f</b>) pulse energy correspond to different saturated absorbers.</p>
Full article ">Figure 8
<p>Passively Q-switched pulse trains and single waveform in pulse trains of (<b>a</b>) CsPbI<sub>3</sub> SA and (<b>b</b>) Au-doped CsPbI<sub>3</sub> SA using <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>o</mi> <mi>c</mi> </mrow> </msub> </mrow> </semantics></math> = 1%; (<b>c</b>) CsPbI<sub>3</sub> SA and (<b>d</b>) Au-doped CsPbI<sub>3</sub> SAs using <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>o</mi> <mi>c</mi> </mrow> </msub> </mrow> </semantics></math> = 4% under pump power of 4.6 W.</p>
Full article ">Figure 9
<p>(<b>a</b>) Average output power fluctuations over time. (<b>b</b>) Beam quality of a passively Q-switched Er:YAP laser at an absorbed pump power of 4.6 W.</p>
Full article ">
19 pages, 1864 KiB  
Article
Effects of Pulsed Electromagnetic Field Treatment on Skeletal Muscle Tissue Recovery in a Rat Model of Collagenase-Induced Tendinopathy: Results from a Proteome Analysis
by Enrica Torretta, Manuela Moriggi, Daniele Capitanio, Carlotta Perucca Orfei, Vincenzo Raffo, Stefania Setti, Ruggero Cadossi, Laura de Girolamo and Cecilia Gelfi
Int. J. Mol. Sci. 2024, 25(16), 8852; https://doi.org/10.3390/ijms25168852 - 14 Aug 2024
Viewed by 453
Abstract
Tendon disorders often result in decreased muscle function and atrophy. Pulsed Electromagnetic Fields (PEMFs) have shown potential in improving tendon fiber structure and muscle recovery. However, the molecular effects of PEMF therapy on skeletal muscle, beyond conventional metrics like MRI or markers of [...] Read more.
Tendon disorders often result in decreased muscle function and atrophy. Pulsed Electromagnetic Fields (PEMFs) have shown potential in improving tendon fiber structure and muscle recovery. However, the molecular effects of PEMF therapy on skeletal muscle, beyond conventional metrics like MRI or markers of muscle decline, remain largely unexplored. This study investigates the metabolic and structural changes in PEMF-treated muscle tissue using proteomics in a rat model of Achilles tendinopathy induced by collagenase. Sprague Dawley rats were unilaterally induced for tendinopathy with type I collagenase injection and exposed to PEMFs for 8 h/day. Gastrocnemius extracts from untreated or PEMF-treated rats were analyzed with LC-MS/MS, and proteomics differential analysis was conducted through label-free quantitation. PEMF-treated animals exhibited decreased glycolysis and increased LDHB expression, enhancing NAD signaling and ATP production, which boosted respiratory chain activity and fatty acid beta-oxidation. Antioxidant protein levels increased, controlling ROS production. PEMF therapy restored PGC1alpha and YAP levels, decreased by tendinopathy. Additionally, myosins regulating slow-twitch fibers and proteins involved in fiber alignment and force transmission increased, supporting muscle recovery and contractile function. Our findings show that PEMF treatment modulates NAD signaling and oxidative phosphorylation, aiding muscle recovery through the upregulation of YAP and PGC1alpha and increasing slow myosin isoforms, thus speeding up physiological recovery. Full article
Show Figures

Figure 1

Figure 1
<p>Experimental design and label-free LC–ESI–MS/MS results. (<b>A</b>) Untreated rats. Type I collagenase was injected into the right limbs (COL), while PBS was injected into their left counterparts (PBS). Animals were sacrificed on day 21, 30, and 45 after the injection. Based on this, right muscles (COL) were named COL21, COL30, and COL45, while left muscles (PBS) were named PBS21, PBS30, and PBS45. (<b>B</b>) PEMF-treated rats. Animals were treated with PEMFs (1.5 mT SD 0.2; 75 Hz) for eight hours/day. Right muscles (COL + PEMF) were named COL + PEMF21, COL + PEMF30, and COL + PEMF45. (<b>C</b>) Venn diagram. Comparisons of shared and distinct significantly changed proteins in COL vs. PBS (dusty blue circles) and in COL + PEMF and COL (pink circles). Data resulted from label-free quantitation after LC–ESI–MS/MS analysis (ANOVA followed by Tukey’s multiple comparison test, <span class="html-italic">p</span>-value &lt; 0.05). The graphical illustration was generated using BioRender (version 4).</p>
Full article ">Figure 2
<p>Glucose and stress metabolism. (<b>A</b>) Summary of changed metabolic enzymes in glucose metabolism in muscles attached to inflamed tendons, both without (COL) and with PEMF treatment (PEMF). (<b>B</b>) Heatmap illustrating the expression profile of significantly increased (in red) or decreased (in green) enzymes (ANOVA and Tukey’s test, <span class="html-italic">p</span>-value &lt; 0.05) involved in glucose metabolism and stress response pathways. The graphical illustration was generated using BioRender (version 4).</p>
Full article ">Figure 3
<p>TCA cycle, fatty acid oxidation, and oxidative phosphorylation pathways. (<b>A</b>,<b>B</b>) Heatmaps illustrating the expression profile of significantly increased (in red) or decreased (in green) enzymes (ANOVA and Tukey’s test, <span class="html-italic">p</span>-value &lt; 0.05) involved in TCA cycle, fatty acid oxidation (<b>A</b>), and oxidative phosphorylation (<b>B</b>) pathways. (<b>C</b>) Graphical representation of metabolic canonical pathways: activated (z-score &gt; 2; orange arrows) or inhibited (z-blue &lt; 2; blue arrows) in muscles attached to inflamed tendons, without (COL) and with PEMF treatment (PEMF). (<b>D</b>) Heatmap displaying the expression profile of enzymes involved in the NAD signaling pathway. (<b>E</b>) Heatmap presenting the most significant upstream regulators. Orange- and blue-colored rectangles indicate predicted regulator activation or inhibition, respectively, via the z-score statistic. (<b>F</b>,<b>G</b>) Bar graphs depicting the expression of Yes-Associated Protein (YAP) (<b>F</b>) and peroxisome proliferator-activated receptor gamma coactivator 1-alpha (PGC1alpha) (<b>G</b>) in the <span class="html-italic">gastrocnemius</span> muscle from PBS, COL, and COL + PEMF groups (mean ± SD; * = significant difference, ANOVA and Tukey’s test, <span class="html-italic">n</span> = 2, * <span class="html-italic">p</span>-value &lt; 0.05; ** <span class="html-italic">p</span>-value &lt; 0.01; *** <span class="html-italic">p</span>-value &lt; 0.001). Full-length images are available in <a href="#app1-ijms-25-08852" class="html-app">Supplementary Figure S1</a>. The graphical illustration was generated using BioRender (version 4).</p>
Full article ">Figure 4
<p>Muscle fiber characterization and contractile proteins. (<b>A</b>–<b>C</b>) Bar graphs showing the distribution of myosin heavy chain (MyHC) isoforms in untreated and PEMF-treated animals sacrificed at 21 days (<b>A</b>), 30 days (<b>B</b>), and 45 days (<b>C</b>) after collagenase injection. ANOVA and Tukey’s test, <span class="html-italic">n</span> = 3, * <span class="html-italic">p</span>-value &lt; 0.05; ** <span class="html-italic">p</span>-value &lt; 0.01; *** <span class="html-italic">p</span>-value &lt; 0.001. Representative gel images are displayed. (<b>D</b>–<b>G</b>) Heatmaps illustrating the expression profile and % fold changes in significantly increased (in red) or decreased (in green) proteins (ANOVA and Tukey’s test, <span class="html-italic">p</span>-value &lt; 0.05) in sarcomere structures: thin filaments (<b>D</b>), thick filaments (<b>E</b>), M line (<b>F</b>), and Z disk (<b>G</b>). Full-length images are available in <a href="#app1-ijms-25-08852" class="html-app">Supplementary Figure S2</a>. The graphical illustration was generated using BioRender (version 4).</p>
Full article ">
18 pages, 6372 KiB  
Article
Activation of VGLL4 Suppresses Cardiomyocyte Maturational Hypertrophic Growth
by Aaron Farley, Yunan Gao, Yan Sun, Sylvia Zohrabian, William T. Pu and Zhiqiang Lin
Cells 2024, 13(16), 1342; https://doi.org/10.3390/cells13161342 - 13 Aug 2024
Viewed by 573
Abstract
From birth to adulthood, the mammalian heart grows primarily through increasing cardiomyocyte (CM) size, which is known as maturational hypertrophic growth. The Hippo-YAP signaling pathway is well known for regulating heart development and regeneration, but its roles in CM maturational hypertrophy have not [...] Read more.
From birth to adulthood, the mammalian heart grows primarily through increasing cardiomyocyte (CM) size, which is known as maturational hypertrophic growth. The Hippo-YAP signaling pathway is well known for regulating heart development and regeneration, but its roles in CM maturational hypertrophy have not been clearly addressed. Vestigial-like 4 (VGLL4) is a crucial component of the Hippo-YAP pathway, and it functions as a suppressor of YAP/TAZ, the terminal transcriptional effectors of this signaling pathway. To develop an in vitro model for studying CM maturational hypertrophy, we compared the biological effects of T3 (triiodothyronine), Dex (dexamethasone), and T3/Dex in cultured neonatal rat ventricular myocytes (NRVMs). The T3/Dex combination treatment stimulated greater maturational hypertrophy than either the T3 or Dex single treatment. Using T3/Dex treatment of NRVMs as an in vitro model, we found that activation of VGLL4 suppressed CM maturational hypertrophy. In the postnatal heart, activation of VGLL4 suppressed heart growth, impaired heart function, and decreased CM size. On the molecular level, activation of VGLL4 inhibited the PI3K-AKT pathway, and disrupting VGLL4 and TEAD interaction abolished this inhibition. In conclusion, our data suggest that VGLL4 suppresses CM maturational hypertrophy by inhibiting the YAP/TAZ-TEAD complex and its downstream activation of the PI3K-AKT pathway. Full article
(This article belongs to the Section Cells of the Cardiovascular System)
Show Figures

Figure 1

Figure 1
<p>The expression of cardiac lipid metabolism genes increases with age. (<b>A</b>). Heat map illustrating the differential expression data of selected cardiac lipid metabolism genes. E12, embryo age 12.5 days; P1, postnatal day 1; P42, postnatal day 42. (<b>B</b>–<b>D</b>). qRT-PCR measurement of cardiac lipid metabolism genes. RNA isolated from the hearts of different age mice were tested. For each group, four hearts were included. Statistical analysis was performed with one-way ANOVA followed by Tukey’s multiple comparison test. * indicates comparisons between E18.5 hearts and the other age hearts. *, <span class="html-italic">p</span> &lt; 0.05; **, <span class="html-italic">p</span> &lt; 0.01; ***, <span class="html-italic">p</span> &lt; 0.001; ****, <span class="html-italic">p</span> &lt; 0.0001. # indicates comparisons between P14 and P42 hearts. ##, <span class="html-italic">p</span> &lt; 0.01; ###, <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 2
<p>T3/Dex treatment promotes CM maturational hypertrophic growth. (<b>A</b>) Immunofluorescence images of NRVMs. Scale bar = 50 µm. White arrows indicate slim cell body; yellow arrowheads indicate protrusions extending out of the cell bodies. (<b>B</b>–<b>E</b>) Quantification of cell size (<b>B</b>), length (<b>C</b>), width (<b>D</b>), and length-to-width ratio (<b>E</b>). (<b>F</b>) Quantification of protrusion number per CM. Violin plot was used to display the distribution of CM populations that were categorized by protrusion number per cell. NS, not significant. (<b>B</b>–<b>F</b>) statistical analysis was performed with Kruskal–Wallis ranks test followed by Dunn’s multiple comparison test. * indicates comparisons between control and the other groups. * <span class="html-italic">p</span> &lt; 0.05; **, <span class="html-italic">p</span> &lt; 0.01; ***, <span class="html-italic">p</span> &lt; 0.001; ****, <span class="html-italic">p</span> &lt; 0.0001. # indicates comparisons between indicated groups. ###, <span class="html-italic">p</span> &lt; 0.001. For each group, 400–450 CMs were analyzed. (<b>G</b>,<b>H</b>) qRT-PCR measurement of NRVMs gene expression levels. Statistical analysis was performed with one-way ANOVA followed by Tukey’s multiple comparison test. * indicates comparisons between control and the other groups. **, <span class="html-italic">p</span> &lt; 0.01; ****, <span class="html-italic">p</span> &lt; 0.0001. # indicates comparisons between indicated groups. #, <span class="html-italic">p</span> &lt; 0.05; ###, <span class="html-italic">p</span> &lt; 0.001. For each group, <span class="html-italic">n</span> = 4. (<b>I</b>) Western blot of ATGL. Proteins isolated from NRVMs were analyzed. GAPDH was used as loading control. In the bar graph, ATGL protein level was normalized to GAPDH. (<b>J</b>,<b>K</b>) Mitochondria respiration activity measured by Seahorse XF. A total of 500 NRVMS were seeded and treated with indicated hormones for 2 days before being subjected to mitochondria stress test. OCR: oxygen consumption rate. (<b>J</b>) real-time measurement of OCR. (<b>K</b>) quantification of OCR under baseline and different stress conditions. Statistical analysis was performed with one-way ANOVA followed by Tukey’s multiple comparison test. * indicates comparisons between control and the other groups. *, <span class="html-italic">p</span> &lt; 0.05; **, <span class="html-italic">p</span> &lt; 0.01. # indicates comparisons between indicated groups. #, <span class="html-italic">p</span> &lt; 0.05. For each group, <span class="html-italic">n</span> = 6.</p>
Full article ">Figure 3
<p>T3/Dex induction of CM maturational hypertrophy depends on YAP/TAZ-TEAD complex. (<b>A</b>) YAP immunofluorescence staining. (<b>B</b>) Representative immunofluorescence images of NRVMs. (<b>A</b>,<b>B</b>) scale bar: 50 µm. (<b>C</b>) Quantification of NRVMs surface area. For each group, 150–200 cells from 4 biological replicates were measured. Kruskal–Wallis ranks test followed by Dunn’s multiple comparison test, *, <span class="html-italic">p</span> &lt; 0.05; ***, <span class="html-italic">p</span> &lt; 0.001. (<b>D</b>) qRT-PCR measurement of gene expression. 2 days after T3/Dex and YTIP treatment, NRVMs were collected for gene expression analysis. One-way ANOVA test followed by Tukey’s multiple comparison test: *, <span class="html-italic">p</span> &lt; 0.05; <span class="html-italic">n</span> = 4.</p>
Full article ">Figure 4
<p>Activation of VGLL4 suppresses T3/Dex-induced hypertrophic growth. NRVMs were treated with 100 nM T3 + 1 µM Dex (T/D) for 40–48 h to induce hypertrophy. Ad.VGLL4-GFP was used to activate VGLL4. (<b>A</b>) NRVMs stained with TNNI3 antibody. Scale bar: 50 µm. (<b>B</b>) Measurement of NRVMs surface area. A total of 200–350 cells from 4 biological replicates were measured. Kruskal–Wallis ranks test followed by Dunn’s multiple comparison test: ***, <span class="html-italic">p</span> &lt; 0.001. (<b>C</b>) Western blot of NRVMs transduced with Ad.VGLL4-GFP. β-actin was used as loading control. (<b>D</b>) qRT-PCR measurement of <span class="html-italic">Myh6</span> and <span class="html-italic">Myh7</span>. One-way ANOVA followed by Tukey’s multiple comparison test: *, <span class="html-italic">p</span> &lt; 0.05; **, <span class="html-italic">p</span> &lt; 0.01; ***, <span class="html-italic">p</span> &lt; 0.001; <span class="html-italic">n</span> = 4.</p>
Full article ">Figure 5
<p>Activation of VGLL4 suppresses CM maturational hypertrophic growth in vivo. AAV9.VGLL4<sup>K225R</sup> was subcutaneously injected into 3-day-old mouse pups. AAV9.Luci-transduced mouse pups was used as control. (<b>A</b>) Fraction shortening (FS%) measured by echocardiography at postnatal day 8 (P8) and day 19 (P19). Measurements were carried out on conscious mice. (<b>B</b>) Gross morphology of hearts. Scale bar: 2 mm. (<b>C</b>) Heart weight. (<b>D</b>) Body weight. (<b>E</b>) Heart-to-body weight ratio. (<b>A</b>,<b>C</b>–<b>E</b>) for each group, 4 mice were included. Student <span class="html-italic">t</span>-test, * <span class="html-italic">p</span> &lt; 0.05. (<b>F</b>) Representative images of heart sections stained with Wheat Germ Agglutinin (WGA). Scale bar: 50 µm. (<b>G</b>) CM cross-section area. A total of 4 hearts from each group were used for quantification. A total of 654 and 862 CMs were measured in Luci and VGLL4<sup>K225R</sup> hearts, respectively. Mann–Whitney test: **, <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 6
<p>VGLL4 suppresses T3/Dex-induced CM hypertrophy through TEAD. (<b>A</b>) Schematic view of the TDU domain sequences of wild-type and HF4A-mutated human VGLL4. Wild-type VGLL4 (V) and not the HF4A-mutated VGLL4 (V<sup>HF4A</sup>) blocks the formation of YAP/TAZ (Y/TAZ)–TEAD (T) complex. (<b>B</b>) Co-immunoprecipitation (Co-IP) assay. Indicated plasmids were transfected into 293T cells. 24 h post transfection, cells were lysed and applied for Co-IP. (<b>C</b>) Dual luciferase reporter assay in 293T cells. Dual luciferase assay was performed 24 h after transfection. 8xGTIIC: TEAD luciferase reporter plasmid. N = 6. (<b>D</b>) Representative immunofluorescence images of NRVMs. Scale bar: 50 µm. (<b>E</b>). Quantification of NRVMs surface area. For each group, 200–300 cells from 4 biological replicates were measured. Kruskal–Wallis ranks test followed by Dunn’s multiple comparison test: *, <span class="html-italic">p</span> &lt; 0.05; *** <span class="html-italic">p</span> &lt; 0.001. NS, not significant. (<b>F</b>,<b>G</b>) qRT-PCR measurements. 2 days after T3/Dex and indicated adenovirus treatment, NRVMs were collected for gene expression analysis. One-way ANOVA followed by Tukey’s multiple comparison test: *, <span class="html-italic">p</span> &lt; 0.05; **, <span class="html-italic">p</span> &lt; 0.01; ***, <span class="html-italic">p</span> &lt; 0.001, <span class="html-italic">n</span> = 5. NS, not significant. (<b>H</b>,<b>I</b>) Mitochondria respiration activity measured by Seahorse XF. (<b>H</b>) real-time measurement of OCR. (<b>I</b>) quantification of OCR under baseline and different stress conditions. Statistical analysis was performed with one-way ANOVA followed by Tukey’s multiple comparison test. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01, ***, <span class="html-italic">p</span> &lt; 0.001; <span class="html-italic">n</span> = 6.</p>
Full article ">Figure 7
<p>VGLL4 suppresses PI3K-Akt pathway. (<b>A</b>) qRT-PCR measurement of <span class="html-italic">Pik3ca</span> and <span class="html-italic">Pik3cb</span> expression levels. N = 4. (<b>B</b>) Western blot of phospho Akt (Ser 473) and total Akt. β-actin was used as loading control. (<b>C</b>) Densitometric quantification of total Akt. The total Akt densitometric value was normalized to that of β-actin. (<b>D</b>) Ratio of phospho Akt and total Akt. One-way ANOVA followed by Tukey’s multiple comparison test, *, <span class="html-italic">p</span> &lt; 0.05; <span class="html-italic">n</span> = 3.</p>
Full article ">Figure 8
<p>VGLL4 promotes TEAD1 degradation through CTSB. (<b>A</b>) Western blot of TEAD1 and VGLL4. 293T cells were transfected with indicated plasmids. 24 h after transfection, cells were treated with 10 µM CA074 for 12 h before being collected for Western blot analysis. GAPDH was used as loading control. (<b>B</b>) Densitometric quantification of TEAD1. TEAD1 densitometric value was normalized to that of GAPDH. One-way ANOVA followed by Tukey’s multiple comparison test, *, <span class="html-italic">p</span> &lt; 0.05. N = 3. (<b>C</b>) Western blot of TEAD1 and CTSB. 293T cells were transfected with indicated plasmids. 24 h after transfection, cells were collected for Western blot analysis. (<b>D</b>) Co-immunoprecipitation (Co-IP) assay. Indicated plasmids were transfected into 293T cells. 24 h post-transfection, cells were lysed and applied for Flag Ab Co-IP. (<b>E</b>) NRVMs Western blots. Serum-starved NRVMs were treated with indicated virus for 36–40 h before being collected for Western blot analysis. GAPDH was used as loading control. (<b>F</b>) Summary of the current study.</p>
Full article ">
37 pages, 4373 KiB  
Review
The Myofibroblast Fate of Therapeutic Mesenchymal Stromal Cells: Regeneration, Repair, or Despair?
by Fereshteh Sadat Younesi and Boris Hinz
Int. J. Mol. Sci. 2024, 25(16), 8712; https://doi.org/10.3390/ijms25168712 - 9 Aug 2024
Viewed by 747
Abstract
Mesenchymal stromal cells (MSCs) can be isolated from various tissues of healthy or patient donors to be retransplanted in cell therapies. Because the number of MSCs obtained from biopsies is typically too low for direct clinical application, MSC expansion in cell culture is [...] Read more.
Mesenchymal stromal cells (MSCs) can be isolated from various tissues of healthy or patient donors to be retransplanted in cell therapies. Because the number of MSCs obtained from biopsies is typically too low for direct clinical application, MSC expansion in cell culture is required. However, ex vivo amplification often reduces the desired MSC regenerative potential and enhances undesired traits, such as activation into fibrogenic myofibroblasts. Transiently activated myofibroblasts restore tissue integrity after organ injury by producing and contracting extracellular matrix into scar tissue. In contrast, persistent myofibroblasts cause excessive scarring—called fibrosis—that destroys organ function. In this review, we focus on the relevance and molecular mechanisms of myofibroblast activation upon contact with stiff cell culture plastic or recipient scar tissue, such as hypertrophic scars of large skin burns. We discuss cell mechanoperception mechanisms such as integrins and stretch-activated channels, mechanotransduction through the contractile actin cytoskeleton, and conversion of mechanical signals into transcriptional programs via mechanosensitive co-transcription factors, such as YAP, TAZ, and MRTF. We further elaborate how prolonged mechanical stress can create persistent myofibroblast memory by direct mechanotransduction to the nucleus that can evoke lasting epigenetic modifications at the DNA level, such as histone methylation and acetylation. We conclude by projecting how cell culture mechanics can be modulated to generate MSCs, which epigenetically protected against myofibroblast activation and transport desired regeneration potential to the recipient tissue environment in clinical therapies. Full article
Show Figures

Figure 1

Figure 1
<p>Tissue sources of therapeutic MSCs and stiffness-dependent differentiation. (<b>A</b>) The most prominently used tissue sources to isolate therapeutic MSCs from human biopsies include adipose tissues, bone marrow, and umbilical cord Wharton’s jelly. (<b>B</b>) The mechanical properties, i.e., softness or stiffness of tissues and those of the culture substrates used to grow and expand adhesive MSCs can influence MSC differentiation capacity and fate. MSCs cultured in soft cell culture environments matched to the elastic modulus (indicated in kPa) of normal fat and muscle tissue exhibit a high propensity for adipogenic and myogenic differentiation. In contrast, growth on stiffer culture substrates promotes the lineage commitment of MSCs towards cartilage and bone. One MSC fate, either representing a transitional state to osteogenesis or an independent scar-forming phenotype, is the activation of MSCs into fibrogenic myofibroblasts. Notably, the scars forming in response to the injury of soft tissues are always stiffer than the normal tissue texture (here schematized for skin), which drives the mechanically induced myofibroblast activation from resident and delivered mesenchymal cells. Cell culture plastic dishes are even stiffer (~10,000-times) than the stiff scar, which results in MSC-to-myofibroblast activation in vitro. Scheme produced with Biorender.</p>
Full article ">Figure 2
<p>MSC mechanoperception and nuclear mechanics. (<b>A</b>) The spreading area of mesenchymal stromal cells (MSCs) attaching to an adhesive substrate can be controlled using micropatterning; for instance, by transferring fibronectin protein (blue staining) in square shapes of different areas onto glass or plastic substrates using polydimethylsiloxane (PDMS) stamps. (<b>B</b>) Restricting MSC spreading limits the number and size of focal adhesions (green vinculin staining) and F-actin stress fibers (phalloidin, red), thus overall reducing MSC stress. (<b>C</b>) Another way to reduce stress on MSCs in culture is manipulating the elastic modulus of their substrate. MSCs perceive mechanical cues from the extracellular matrix (ECM) via transmembrane integrins; binding to extracellular ligands and intracellular F-actin shifts integrins from a low affinity inactive to a high-affinity active configuration. This integrin conformational switch prompts the assembly of complex focal adhesion structures comprising the cytosolic proteins talin, vinculin, focal adhesion kinase, paxillin, and filamin. Focal adhesions serve as hubs for mechanotransduction pathways, orchestrating the polymerization of G- into F-actin and the organization of vimentin monomers into intermediate filaments. Mechanical stress also opens stretch-activated channels (SACs) to allow the influx of Ca<sup>2+</sup> into the cytosol to trigger distinct signaling cascades. (<b>D</b>) The nuclei of MSCs grown on stiff surfaces are characterized by higher lamin A:C ratios in the inner nuclear membrane, more decondensed chromatin and higher histone acetylation compared to soft environments. A direct connection between ECM adhesions and the nucleus is established through the nucleoskeleton and cytoskeleton complex (LINC), containing nuclear envelope spectrin repeat proteins (nesprins) and Sad1p and UNC-84 homology (SUN) proteins that span the nuclear envelope. Nesprin-3 attaches SUN proteins to F-actin, whereas nesprins-1 and -2 link to intermediate filaments. Within the inner nuclear membrane, SUN dimers interact with lamin A bound to chromatin, causing organized DNA to unfold under high mechanical stress. High stress enhances the nuclear translocation of mechanosensitive transcription factors, such as MRTF-A, Runt-related transcription factor 2 (RUNX2), Yes-associated protein (YAP), and transcriptional coactivator with PDZ-binding motif (TAZ) via opening of the nuclear pore complex (NPC). The promoter binding of these transcription factors drives the expression of pro-fibrotic and osteogenic genes. Scheme elements produced using Biorender, immunofluorescence images produced by Nicole Berezyuk (Hinzlab).</p>
Full article ">Figure 3
<p>In vitro systems and mechanisms to generate mechanical memory in MSCs. (<b>A</b>) Seminal studies generated ‘long-term’ mechanical memory of lung fibroblasts [<a href="#B310-ijms-25-08712" class="html-bibr">310</a>] and MSC(M) [<a href="#B125-ijms-25-08712" class="html-bibr">125</a>] by culturing and adapting (‘priming’) cells for up to 3 weeks on either soft or stiff silicone elastomer substrates. Mechanical memory was defined as the capacity of MSCs to retain regenerative (soft) or pro-fibrotic and/or pro-osteogenic (stiff) features after switching to the respective substrate for another 2 weeks. (<b>B</b>) In the same study, growth on stiff culture substrates was shown to induce nuclear translocation of the mechanosensitive co-transcription factor myocardin-related transcription factor A (MRTF-A), where it drives the transcription of the profibrotic microRNA miR-21 [<a href="#B125-ijms-25-08712" class="html-bibr">125</a>]. Cytoplasmic miR-21 levels remain elevated for up to 2 weeks even after switching to soft substrates, whereas MRTF-A relocates to the cytosol within minutes. (<b>C</b>) In a different experimental approach to generate ‘short-term’ mechanical memory, MSCs and fibroblasts were cultured on stiff phototunable hydrogels for 10 d to acquire high levels of histone acetylation and low condensed chromatin [<a href="#B301-ijms-25-08712" class="html-bibr">301</a>,<a href="#B303-ijms-25-08712" class="html-bibr">303</a>]. Following in situ softening of the hydrogels using a light reaction, MSCs maintained high histone acetylation levels while showing increased chromatin condensation. The preserved histone acetylation can regulate chromatin accessibility and transcription profiles.</p>
Full article ">Figure 4
<p>Therapeutic effects of mechanically primed MSCs on rat wound healing. Skin wound healing was the first preclinical example to show a differential effect of soft- versus stiff-primed MSCs on tissue repair after transplantation [<a href="#B125-ijms-25-08712" class="html-bibr">125</a>]. Rat bone-marrow-derived MSCs (MSC(M)), primed for 3 weeks on either soft (5 kPa) or stiff (100 kPa) silicone culture substrates, were applied in a fibrin matrix to rat skin wounds, kept open, and made hypertrophic by a plastic frame splint. Shown are immunofluorescence images of 9-day-old wound tissue cross-sections. In this experimental model, soft primed MSC(M) suppress scar features such as enhanced wound tension which is not shown in the figure but in the published work [<a href="#B125-ijms-25-08712" class="html-bibr">125</a>], myofibroblast accumulation (red only, α-SMA), high vascularization (yellow, from co-staining of vascular smooth muscle for desmin, green, and α-SMA, red), and alignment of dense collagen extracellular matrix (only shown in the schematic). All these features are enhanced after delivery of stiff-primed MSC and even further accentuated in wounds that did not receive any MSCs.</p>
Full article ">
Back to TopTop