[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

Journals

Article Types

Countries / Regions

Search Results (268)

Search Parameters:
Keywords = hepatic stellate cells (HSC)

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
15 pages, 1387 KiB  
Review
Transplant Immunology in Liver Transplant, Rejection, and Tolerance
by Masaya Yokoyama, Daisuke Imai, Samuel Wolfe, Ligee George, Yuzuru Sambommatsu, Aamir A. Khan, Seung Duk Lee, Muhammad I. Saeed, Amit Sharma, Vinay Kumaran, Adrian H. Cotterell, Marlon F. Levy and David A. Bruno
Livers 2024, 4(3), 420-434; https://doi.org/10.3390/livers4030031 - 9 Sep 2024
Viewed by 454
Abstract
Liver transplantation is the most effective treatment for end-stage liver disease. Despite improvements in surgical techniques, transplant rejection remains a significant concern. The liver is considered an immune-privileged organ due to its unique microenvironment and complex interactions among various cell types. Alloimmune responses [...] Read more.
Liver transplantation is the most effective treatment for end-stage liver disease. Despite improvements in surgical techniques, transplant rejection remains a significant concern. The liver is considered an immune-privileged organ due to its unique microenvironment and complex interactions among various cell types. Alloimmune responses mediated by T cells and antigen-presenting cells (APCs) play crucial roles in transplant rejection. The liver’s dual blood supply and unique composition of its sinusoidal endothelial cells (LSECs), Kupffer cells (KCs), hepatocytes, and hepatic stellate cells (HSCs) contribute to its immune privilege. Alloantigen recognition by T cells occurs through direct, indirect, and semidirect pathways, leading to acute cellular rejection (ACR) and chronic rejection. ACR is a T cell-mediated process that typically occurs within the first few weeks to months after transplantation. Chronic rejection, on the other hand, is a gradual process characterized by progressive fibrosis and graft dysfunction, often leading to graft loss. Acute antibody-mediated rejection (AMR) is less common following surgery compared to other solid organ transplants due to the liver’s unique anatomy and immune privilege. However, when it does occur, AMR can be aggressive and lead to rapid graft dysfunction. Despite improvements in immunosuppression, rejection remains a challenge, particularly chronic rejection. Understanding the mechanisms of rejection and immune tolerance, including the roles of regulatory T cells (Tregs) and hepatic dendritic cells (DCs), is crucial for improving transplant outcomes. Strategies to induce immune tolerance, such as modulating DC function or promoting Treg activity, hold promise for reducing rejection and improving long-term graft survival. This review focuses on the liver’s unique predisposition to rejection and tolerance, highlighting the roles of individual cell types in these processes. Continued research into the mechanisms of alloimmune responses and immune tolerance in liver transplantation is essential for developing more effective therapies and improving long-term outcomes for patients with end-stage liver disease. Full article
(This article belongs to the Special Issue The Liver as the Center of the Internal Defence System of the Body)
Show Figures

Figure 1

Figure 1
<p>Immunological basis of T cell-mediated rejection.</p>
Full article ">Figure 2
<p>Recognition of alloantigen presentation.</p>
Full article ">Figure 3
<p>Pathways of antibody-mediated rejection.</p>
Full article ">
16 pages, 3625 KiB  
Article
In Vitro Investigation of the Anti-Fibrotic Effects of 1-Phenyl-2-Pentanol, Identified from Moringa oleifera Lam., on Hepatic Stellate Cells
by Watunyoo Buakaew, Sucheewin Krobthong, Yodying Yingchutrakul, Nopawit Khamto, Pornsuda Sutana, Pachuen Potup, Yordhathai Thongsri, Krai Daowtak, Antonio Ferrante, Catherine Léon and Kanchana Usuwanthim
Int. J. Mol. Sci. 2024, 25(16), 8995; https://doi.org/10.3390/ijms25168995 - 19 Aug 2024
Viewed by 758
Abstract
Liver fibrosis, characterized by excessive extracellular matrix deposition, is driven by activated hepatic stellate cells (HSCs). Due to the limited availability of anti-fibrotic drugs, the research on therapeutic agents continues. Here we have investigated Moringa oleifera Lam. (MO), known for its various bioactive [...] Read more.
Liver fibrosis, characterized by excessive extracellular matrix deposition, is driven by activated hepatic stellate cells (HSCs). Due to the limited availability of anti-fibrotic drugs, the research on therapeutic agents continues. Here we have investigated Moringa oleifera Lam. (MO), known for its various bioactive properties, for anti-fibrotic effects. This study has focused on 1-phenyl-2-pentanol (1-PHE), a compound derived from MO leaves, and its effects on LX-2 human hepatic stellate cell activation. TGF-β1-stimulated LX-2 cells were treated with MO extract or 1-PHE, and the changes in liver fibrosis markers were assessed at both gene and protein levels. Proteomic analysis and molecular docking were employed to identify potential protein targets and signaling pathways affected by 1-PHE. Treatment with 1-PHE downregulated fibrosis markers, including collagen type I alpha 1 chain (COL1A1), collagen type IV alpha 1 chain (COL4A1), mothers against decapentaplegic homologs 2 and 3 (SMAD2/3), and matrix metalloproteinase-2 (MMP2), and reduced the secretion of matrix metalloproteinase-9 (MMP-9). Proteomic analysis data showed that 1-PHE modulates the Wnt/β-catenin pathway, providing a possible mechanism for its effects. Our results suggest that 1-PHE inhibits the TGF-β1 and Wnt/β-catenin signaling pathways and HSC activation, indicating its potential as an anti-liver-fibrosis agent. Full article
Show Figures

Figure 1

Figure 1
<p>The dose–response curves of LX-2 cell viability after treatment with varying concentrations of crude Moringa extract or 1-phenyl-2-pentanol.</p>
Full article ">Figure 2
<p>The expression of liver fibrotic-associated genes and the MMP-9 level in crude MO extract and 1-PHE treated cells. The LX-2 cells at 2 × 10<sup>5</sup> cells/well were incubated with different concentrations of crude MO extract or 1-PHE in the presence of 10 ng/mL TGF-β1 for 48 h. The cells were harvested and the mRNA expression measured using real-time qRT-PCR. (<b>A</b>) The candidate liver fibrotic-associated genes, as shown above, include <span class="html-italic">COL1A1</span>, <span class="html-italic">TIMP1</span>, <span class="html-italic">MMP2</span>, <span class="html-italic">SMAD2</span>, and <span class="html-italic">SMAD3</span>. The relative gene expression was normalized to <span class="html-italic">GAPDH</span>. (<b>B</b>) The cell culture supernatant was collected and evaluated for the MMP-9 level. The data are presented as mean ± SD. <span class="html-italic">p</span>-value &lt; 0.0332 (*), <span class="html-italic">p</span>-value &lt; 0.0021 (**), <span class="html-italic">p</span>-value &lt; 0.0002 (***), <span class="html-italic">p</span>-value &lt; 0.0001 (****).</p>
Full article ">Figure 3
<p>The proteomic profiling of DEPs and GO annotation analysis. In total, 1570 DEPs were identified following treatment of LX-2 cells with 1-PHE and TGF-β1 (<b>A</b>). Statistical thresholds (log<sub>2</sub> fold change ≤1.5 or ≥1.5; <span class="html-italic">p</span>-value &lt; 0.05) were applied to distinguish significantly upregulated (n = 68; red dots) and downregulated (n = 30; blue dots) proteins (<b>B</b>). Protein–protein interaction (PPI) network analysis demonstrated interconnectedness within the sets of upregulated and downregulated DEPs (<b>C</b>,<b>D</b>). To illuminate the potential functions of these DEPs, Gene Ontology (GO) annotation was employed, categorizing proteins by biological process, cellular component, and molecular function (<b>E</b>,<b>F</b>).</p>
Full article ">Figure 4
<p>KEGG signaling pathway enrichment analysis of DEPs. KEGG signaling pathway enrichment analysis was performed on DEPs to elucidate the potential impact of 1-PHE on signaling pathways within LX-2 cells. Notably, Parkinson’s disease emerged as the most significantly enriched pathway associated with upregulated DEPs (<b>A</b>). In contrast, the renin secretion signaling pathway demonstrated the strongest enrichment among downregulated proteins (<b>B</b>). Intriguingly, the Wnt signaling pathway, a pathway strongly implicated in liver fibrosis and HSC activation, was also found to be downregulated (<b>C</b>). Downregulated genes were indicated by the color blue, while other genes within the pathway were represented by green. <span class="html-italic">LRP5</span>, LDL-receptor-related protein 5; <span class="html-italic">PRKACA</span>, protein kinase cAMP-activated catalytic subunit alpha.</p>
Full article ">Figure 5
<p>The 2D and 3D structural characteristics of ligand-protein complexes. Specifically, the interactions between PRKACA and two ligands, 1-PHE and 3SB, were visualized (<b>A</b>,<b>B</b>). Additionally, the interaction between 1-PHE and LRP5 was investigated (<b>C</b>).</p>
Full article ">Figure 6
<p>Illustration of the results of 250-nanosecond molecular dynamics simulations comparing the interactions of 1-PHE with PRKACA (<b>A</b>–<b>D</b>) and LRP5 (<b>E</b>–<b>H</b>). (<b>A</b>,<b>D</b>) present root mean square deviation (RMSD) plots, (<b>B</b>,<b>F</b>) display root mean square fluctuation (RMSF) plots, (<b>C</b>,<b>G</b>) depict numbers of hydrogen bonds, and (<b>D</b>,<b>H</b>) showcase radius of gyration (Rg) plots.</p>
Full article ">
19 pages, 9685 KiB  
Article
Association of IL-9 Cytokines with Hepatic Injury in Echinococcus granulosus Infection
by Tanfang Zhou, Xinlu Xu, Jiang Zhu, Mayire Aizezi, Aili Aierken, Menggen Meng, Rongdong He, Kalibixiati Aimulajiang and Hao Wen
Biomolecules 2024, 14(8), 1007; https://doi.org/10.3390/biom14081007 - 14 Aug 2024
Viewed by 580
Abstract
Cystic echinococcosis (CE) is a zoonotic disease caused by the parasite Echinococcus granulosus (E. granulosus), which can lead to the formation of liver lesions. Research indicates that E. granulosus releases both Toll-like receptor 2 (TLR2) and Interleukin-9 (IL-9), which can potentially [...] Read more.
Cystic echinococcosis (CE) is a zoonotic disease caused by the parasite Echinococcus granulosus (E. granulosus), which can lead to the formation of liver lesions. Research indicates that E. granulosus releases both Toll-like receptor 2 (TLR2) and Interleukin-9 (IL-9), which can potentially impair the body’s innate immune defenses and compromise the liver’s ability to fight against diseases. To investigate the role of TLR2 and IL-9 in liver damage caused by E. granulosus infection, samples were initially collected from individuals diagnosed with CE. Subsequently, BALB/c mice were infected with E. granulosus at multiple time points (4 weeks, 12 weeks, 32 weeks) and the expression levels of these markers was then assessed at each of these phases. Furthermore, a BALB/c mouse model was generated and administered anti-IL-9 antibody via intraperitoneal injection. The subsequent analysis focused on the TLR2/MyD88/NF-κB signaling pathway and the expression of IL-9 in E. granulosus was examined. A co-culture experiment was conducted using mouse mononuclear macrophage cells (RAW264.7) and hepatic stellate cells (HSCs) in the presence of E. granulosus Protein (EgP). The findings indicated elevated levels of IL-9 and TLR2 in patients with CE, with the activation of the signaling pathway significantly increased as the duration of infection progressed. Administration of anti-IL-9 in mice reduced the activation of the TLR2/MyD88/NF-κB signaling pathway, exacerbating liver injury. Moreover, EgP stimulates the TLR2/MyD88/NF-κB signaling pathway, resulting in the synthesis of α-SMA and Collagen I. The data suggest that infection with E. granulosus may stimulate the production of IL-9 through the activation of the TLR2/MyD88/NF-κB signaling pathway, which is mediated by TLR2. This activation stimulates RAW264.7 and HSCs, exacerbating liver injury and fibrosis. Full article
Show Figures

Figure 1

Figure 1
<p>The expression of TLR2 and IL-9 increased in CE. To confirm this, we used the ELISA method to accurately measure the levels of TLR2 (<b>A</b>), MyD88 (<b>B</b>), NF-κB p65 (<b>C</b>), and IL-9 (<b>D</b>) in serum samples obtained from a group of HC (n = 20) and another group of patients with CE (n = 20). Furthermore, PBMCs were extracted from the plasma of both groups, and the expression levels of TLR2 (<b>E</b>), MyD88 (<b>F</b>), NF-κB p65 (<b>G</b>), and IL-9 (<b>H</b>) mRNA were examined by RT-qPCR. In addition, we performed IHC labeling of TLR2, MyD88, and NF-κB p65 on liver sections obtained from both HC and CE groups. The resulting images (<b>I</b>) were magnified 200×, with the left side showing the overall view and the right side showing a close-up view. The percentage of positively stained cells was calculated to assess the expression of TLR2 (<b>J</b>), MyD88 (<b>K</b>), and NF-κB p65 (<b>L</b>) (n = 20). Statistics are displayed as the mean ± SD. ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 and **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 2
<p>The expression of TLR2 increased with the time of infection. For this investigation, we selected female BALB/c mice and established a mouse model of <span class="html-italic">E. granulosus</span> infection by injecting either 2000 PSCs or an equal amount of PBS through the hepatic portal vein (n = 6). Mice were euthanized at 4-, 12-, and 32-week post-infection to examine the immune response and pathological changes at different time intervals. The establishment of a mouse model and the experimental design. We used the ELISA technique to quantify the levels of IL-9 in the blood serum of mice at different time intervals (4 weeks, 12 weeks, 32 weeks) (<b>A</b>) (n = 6). Figure (<b>B</b>) displays the results. We also performed RT-qPCR analysis to examine the expression levels of TLR2 (<b>C</b>), MyD88 (<b>D</b>), and NF-κB p65 (<b>E</b>) mRNA in the livers of mice at different stages of infection (n = 6). To assess pathological changes and protein expression, we performed H&amp;E, PAS, and IHC staining (<b>F</b>) on mouse liver tissue sections. The liver tissue samples were analyzed to determine the expression levels of TLR2 (<b>G</b>), MyD88 (<b>H</b>), and NF-κB p65 (<b>I</b>) by IHC staining (n = 6). The percentage of positive regions was calculated to evaluate the expression levels. Statistics are displayed as the mean ± SD. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 and **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 3
<p>Injection of Anti-IL-9 reduced TLR2 in <span class="html-italic">E. granulosus</span>-infected mice. Our mIHC studies show that cells near of the lesion can be simultaneously labeled with DAPI (used for nuclear labeling), TLR2, and IL-9 on the same tissue slide (<b>A</b>). These results indicate a strong correlation between the expression of TLR2 and IL-9 and nuclear dispersion. In the experiment, female BALB/c mice were infected by injecting 2000 PSCs directly into the liver or an equal volume of PBS. After three days, the mice were randomly assigned to either the model or intervention group. The model group received intraperitoneal injections of the same volume of IgG, while the intervention group received injections of anti-IL-9. Injections were given at 3-day intervals for up to 12 weeks, followed by the euthanasia of the mice (<b>B</b>). Using the KEGG database for protein sequencing of the liver of mice injected with anti-IL-9, we discovered a remarkable association with the Toll-like receptor pathway (<b>C</b>). This suggests that IL-9 may modulate the immune response by affecting this specific pathway. We also quantified TLR2 levels in the plasma of three groups of mice using ELISA. The results showed a significant decrease in TLR2 levels in the plasma of mice injected with anti-IL-9 compared to the model group (<span class="html-italic">p</span> &lt; 0.05). This suggests that Anti-IL-9 may reduce the immune response by decreasing the level of TLR2 (<b>D</b>) (n = 6). Statistics are displayed as the mean ± SD. * <span class="html-italic">p</span> &lt; 0.05, **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 4
<p>Injection of anti-IL-9 reduced TLR2 in <span class="html-italic">E. granulosus</span>-infected mice. To investigate this further, we used RT-qPCR to evaluate the relative expression levels of TLR2 (<b>A</b>), MyD88 (<b>B</b>), NF-κB p65 (<b>C</b>), and IL-9 (<b>D</b>) mRNA in several experimental groups (n = 6). Furthermore, IHC, H&amp;E, and PAS staining techniques were used on mouse liver tissue sections to assess both pathological changes and protein expression levels (<b>E</b>). The hepatic expression levels of TLR2 (<b>F</b>), MyD88 (<b>G</b>), NF-κB p65 (<b>H</b>), and IL-9 (<b>I</b>) were quantitatively assessed by determining the proportion of positive cells in IHC staining (n = 6). Statistics are displayed as the mean ± SD. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 and **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 5
<p>The effects of immunity and fibrosis play an important role in RAW264.7 and HSCs. We first incubated RAW264.7 macrophages and hepatic stellate cells (HSCs) with 15 µg/mL of EgP for 24 h. Next, 2 µL of Diprovocim and 2 µL of C29 were added to the culture system. This process took 1 h, after which the cells were collected for further study. We used KEGG analysis to examine the histological results of mouse spleen lymphocytes induced by worm proteins. The results showed a strong correlation among the Toll-like receptor pathway, IL-9, the fibrosis markers COL1A1 and COL5A3, and the macrophage surface marker CD68 (n = 3) (<b>A</b>). We then used an inverted microscope to capture images of cells from different treatment groups to examine changes in cell shape (<b>B</b>). The expression of TLR2 (<b>C</b>), MyD88 (<b>D</b>), NF-κB p65 (<b>E</b>), and IL-9 (<b>F</b>) in RAW264.7 was analyzed by qRT-PCR (n = 3). The levels of TLR2, MyD88, NF-κB p65, and IL-9 in RAW264.7 cells were measured by western blotting (n = 3) (<b>G</b>). The mRNA expression levels of α-SMA (<b>H</b>) and Collagen I (<b>I</b>) in HSCs were investigated by qRT-PCR. The protein expression of α-SMA and Collagen I in HSCs was simultaneously measured by Western blotting (n = 3) (<b>J</b>). Statistics are displayed as the mean ± SD. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 and **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">
17 pages, 10148 KiB  
Article
Mesenchymal Stromal Cell-Derived Extracellular Vesicles for Reversing Hepatic Fibrosis in 3D Liver Spheroids
by Giulia Chiabotto, Armina Semnani, Elena Ceccotti, Marco Guenza, Giovanni Camussi and Stefania Bruno
Biomedicines 2024, 12(8), 1849; https://doi.org/10.3390/biomedicines12081849 - 14 Aug 2024
Viewed by 951
Abstract
Hepatic fibrosis, arising from prolonged liver injury, entails the activation of hepatic stellate cells (HSCs) into myofibroblast-like cells expressing alpha-smooth muscle actin (α-SMA), thereby driving extracellular matrix deposition and fibrosis progression. Strategies targeting activated HSC reversal and hepatocyte regeneration show promise for fibrosis [...] Read more.
Hepatic fibrosis, arising from prolonged liver injury, entails the activation of hepatic stellate cells (HSCs) into myofibroblast-like cells expressing alpha-smooth muscle actin (α-SMA), thereby driving extracellular matrix deposition and fibrosis progression. Strategies targeting activated HSC reversal and hepatocyte regeneration show promise for fibrosis management. Previous studies suggest that extracellular vesicles (EVs) from mesenchymal stromal cells (MSCs) can suppress HSC activation, but ensuring EV purity is essential for clinical use. This study investigated the effects of MSC-derived EVs cultured in chemically defined conditions on liver spheroids and activated HSCs. Umbilical cord- and bone marrow-derived MSCs were expanded in chemically defined media, and EVs were isolated using filtration and differential ultracentrifugation. The impact of MSC-EVs was evaluated on liver spheroids generated in Sphericalplate 5D™ and on human HSCs, both activated by transforming growth factor beta 1 (TGF-β1). MSC-EVs effectively reduced the expression of profibrotic markers in liver spheroids and activated HSCs induced by TGF-β1 stimulation. These results highlight the potential of MSC-EVs collected under chemically defined conditions to mitigate the activated phenotype of HSCs and liver spheroids, suggesting MSC-EVs as a promising treatment for hepatic fibrosis. Full article
(This article belongs to the Special Issue 3D Cell Culture Systems for Biomedical Research)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Set up of liver spheroids using HepG2 and LX-2. (<b>A</b>) Diagram illustrating the experimental setup performed in vitro to generate liver spheroids using the co-culture of HepG2 and LX-2 or HepG2 only; created with BioRender.com. (<b>B</b>) Morphological observation in light microscopy of liver spheroids from day 1 to day 6 after cell seeding (scale bar, 100 µm). (<b>C</b>) Quantitative RT-PCR analysis of profibrotic gene expression in liver spheroids up to 6 days post-seeding. Expression levels were normalized to the <span class="html-italic">TBP</span> reference gene. Liver spheroids of HepG2 and LX-2 not activated by TGF-β1 served as reference control. Data from at least three independent experiments were analyzed using a two-way ANOVA 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; and **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 1 Cont.
<p>Set up of liver spheroids using HepG2 and LX-2. (<b>A</b>) Diagram illustrating the experimental setup performed in vitro to generate liver spheroids using the co-culture of HepG2 and LX-2 or HepG2 only; created with BioRender.com. (<b>B</b>) Morphological observation in light microscopy of liver spheroids from day 1 to day 6 after cell seeding (scale bar, 100 µm). (<b>C</b>) Quantitative RT-PCR analysis of profibrotic gene expression in liver spheroids up to 6 days post-seeding. Expression levels were normalized to the <span class="html-italic">TBP</span> reference gene. Liver spheroids of HepG2 and LX-2 not activated by TGF-β1 served as reference control. Data from at least three independent experiments were analyzed using a two-way ANOVA 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; and **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 2
<p>Set up of liver spheroids using UpHep and LX-2. (<b>A</b>) Diagram illustrating the experimental setup performed in vitro to generate liver spheroids using the co-culture of UpHep and LX-2 or UpHep only; created with BioRender.com. (<b>B</b>) Morphological observation in light microscopy of liver spheroids from day 1 to day 6 after cell seeding (scale bar, 100 µm). (<b>C</b>,<b>D</b>) Quantitative RT-PCR analysis evaluating the expression of profibrotic (<b>C</b>) and hepatocyte-specific (<b>D</b>) genes in liver spheroids up to 6 days after seeding. Expression levels were normalized to the <span class="html-italic">TBP</span> reference gene. Liver spheroids of UpHep not activated by TGF-β1 served as reference control. Data from at least three independent experiments were analyzed using a two-way ANOVA 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; and **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 2 Cont.
<p>Set up of liver spheroids using UpHep and LX-2. (<b>A</b>) Diagram illustrating the experimental setup performed in vitro to generate liver spheroids using the co-culture of UpHep and LX-2 or UpHep only; created with BioRender.com. (<b>B</b>) Morphological observation in light microscopy of liver spheroids from day 1 to day 6 after cell seeding (scale bar, 100 µm). (<b>C</b>,<b>D</b>) Quantitative RT-PCR analysis evaluating the expression of profibrotic (<b>C</b>) and hepatocyte-specific (<b>D</b>) genes in liver spheroids up to 6 days after seeding. Expression levels were normalized to the <span class="html-italic">TBP</span> reference gene. Liver spheroids of UpHep not activated by TGF-β1 served as reference control. Data from at least three independent experiments were analyzed using a two-way ANOVA 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; and **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 3
<p>Characterization of BM-MSC-EVs and UC-MSC-EVs. (<b>A</b>,<b>B</b>) Representative graphs illustrating EV size distribution via Nanoparticle Tracking Analysis (red lines indicate the mean of the three recorded videos), alongside micrographs of transmission electron microscopy showcasing intact BM-MSC-EVs (<b>A</b>) and UC-MSC-EVs (<b>B</b>). EVs underwent negative staining using NanoVan (scale bar, 200 nm). (<b>C</b>,<b>D</b>) The molecular surface profiles of BM-MSC-EVs (<b>C</b>) and UC-MSC-EVs (<b>D</b>) were assessed via a multiplex bead-based flow cytometry assay. Exosomal markers and mesenchymal markers are indicated by black bars and light blue bars, respectively. The two graphs present a quantification of median APC fluorescence values for 39 different bead populations following background correction. Notably, no statistically significant variances were detected among three distinct preparations of BM-MSC-EVs and UC-MSC-EVs. (<b>E</b>) Representative Western blot analysis depicting the presence of exosomal markers in EVs. The cis-Golgi marker GM130, absent in EVs but expressed in cells, served as the negative control.</p>
Full article ">Figure 4
<p>Antifibrotic effect of EVs on hepatic spheroids treated with TGF-β1. (<b>A</b>) Diagram illustrating the experimental setup conducted in vitro to assess the impact of EVs derived from BM-MSCs and UC-MSCs on spheroids generated using HepG2 and LX-2 or UpHep only; created with BioRender.com. (<b>B</b>,<b>C</b>) Quantitative RT-PCR analysis evaluating the expression of profibrotic genes in spheroids generated using HepG2 and LX-2 (<b>B</b>) or UpHep only (<b>C</b>) after stimulation with EVs obtained from BM-MSCs and UC-MSCs. Expression levels were normalized to the TBP reference gene. (<b>D</b>,<b>E</b>) Quantification of protein bands and representative images from Western blot analysis showing profibrotic markers in spheroids formed by HepG2 and LX-2 (<b>D</b>) or UpHep only (<b>E</b>). Protein band intensities were normalized to vinculin expression. All the experiments’ reference controls consisted of liver spheroids activated with TGF-β1 (10 ng/mL) but not treated with EVs. The negative control (CTR) consisted of liver spheroids cultured in the absence of TGF-β1. Data from at least three independent experiments were analyzed using a two-way ANOVA 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; and **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 5
<p>Antifibrotic effect of EVs on TGF-β1-activated LX-2. (<b>A</b>) Diagram illustrating the experimental setup conducted in vitro to assess the impact of EVs derived from BM-MSCs and UC-MSCs on LX-2 cells; created with BioRender.com. (<b>B</b>) Quantitative RT-PCR analysis of profibrotic gene expression in LX-2 cells after a 24 h incubation period with EVs obtained from BM-MSCs and UC-MSCs. Expression levels were normalized to the <span class="html-italic">TBP</span> reference gene. Reference control consisted of LX-2 cells activated with TGF-β1 (10 ng/mL) but not treated with EVs. The negative control (CTR) consisted of LX-2 cells cultured in the absence of TGF-β1. Data from at least three independent experiments were analyzed using a two-way ANOVA test: * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; and **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">
22 pages, 777 KiB  
Review
Fibrosis and Hepatocarcinogenesis: Role of Gene-Environment Interactions in Liver Disease Progression
by Anindita Banerjee and Patrizia Farci
Int. J. Mol. Sci. 2024, 25(16), 8641; https://doi.org/10.3390/ijms25168641 - 8 Aug 2024
Viewed by 576
Abstract
The liver is a complex organ that performs vital functions in the body. Despite its extraordinary regenerative capacity compared to other organs, exposure to chemical, infectious, metabolic and immunologic insults and toxins renders the liver vulnerable to inflammation, degeneration and fibrosis. Abnormal wound [...] Read more.
The liver is a complex organ that performs vital functions in the body. Despite its extraordinary regenerative capacity compared to other organs, exposure to chemical, infectious, metabolic and immunologic insults and toxins renders the liver vulnerable to inflammation, degeneration and fibrosis. Abnormal wound healing response mediated by aberrant signaling pathways causes chronic activation of hepatic stellate cells (HSCs) and excessive accumulation of extracellular matrix (ECM), leading to hepatic fibrosis and cirrhosis. Fibrosis plays a key role in liver carcinogenesis. Once thought to be irreversible, recent clinical studies show that hepatic fibrosis can be reversed, even in the advanced stage. Experimental evidence shows that removal of the insult or injury can inactivate HSCs and reduce the inflammatory response, eventually leading to activation of fibrolysis and degradation of ECM. Thus, it is critical to understand the role of gene-environment interactions in the context of liver fibrosis progression and regression in order to identify specific therapeutic targets for optimized treatment to induce fibrosis regression, prevent HCC development and, ultimately, improve the clinical outcome. Full article
(This article belongs to the Special Issue Molecular Research of Hepatocellular Carcinoma)
Show Figures

Figure 1

Figure 1
<p>Pathogenesis and dynamics of liver fibrosis progression and regression. Hepatocyte injury leads to activation of inflammatory mediators and transformation of HSCs into myofibroblasts that, through an autocrine loop, aberrantly produce collagen alongside a disruption of the MMP/TIMP homeostasis. NF-kB, nuclear factor kappa B; TGF-β, transforming growth factor beta; Hh, Hedgehog pathway; IL, interleukin; TNFα, tumor necrosis factor alpha; PI3K, phosphoinositide 3-kinase; CXCR7, chemokine receptor 7; FGF1R, fibroblast growth factor 1 receptor; HGF, hepatocyte growth factor; VEGF, vascular endothelial growth factor; MMP, matrix metalloproteases; TIMP, tissue inhibitors of metalloproteinases. Image created in BioRender.com (accessed on 20 June 2024).</p>
Full article ">
19 pages, 2347 KiB  
Review
Noncoding RNA-Mediated Epigenetic Regulation in Hepatic Stellate Cells of Liver Fibrosis
by Ruoyu Gao and Jingwei Mao
Non-Coding RNA 2024, 10(4), 44; https://doi.org/10.3390/ncrna10040044 - 7 Aug 2024
Viewed by 724
Abstract
Liver fibrosis is a significant contributor to liver-related disease mortality on a global scale. Despite this, there remains a dearth of effective therapeutic interventions capable of reversing this condition. Consequently, it is imperative that we gain a comprehensive understanding of the underlying mechanisms [...] Read more.
Liver fibrosis is a significant contributor to liver-related disease mortality on a global scale. Despite this, there remains a dearth of effective therapeutic interventions capable of reversing this condition. Consequently, it is imperative that we gain a comprehensive understanding of the underlying mechanisms driving liver fibrosis. In this regard, the activation of hepatic stellate cells (HSCs) is recognized as a pivotal factor in the development and progression of liver fibrosis. The role of noncoding RNAs (ncRNAs) in epigenetic regulation of HSCs transdifferentiation into myofibroblasts has been established, providing new insights into gene expression changes during HSCs activation. NcRNAs play a crucial role in mediating the epigenetics of HSCs, serving as novel regulators in the pathogenesis of liver fibrosis. As research on epigenetics expands, the connection between ncRNAs involved in HSCs activation and epigenetic mechanisms becomes more evident. These changes in gene regulation have attracted considerable attention from researchers in the field. Furthermore, epigenetics has contributed valuable insights to drug discovery and the identification of therapeutic targets for individuals suffering from liver fibrosis and cirrhosis. As such, this review offers a thorough discussion on the role of ncRNAs in the HSCs activation of liver fibrosis. Full article
(This article belongs to the Collection Feature Papers in Non-Coding RNA)
Show Figures

Figure 1

Figure 1
<p>The biological pathways and interactions of three ncRNAs. MiRNAs are produced from pri-miRNAs transcribed by RNA polymerase II from independent genes or introns of protein genes. The pri-miRNAs are processed by a complex involving Drosha and DiGeorge syndrome critical region 8 (DGCR8) to create pre-miRNAs, which are then exported to the cytoplasm by Exportin 5. In the cytoplasm, Dicer cleaves the pre-miRNA to form a miRNA duplex, with one strand being degraded and the other becoming the mature miRNA. The mature miRNA, along with Argonaute2 (AGO2) and glycine-tryptophan repeat-containing protein of 182 KDa (GW182), binds to the 3′-untranslated region (3′-UTR) of targeted mRNA to inhibit its translation. LncRNAs are transcribed by RNA polymerase II and have a structure similar to mRNA. They can be transported from the nucleus to the cytoplasm by nuclear RNA export factor 1 (NXF1). Some lncRNAs act as competitive endogenous RNAs (ceRNAs), binding to miRNAs to affect gene expression. LncRNAs can also interact with RBPs to stabilize and promote mRNA translation. CircRNAs are produced from pre-mRNA through backsplicing and can regulate gene expression by acting as ceRNAs to inhibit miRNA and increase expression of targeted genes.</p>
Full article ">Figure 2
<p>TGF-β1/Smad pathway mediated by identified ncRNAs. The TGF-β signaling pathway involves three isoforms, namely TGF-β1, TGF-β2, and TGF-β3. This discussion will focus specifically on the mechanism of TGF-β1. The precursor form of TGF-β contains a latency-associated peptide (LAP), which can be cleaved to allow binding to the mature TGF-β homodimer. This complex, along with the latent TGF-β-binding protein (LTBP), forms the latent TGF-β/LAP/LTBP complex, maintaining TGF-β in an inactive state unable to interact with TGF-βR I and TGF-βR II. Upon release of the TGF-β homodimer from the complex, it becomes active and can interact with TGF receptor II (TGFR II), leading to TGFR I activation and subsequent phosphorylation of Smad2 and Smad3. The phosphorylated forms of Smad2 and Smad3 form a complex with Smad4, translocate to the nucleus, and bind to the promoter regions of fibrotic genes, thereby inducing their transcription. Additionally, Smad7, acting as an inhibitory Smad, competitively interacts with TGFR I to inhibit the binding between TGFR I and Smad2/3, thereby suppressing the activation of the TGF-β1/Smad pathway. Furthermore, Smad7 recruits E3 ubiquitin ligase and phosphatases to facilitate the degradation and dephosphorylation of Smad2/3. Left–right determination factor 2 (LEFTY2) is a member of the TGF-β protein superfamily, which can inhibit TGF-β1/Smad3 signaling. Follistatin-like 1 (FSTL1) is an inducer of TGF-β1/Smad3 signaling through promoting the Smad3 phosphorylation. Some identified ncRNAs can also regulate the TGF-β/Smad pathway by targeting various components of this signaling cascade.</p>
Full article ">Figure 3
<p>Hippo and Hedgehog pathways mediated by identified ncRNAs. The Hippo pathway is activated by TAO kinase, which subsequently phosphorylates mammalian Ste20-like kinase 1/2 (MST1/2). Phosphorylated MST1/2 then phosphorylates scaffold proteins SAV1 and MOB1A/B, facilitating the recruitment and phosphorylation of the large tumor suppressor 1/2 (LATS1/2). Phosphorylated LATS1/2 facilitates the phosphorylation of transcriptional co-activators Yes-associated protein (YAP) and transcriptional co-activator with PDZ-binding motif (TAZ), leading to their cytoplasmic retention mediated by 14-3-3 and subsequent degradation. In the absence of Hippo pathway activation, the kinase cascades are not initiated, allowing YAP/TAZ co-activators to translocate into the nucleus and interact with the TEAD transcription factor family to modulate gene expression. In the absence of Hedgehog ligands, the twelve-pass transmembrane receptor Ptched (Ptch) is able to inhibit the activity of the seven-pass transmembrane receptor Smoothene (Smo). This inhibition of Smo leads to the inactivation of glioma-associated oncogene transcription factors (Glis) through the formation of a complex with suppressor of Fused (SuFu) and Kif7. This complex then promotes the phosphorylation of Glis by protein kinase (PKA), casein kinaseⅠɑ (CK1ɑ), and glycogen synthase kinase-3β (GSK3β), resulting in the repression of subsequent transcription. When the Hh protein binds to the Ptch receptor, it leads to a reduction in the inhibition of Smo, allowing for the phosphorylation of Smo by CK1 family kinase and GRK2. This phosphorylation event induces the release of Gli from the complex, enabling its translocation into the nucleus to regulate the expression of targeted genes. Additionally, the regulation of these pathways is influenced by ncRNAs, as depicted in this figure.</p>
Full article ">Figure 4
<p>DNA methylation usually occurs in the region enrichment with cytosine–phosphate–guanine (CpG) dinucleotides that are also called CpG islands. DNMTs catalyze the methyl group transference from S-adenyl methionine (SAM) to the fifth carbon of cytosine residue to form 5-methylcytosine (5mc). DNMT3a and DNMT3b catalyze the de novo methylation, namely, add the 5mc to the DNA directly. DNMT1 plays a role in maintaining methylation in DNA replication. A family of TET enzymes can superinduce a hydroxyl group to the 5mc; this process converts the 5mc to 5-hydroxymethylcytosine (5hmc). MecP2 can bind with methylated CpG to suppress or enhance gene expression. The regulation of these processes is influenced by ncRNAs, as depicted in this figure.</p>
Full article ">Figure 5
<p>Wnt/β-catenin pathway and associated ncRNAs. When the Wnt signaling pathway is not effectively inhibited by Dickkopf (DKK) or Wnt inhibitory protein (WIF), which interact with lipoprotein receptor-related 5/6 (LRP5/6) to disrupt LRP5/6 and Frizzled (FZD) receptor dimers, Wnt proteins are able to bind to their FZD receptors, leading to dimerization of FZD and LRP5/6 receptors. The formation of FZD/LRP heterodimers induces a conformational change in the receptors, leading to the binding of the cytoplasmic portion of FZD to disheveled (DVL) and phosphorylation of the LRP5/6 tail by GSK3. This phosphorylation event facilitates the recruitment of the scaffold protein Axin. DVL serves as a platform for enhanced interaction between Axin and the LRP5/6 tail. The interaction between Axin, LRP5/6, and DVL disrupts the Destructive Complex (DC), releasing β-catenin. Subsequently, β-catenin translocates into the nucleus, where it displaces corepressor Groucho/transducin with TCF/LEF to form a complex that mediates gene expression. The destruction complex (DC), composed of Axin, adenomatous polyposis (APC), glycogen synthase kinase 3β (GSK3β), and casein kinase 1ɑ (CK1ɑ), functions to maintain β-catenin in an inactive state through phosphorylation by CK1ɑ, facilitating GSK3β-mediated phosphorylation of β-catenin. Subsequent phosphorylation of β-catenin leads to recruitment of β-transducin repeat containing protein (β-Trcp), an E3 ubiquitin ligase that promotes ubiquitination of β-catenin for degradation. The Wnt/β-catenin pathway is also regulated by various ncRNAs.</p>
Full article ">
18 pages, 3536 KiB  
Article
Enhanced In Vitro Efficacy of Verbascoside in Suppressing Hepatic Stellate Cell Activation via ROS Scavenging with Reverse Microemulsion
by Xiao Xiao, Feiyu Yang, Yuling Huang, Shaohui Liu, Zhenhua Hu, Shanggao Liao and Yuanyuan Li
Antioxidants 2024, 13(8), 907; https://doi.org/10.3390/antiox13080907 - 27 Jul 2024
Viewed by 485
Abstract
Numerous approaches targeting hepatic stellate cells (HSCs) have emerged as pivotal therapeutic strategies to mitigate liver fibrosis and are currently undergoing clinical trials. The investigation of herbal drugs or isolated natural active compounds is particularly valuable, due to their multifaceted functions and low [...] Read more.
Numerous approaches targeting hepatic stellate cells (HSCs) have emerged as pivotal therapeutic strategies to mitigate liver fibrosis and are currently undergoing clinical trials. The investigation of herbal drugs or isolated natural active compounds is particularly valuable, due to their multifaceted functions and low risk of side effects. Recent studies have hinted at the potential efficacy of verbascoside (VB) in ameliorating renal and lung fibrosis, yet its impact on hepatic fibrosis remains to be elucidated. This study aims to evaluate the potential effects of VB on liver fibrosis by assessing its ability to inhibit HSC activation. VB demonstrated significant efficacy in suppressing the expression of fibrogenic genes in activated LX-2 cells. Additionally, VB inhibited the migration and proliferation of these activated HSCs by scavenging reactive oxygen species (ROS) and downregulating the AMPK pathway. Furthermore, a biosafe reverse microemulsion loaded with VB (VB-ME) was developed to improve VB’s instability and low bioavailability. The optimal formulation of VB-ME was meticulously characterized, revealing substantial enhancements in cellular uptake, ROS-scavenging capacity, and the suppression of HSC activation. Full article
Show Figures

Figure 1

Figure 1
<p>VB suppresses mRNA levels of fibrogenic genes in activated LX-2 cells. (<b>A</b>) Chemical information of VB. (<b>B</b>) LX-2 cells were treated with VB at concentrations of 3.125, 6.25, 12.5, 25, 50, and 100 µM for 24 h, and cell viability was assessed using the MTT assay. The control group was treated with PBS. The data are presented as mean ± SEM (<span class="html-italic">n</span> = 4), with ‘ns’ indicating no significant differences. (<b>C</b>) LX-2 cells were pretreated with VB at concentrations of 0, 50, or 100 μM for 6 h, followed by treatment with TGFβ1 (5 ng/mL) for 24 h. mRNA expression levels of fibrogenic genes, including αSMA, COL1A1, COL1A2, FN1, and PDGFβ, were measured using qPCR. Inactive LX-2 cells served as the control group. The data presented are representative of three independent replicate experiments. Data are presented as mean ± SEM (<span class="html-italic">n</span> = 3). *: <span class="html-italic">p</span> &lt; 0.05, **: <span class="html-italic">p</span> &lt; 0.01, ***: <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 2
<p>VB inhibits the protein expression of fibrogenic genes at the protein level. LX-2 cells were pretreated with VB at different concentrations for 6 h, followed by treatment with TGFβ1 (5 ng/mL) for 24 h. Inactive LX-2 cells served as the control group. (<b>A</b>) Confocal images of LX-2 cells stained with COL1A1 antibody and DAPI. Scale bar, 20 μm. (<b>B</b>) Western blot assay images depicting protein levels of fibrogenic genes, including COL1A1 and FN1, in LX-2 cells. (<b>C</b>) Quantitative analysis of fluorescence levels in (<b>A</b>). (<b>D</b>) Quantitative analysis of the bands of Western blot in (<b>C</b>). The data presented are representative of three independent replicate experiments. Data are presented as mean ± SEM (<span class="html-italic">n</span> = 3). *: <span class="html-italic">p</span> &lt; 0.05, **: <span class="html-italic">p</span> &lt; 0.01, ***: <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 3
<p>VB scavenges ROS in activated LX-2 cells and downregulates the MAPK pathway. (<b>A</b>) LX-2 cells were pretreated with VB at different concentrations for 6 h, followed by treatment with TGFβ1 (5 ng/mL) for 24 h. Inactive LX-2 cells served as the control group. Intracellular ROS levels were determined using the DCFH-DA assay. Green represents intracellular ROS, and blue represents the cell nucleus. Scale bar, 100 µM. (<b>B</b>) Western blot assay images depicting protein levels of p38, JNK, and ERK. (<b>C</b>) The fluorescence intensity of DCFH-DA was normalized to the control group. (<b>D</b>) Quantitative analysis of the bands of Western blot in (<b>B</b>). (<b>E</b>) Intracellular SOD activities were determined using a superoxide dismutase assay kit. (<b>F</b>) The mRNA levels of Nrf2 in LX-2 cells were measured under different treatment conditions. LX-2 cells were treated as described in (<b>A</b>). In the VB group, cells were treated with 100 μM VB; in the ML385 group, cells were treated with 5 μM ML385; in the VB + ML385 group, cells were treated with 100 μM VB and 5 μM ML385. The data presented are representative of three independent replicate experiments. Data are presented as mean ± SEM (<span class="html-italic">n</span> = 3), *: <span class="html-italic">p</span> &lt; 0.05, **: <span class="html-italic">p</span> &lt; 0.01, ***: <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 4
<p>Development and characterization of VB-ME. (<b>A</b>) Pseudo-ternary phase diagram of the microemulsion system: IPM/Water/SoyPC in EtOH at 25 °C. IPM: isopropyl myristate, SoyPC: soy phosphatidylcholine, EtOH: ethanol. Blue area presents microemulsion zone. (<b>B</b>) The visual image of VB-ME. (<b>C</b>) The droplet-size distribution of VB-ME. (<b>D</b>) The averaged hydrodynamic diameter, PDI, and zeta potential of VB-ME. (<b>E</b>) The infrared spectrograms of isopropyl myristate, phosphatidylcholine (soy), VB, blank microemulsion, and VB microemulsion. The data presented are representative of three independent replicate experiments.</p>
Full article ">Figure 5
<p>VB-ME increased cellular uptake in Caco-2 cells. (<b>A</b>) Caco2 cells were treated with FITC-dextran (F-dextran), FITC-dextran mixed with VB solution (mix group), or F-dextran co-loaded with VB-ME (F-VB-ME group) for 2 h. Fluorescence signals were observed using a fluorescence microscope. Scale bar, 100 µM. (<b>B</b>) The fluorescence intensity was normalized to the F-dextran group. (<b>C</b>) The stability of VB-ME and VB in PBS (pH 7.4) solution at 37 °C for 24 h. The relative degradation level was tested using the HPLC method. The data presented are representative of three independent replicate experiments. Data are presented as mean ± SEM (<span class="html-italic">n</span> = 5), *: <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 6
<p>VB-ME enhanced the suppression of fibrogenic gene expression and migration in activated LX-2 cells. LX-2 cells were pretreated with either VB or VB-ME at a concentration of 50 μM VB for 6 h, followed by treatment with TGFβ1 (5 ng/mL) for 24 h. (<b>A</b>) The mRNA expression levels of fibrogenic genes, including αSMA, COL1A1, COL1A2, and PDGFβ, were measured using the qPCR method. Inactive LX-2 cells served as the control group. The data presented are representative of three independent replicate experiments. Data are presented as mean ± SEM (<span class="html-italic">n</span> = 3). *: <span class="html-italic">p</span> &lt; 0.05, **: <span class="html-italic">p</span> &lt; 0.01, ***: <span class="html-italic">p</span> &lt; 0.001. (<b>B</b>) A scratch wound healing assay was performed after the same treatment process as in the previous figure. The red line indicates the scratched cell edge, and the blue arrow shows the range of cell migration. Scale bar, 500 µM. (<b>C</b>) The cell migration level was calculated as (Range (0 h) − Range (12 h))/Range (0 h), and the result was normalized to the TGFβ-treated group. The data presented are representative of three independent replicate experiments. Data are presented as mean ± SEM (<span class="html-italic">n</span> = 4), ***: <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 7
<p>VB-ME enhanced the inhibition of proliferation in activated LX-2 cells. (<b>A</b>) Following the same treatment process as in the previous figure, cell proliferation levels were assessed using the Ki67 assay. Ki67 protein is shown in red, and nuclear staining with DAPI is in blue. Scale bar, 100 µM. (<b>B</b>) The relative proliferation rate was calculated by dividing the number of Ki67-positive cells by the total cell count, and the result was normalized to the TGFβ-treated group. The data presented are representative of three independent replicate experiments. Data are presented as mean ± SEM (<span class="html-italic">n</span> = 10), **: <span class="html-italic">p</span> &lt; 0.01, ***: <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 8
<p>VB-ME enhanced the scavenging of ROS in LX-2 cells. LX-2 cells were pretreated with VB at different concentrations for 6 h, followed by treatment with TGFβ1 (5 ng/mL) for 24 h. Inactive LX-2 cells served as the control group. (<b>A</b>) Intracellular ROS levels were determined using the DCFH-DA assay. Green represents intracellular ROS, and blue represents the cell nucleus. Scale bar, 100 µM. (<b>B</b>) The fluorescence intensity of DCFH-DA was normalized to the control group. The data presented are representative of three independent replicate experiments. Data are presented as mean ± SEM (<span class="html-italic">n</span> = 5), *: <span class="html-italic">p</span> &lt; 0.05, ***: <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">
33 pages, 2382 KiB  
Review
Liver Fibrosis: From Basic Science towards Clinical Progress, Focusing on the Central Role of Hepatic Stellate Cells
by Hikmet Akkız, Robert K. Gieseler and Ali Canbay
Int. J. Mol. Sci. 2024, 25(14), 7873; https://doi.org/10.3390/ijms25147873 - 18 Jul 2024
Cited by 1 | Viewed by 1281
Abstract
The burden of chronic liver disease is globally increasing at an alarming rate. Chronic liver injury leads to liver inflammation and fibrosis (LF) as critical determinants of long-term outcomes such as cirrhosis, liver cancer, and mortality. LF is a wound-healing process characterized by [...] Read more.
The burden of chronic liver disease is globally increasing at an alarming rate. Chronic liver injury leads to liver inflammation and fibrosis (LF) as critical determinants of long-term outcomes such as cirrhosis, liver cancer, and mortality. LF is a wound-healing process characterized by excessive deposition of extracellular matrix (ECM) proteins due to the activation of hepatic stellate cells (HSCs). In the healthy liver, quiescent HSCs metabolize and store retinoids. Upon fibrogenic activation, quiescent HSCs transdifferentiate into myofibroblasts; lose their vitamin A; upregulate α-smooth muscle actin; and produce proinflammatory soluble mediators, collagens, and inhibitors of ECM degradation. Activated HSCs are the main effector cells during hepatic fibrogenesis. In addition, the accumulation and activation of profibrogenic macrophages in response to hepatocyte death play a critical role in the initiation of HSC activation and survival. The main source of myofibroblasts is resident HSCs. Activated HSCs migrate to the site of active fibrogenesis to initiate the formation of a fibrous scar. Single-cell technologies revealed that quiescent HSCs are highly homogenous, while activated HSCs/myofibroblasts are much more heterogeneous. The complex process of inflammation results from the response of various hepatic cells to hepatocellular death and inflammatory signals related to intrahepatic injury pathways or extrahepatic mediators. Inflammatory processes modulate fibrogenesis by activating HSCs and, in turn, drive immune mechanisms via cytokines and chemokines. Increasing evidence also suggests that cellular stress responses contribute to fibrogenesis. Recent data demonstrated that LF can revert even at advanced stages of cirrhosis if the underlying cause is eliminated, which inhibits the inflammatory and profibrogenic cells. However, despite numerous clinical studies on plausible drug candidates, an approved antifibrotic therapy still remains elusive. This state-of-the-art review presents cellular and molecular mechanisms involved in hepatic fibrogenesis and its resolution, as well as comprehensively discusses the drivers linking liver injury to chronic liver inflammation and LF. Full article
(This article belongs to the Special Issue Molecular Advances in Liver Fibrosis)
Show Figures

Figure 1

Figure 1
<p>Overview of the cellular and molecular mechanisms of liver fibrogenesis. During chronic liver injury, hepatocytes activate signaling via Janus kinase (JNK), Notch, osteopontin, and hedgehog and produce exosomes harboring microRNAs (miRNAs) to initiate HSC activation. Inflammation triggers KCs and recruits monocyte-derived macrophages through C-C motif chemokine receptor (CCR)9 and C-C motif chemokine ligand (CCL)2, CCl<sub>4</sub>, and CCL25. The crosstalk between C-X3-C motif chemokine ligand 1 (CX3CL1) and C-X3-C motif chemokine receptor 1 (CX3CR1) orchestrates macrophage survival, differentiation, and polarization. KCs trigger the HSC activation by TGF-β, PDGF, and IL-1-β. Activated HSCs produce ECM proteins and secrete inflammatory chemokines CCL2, CCL3, and CX3CL1, whereby accumulating proinflammatory monocytes. HSC-originated matrix metalloproteinase (MMP) and tissue inhibitor of metalloproteinase (TIMP) contribute to ECM perpetuation, remodeling, and fibrosis. Activated HSCs lead to portal hypertension by enhancing the hepatic sinusoids’ contractility. Some molecules and pathways, including endothelin 1, TGF-β, Jak2, and the Wnt/β/catenin pathway, affect sinusoidal contractility.</p>
Full article ">Figure 2
<p>Molecular pathways and cellular interactions involved in HSC activation and deactivation. Activated HSCs are the main effector cells during hepatic fibrosis. In the healthy liver, they metabolize and store retinoids. Upon activation by fibrogenic stimuli, quiescent HSCs transdifferentiate into myofibroblasts, lose their vitamin A, upregulate α-smooth muscle actin (αSMA), and produce collagen I. Various factors, including immune cell-derived fibrogenic molecules, growth factors, and lipopolysaccharide, as well as profibrotic lipid mediators such as lysophosphatidylinositol and lysophosphatidic acid, induce HSC activation in the course of chronic liver disease. TGF-β is the most HSC potent activator, which is produced by infiltrating lymphocytes and monocytes, Kupffer cells (KCs), and damaged hepatocytes. IL-17, produced by neutrophils and Th17 cells, sensitizes HSCs to TGF-β by upregulating TGF-β receptor II (TGF-βRII). In addition, platelet-derived growth factor (PDGF), which is produced by endothelial cells and macrophages, further promotes HSC activation. During fibrosis resolution, HSCs either die or revert to an inactive state by upregulating transcription factors such as peroxisome proliferator-activated receptor-γ (PPARγ), GATA-binding factor 4 (GATA4), GATA6, and transcription factor 21 (TCF21). NK and CD8<sup>+</sup> T cells can eliminate activated HSCs by inducing apoptosis (Further abbreviations: GM-CSF, granulocyte/macrophage colony-stimulating factor; HH, hedgehog ligands; IHH, Indian Hedgehog; LPA, lysophosphatidic acid; LPI, lysophosphatidylinositol; LPS, lipopolysaccharide; miRNA, microRNA; MSR1, macrophage scavenger receptor 1; NF-κB, nuclear factor κ-light chain-enhancer of activated B cells; OPN, osteopontin; oxLDL, oxidized low-density lipoprotein; ROS, reactive oxygen species; S1P, sphingosine-1-phosphate; SHH, sonic hedgehog; TLR4, Toll-like receptor 4). Modified from reference [<a href="#B5-ijms-25-07873" class="html-bibr">5</a>].</p>
Full article ">Figure 3
<p>The path towards liver fibrosis: Kupffer cell activation and macrophage recruitment in the chronic inflammatory microenvironment of the diseased liver. (<b>A</b>) Ingestion of fat-laden apoptotic hepatocytes and free cholesterol activates KCs by promoting the production of proinflammatory mediators. (<b>B</b>) The liver’s chronic inflammatory microenvironment recruits monocytes from the circulation, which, due to local proinflammatory signaling, differentiate into monocyte-derived KC-like inflammatory, as well as lipid-associated, macrophages. (<b>C</b>) Macrophage populations are the major contributors in shaping both profibrotic and antifibrotic drivers within the fibrotic niche. Relevant phenotypic markers of the macrophage populations detected in mouse models are indicated in the figure (Abbreviations: CEACAM1, carcinoembryonic antigen-related cell adhesion molecule 1; CLEC4F, C-type lectin domain family 4 member F; LAM, lipid-associated macrophage; Mac1, macrophage-1 antigen; Mar1, macrophage scavenger receptor1; MMP, matrix metalloproteinase; SAM, scar-associated macrophages; SatM, segregated nucleus-containing atypical monocytes; TGF-β, transforming growth factor-β; TNF-α, tumor necrosis factor-α; VSIG4, V-set and immunoglobulin domain containing 4).</p>
Full article ">
11 pages, 2117 KiB  
Article
Depletion of Activated Hepatic Stellate Cells and Capillarized Liver Sinusoidal Endothelial Cells Using a Rationally Designed Protein for Nonalcoholic Steatohepatitis and Alcoholic Hepatitis Treatment
by Falguni Mishra, Yi Yuan, Jenny J. Yang, Bin Li, Payton Chan and Zhiren Liu
Int. J. Mol. Sci. 2024, 25(13), 7447; https://doi.org/10.3390/ijms25137447 - 6 Jul 2024
Viewed by 1136
Abstract
Nonalcoholic steatohepatitis (NASH) and alcoholic hepatitis (AH) affect a large part of the general population worldwide. Dysregulation of lipid metabolism and alcohol toxicity drive disease progression by the activation of hepatic stellate cells and the capillarization of liver sinusoidal endothelial cells. Collagen deposition, [...] Read more.
Nonalcoholic steatohepatitis (NASH) and alcoholic hepatitis (AH) affect a large part of the general population worldwide. Dysregulation of lipid metabolism and alcohol toxicity drive disease progression by the activation of hepatic stellate cells and the capillarization of liver sinusoidal endothelial cells. Collagen deposition, along with sinusoidal remodeling, alters sinusoid structure, resulting in hepatic inflammation, portal hypertension, liver failure, and other complications. Efforts were made to develop treatments for NASH and AH. However, the success of such treatments is limited and unpredictable. We report a strategy for NASH and AH treatment involving the induction of integrin αvβ3-mediated cell apoptosis using a rationally designed protein (ProAgio). Integrin αvβ3 is highly expressed in activated hepatic stellate cells (αHSCs), the angiogenic endothelium, and capillarized liver sinusoidal endothelial cells (caLSECs). ProAgio induces the apoptosis of these disease-driving cells, therefore decreasing collagen fibril, reversing sinusoid remodeling, and reducing immune cell infiltration. The reversal of sinusoid remodeling reduces the expression of leukocyte adhesion molecules on LSECs, thus decreasing leukocyte infiltration/activation in the diseased liver. Our studies present a novel and effective approach for NASH and AH treatment. Full article
(This article belongs to the Special Issue Chronic Liver Disease and Hepatocellular Carcinoma)
Show Figures

Figure 1

Figure 1
<p>ProAgio decreases hepatic fibrosis in NASH mice. (<b>A</b>) The scheme illustrates the ProAgio or vehicle treatment regimen of the NASH mice. (<b>B</b>,<b>C</b>) Representative images (<b>B</b>) and quantifications (<b>C</b>) of Sirius Red staining of liver sections from the treated mice. ET, early treatment; LT late treatment. (<b>D</b>,<b>E</b>) Representative images (<b>D</b>) and quantifications (<b>E</b>) of IHC staining of αSMA in liver sections from the treated mice. Quantitation of collagen by Sirius Red and αSMA levels by IHC of αSMA staining using Fiji software version 2.14.0/1.54f. Four randomly selected tissue sections per animal, three randomly selected view fields in each section, and six randomly selected animals (n = 6) were quantified. The quantity of collagen and αSMA levels are presented as % of staining positive area. (<b>F</b>) Blood glucose levels (mM/L) of five mice (n = 5, showing in different color) treated with the indicated agents were measured at indicated time points before (time 0) and after i.v. injection of 2 g/kg glucose. The mice were fasted overnight before the glucose injection and measurements. (<b>G</b>) Representative images of H&amp;E-stained liver sections from NASH mice treated with the indicated agents. Hepatic ballooning was scored by a hepatic pathologist based on H&amp;E staining (see the arrows in (<b>G</b>) for examples; 5 randomly selected sections per animal were scored n = 5): normal and ProAgio-treated animals score 0; vehicle-treated animals score 2 (number in each panel). The error bars in (<b>C</b>,<b>E</b>) are the standard deviations of five independent mice. Statistical analysis of data was performed by a one-way Student’s <span class="html-italic">t</span>-test. (* <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>ProAgio reduces inflammation in NASH mouse liver. (<b>A</b>,<b>B</b>) Representative images (<b>A</b>) and quantifications (<b>B</b>) of F4/80 staining of liver sections from mice treated with indicated agents. The quantity of total macrophage levels is presented as % of staining positive area. (<b>C</b>,<b>D</b>) Total population (%) of macrophages (<b>C</b>) and polymorphonuclear neutrophils (PMNs) (<b>D</b>) in liver tissues from NASH mice treated with indicated agents were analyzed by FACS. For macrophages in (<b>C</b>), CD45<sup>+</sup>/CD11b<sup>+</sup>/F4/80<sup>+</sup>/Ly-6C<sup>−</sup> cells were used. For PMN in (D), CD45<sup>+</sup>/CD11b<sup>+</sup>/Ly-6G<sup>+</sup> were used. (<b>E</b>,<b>F</b>) Representative images (<b>E</b>) and quantifications (<b>F</b>) of IHC staining of CD44 in liver sections from NASH mice treated with indicated agents. The quantity of CD44 levels is presented as % of staining positive area. Con in (<b>D</b>) and Controls in (<b>E</b>,<b>F</b>) means the mice were normal healthy mice without any disease induction and subsequent treatment. The error bars in (<b>B</b>–<b>D</b>,<b>F</b>) are standard deviations of measurements of 4 mice for (<b>B</b>,<b>C</b>) and 6 mice for (<b>D</b>,<b>F</b>). Statistical analysis of data was performed via a Student’s <span class="html-italic">t</span>-test for a two-group comparison by a one-way ANOVA with Tukey’s multiple comparison test. (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 3
<p>ProAgio decreases hepatic fibrosis in AH mice. (<b>A</b>) The scheme illustrates the ProAgio or vehicle treatment regimen for the AH mice. (<b>B</b>,<b>C</b>) Representative images (<b>B</b>) and quantifications (<b>C</b>) of Sirius Red staining of liver sections from mice treated with indicated agents. (<b>D</b>,<b>E</b>) Representative images (<b>D</b>) and quantifications (<b>E</b>) of IHC staining of αSMA in liver sections from the treated mice. (<b>F</b>,<b>G</b>) Representative images (<b>F</b>) and quantifications (<b>G</b>) of IHC staining of CD44 in liver sections from mice treated with indicated agents. The quantities in (<b>C</b>,<b>E</b>,<b>G</b>) are presented as % of staining positive area. Three randomly selected view fields per section, four randomly selected sections per animal, and six randomly selected animals (n = 6) were quantified. Veh in (<b>G</b>) is the vehicle-treated group. Control in (<b>G</b>) means the mice were normal healthy mice without disease induction and subsequent treatments. The error bars in (<b>C</b>,<b>E</b>,<b>G</b>) are standard deviations of 6 independent mice. Statistical analysis of data was performed by a Student’s <span class="html-italic">t</span>-test for a two-group comparison or a one-way ANOVA with Tukey’s multiple comparison test. (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 4
<p>ProAgio reduces inflammation in AH mouse liver. (<b>A</b>) The total population of neutrophils (Neutrophil%) in the liver tissues from the AH mice that were treated with indicated agents was analyzed by FACS (CD11b<sup>+</sup> Ly-6G<sup>+</sup>). (<b>B</b>,<b>C</b>) Representative images (<b>B</b>) and quantifications (<b>C</b>) of F4/80 staining of liver sections from mice treated with indicated agents. The quantity of total macrophage levels is presented as % of staining positive area. (<b>D</b>,<b>E</b>) Representative images (<b>D</b>) and quantifications (<b>E</b>) of IHC staining of SE-1 in liver sections from the NASH and AH mice that were treated with indicated agents. The quantity of differentiated healthy LSEC levels in (<b>E</b>) is presented as % of SE-1-positive staining area. The error bars in (<b>A</b>,<b>C</b>,<b>E</b>) are standard deviations of measurements of 6 mice. Statistical analysis of data was performed by a one-way ANOVA with Tukey’s multiple comparison test or an unpaired Student’s <span class="html-italic">t</span>-test for two-group comparisons. (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 5
<p>ProAgio decreases immune cell attachment molecules in NASH and AH mice (<b>A</b>–<b>D</b>) Representative images (<b>A</b>,<b>C</b>) and quantifications (<b>B</b>,<b>D</b>) of IHC staining of LYVE-1 (<b>A</b>,<b>B</b>) or VAP-1 (<b>C</b>,<b>D</b>) of liver sections from NASH or AH mice treated with indicated agents. The total LYVE-1 or VAP-1 levels are presented as % of staining positive area. Con and control mean the sections from normal healthy mice without any disease induction or subsequent treatment. (<b>E</b>) Scheme illustrating the drug actions of ProAgio in steatohepatitis. The error bars in (<b>B</b>,<b>D</b>) are standard deviations of measurements of 6 mice. Statistical analysis of data was performed by a one-way ANOVA with Tukey’s multiple comparison test. (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">
20 pages, 676 KiB  
Review
Exploring Fibrosis Pathophysiology in Lean and Obese Metabolic-Associated Fatty Liver Disease: An In-Depth Comparison
by Milena Vesković, Milka Pejović, Nikola Šutulović, Dragan Hrnčić, Aleksandra Rašić-Marković, Olivera Stanojlović and Dušan Mladenović
Int. J. Mol. Sci. 2024, 25(13), 7405; https://doi.org/10.3390/ijms25137405 - 5 Jul 2024
Viewed by 1196
Abstract
While obesity-related nonalcoholic fatty liver disease (NAFLD) is linked with metabolic dysfunctions such as insulin resistance and adipose tissue inflammation, lean NAFLD more often progresses to liver fibrosis even in the absence of metabolic syndrome. This review aims to summarize the current knowledge [...] Read more.
While obesity-related nonalcoholic fatty liver disease (NAFLD) is linked with metabolic dysfunctions such as insulin resistance and adipose tissue inflammation, lean NAFLD more often progresses to liver fibrosis even in the absence of metabolic syndrome. This review aims to summarize the current knowledge regarding the mechanisms of liver fibrosis in lean NAFLD. The most commonly used lean NAFLD models include a methionine/choline-deficient (MCD) diet, a high-fat diet with carbon tetrachloride (CCl4), and a high-fructose and high-cholesterol diet. The major pro-fibrogenic mechanisms in lean NAFLD models include increased activation of the extracellular signal-regulated kinase (ERK) pathway, elevated expression of α-smooth muscle actin (α-SMA), collagen type I, and TGF-β, and modulation of fibrogenic markers such as tenascin-X and metalloproteinase inhibitors. Additionally, activation of macrophage signaling pathways promoting hepatic stellate cell (HSC) activation further contributes to fibrosis development. Animal models cannot cover all clinical features that are evident in patients with lean or obese NAFLD, implicating the need for novel models, as well as for deeper comparisons of clinical and experimental studies. Having in mind the prevalence of fibrosis in lean NAFLD patients, by addressing specific pathways, clinical studies can reveal new targeted therapies along with novel biomarkers for early detection and enhancement of clinical management for lean NAFLD patients. Full article
(This article belongs to the Special Issue Exploring Molecular Mechanisms of Liver Fibrosis)
Show Figures

Figure 1

Figure 1
<p>Insulin resistance in MAFLD pathogenesis. In obesity, dysfunctional adipocytes increase circulating free fatty acids, which are taken up by the liver, leading to triglyceride accumulation and exacerbation of hepatic insulin resistance. Hepatic insulin resistance triggers liver inflammation through proinflammatory cytokines, oxidative stress due to excess reactive oxygen species, and ER stress from protein misfolding, all of which disrupt cellular functions. These stress responses, along with apoptosis and reduced autophagy, cause hepatocytes to release hepatokines, further aggravating systemic insulin resistance and creating a feedback loop that worsens liver dysfunction.</p>
Full article ">
14 pages, 1868 KiB  
Article
Investigating the Antifibrotic Effects of β-Citronellol on a TGF-β1-Stimulated LX-2 Hepatic Stellate Cell Model
by Watunyoo Buakaew, Sucheewin Krobthong, Yodying Yingchutrakul, Pachuen Potup, Yordhathai Thongsri, Krai Daowtak, Antonio Ferrante and Kanchana Usuwanthim
Biomolecules 2024, 14(7), 800; https://doi.org/10.3390/biom14070800 - 5 Jul 2024
Viewed by 1040
Abstract
Liver fibrosis, a consequence of chronic liver damage or inflammation, is characterized by the excessive buildup of extracellular matrix components. This progressive condition significantly raises the risk of severe liver diseases like cirrhosis and hepatocellular carcinoma. The lack of approved therapeutics underscores the [...] Read more.
Liver fibrosis, a consequence of chronic liver damage or inflammation, is characterized by the excessive buildup of extracellular matrix components. This progressive condition significantly raises the risk of severe liver diseases like cirrhosis and hepatocellular carcinoma. The lack of approved therapeutics underscores the urgent need for novel anti-fibrotic drugs. Hepatic stellate cells (HSCs), key players in fibrogenesis, are promising targets for drug discovery. This study investigated the anti-fibrotic potential of Citrus hystrix DC. (KL) and its bioactive compound, β-citronellol (β-CIT), in a human HSC cell line (LX-2). Cells exposed to TGF-β1 to induce fibrogenesis were co-treated with crude KL extract and β-CIT. Gene expression was analyzed by real-time qRT-PCR to assess fibrosis-associated genes (ACTA2, COL1A1, TIMP1, SMAD2). The release of matrix metalloproteinase 9 (MMP-9) was measured by ELISA. Proteomic analysis and molecular docking identified potential signaling proteins and modeled protein–ligand interactions. The results showed that both crude KL extract and β-CIT suppressed HSC activation genes and MMP-9 levels. The MAPK signaling pathway emerged as a potential target of β-CIT. This study demonstrates the ability of KL extract and β-CIT to inhibit HSC activation during TGF-β1-induced fibrogenesis, suggesting a promising role of β-CIT in anti-hepatic fibrosis therapies. Full article
Show Figures

Figure 1

Figure 1
<p>Dose–response curve of crude KL extract and β-citronellol on viability of the LX-2 cells.</p>
Full article ">Figure 2
<p>Crude KL extract and β-CIT attenuate the expression of hepatic-fibrosis-associated genes and mitigate MMP-9 production in LX-2 cells challenged with TGF-β1. Following a 24 h co-treatment protocol, mRNA was extracted and subjected to real-time qRT-PCR analysis. Expression levels were normalized to the housekeeping gene <span class="html-italic">GAPDH</span> (<b>A</b>), revealing a marked downregulation of fibrosis-related genes. Furthermore, a statistically significant reduction in MMP-9 levels was observed in the cell culture supernatant (<b>B</b>). Statistical significance was denoted as follows: <span class="html-italic">p</span> &lt; 0.0332 (*), <span class="html-italic">p</span> &lt; 0.0002 (***), <span class="html-italic">p</span> &lt; 0.0001 (****).</p>
Full article ">Figure 3
<p>The functional enrichment analysis of DEPs. Quantitative LC-MS/MS analysis yielded a total of 1570 differentially expressed proteins (DEPs) (<b>A</b>). Volcano plots visually demonstrated the upregulated (n = 125, red) and downregulated (n = 65, green) proteins within this dataset (<b>B</b>). To elucidate potential functional relationships, protein–protein interaction (PPI) network analyses were performed separately for the upregulated (<b>C</b>) and downregulated (<b>D</b>) protein subsets. Gene ontology (GO) annotations illuminated key biological processes associated with both upregulated (<b>E</b>) and downregulated (<b>F</b>) DEPs.</p>
Full article ">Figure 4
<p>The KEGG analysis of DEPs. Amongst the most significantly enriched KEGG signaling pathways, differential protein expression patterns emerged (upregulated: <b>A</b>; downregulated: <b>B</b>). Notably, the MAPK signaling pathway exhibited a pronounced association with downregulated proteins (<b>C</b>; indicated in blue).</p>
Full article ">Figure 5
<p>The 2D and 3D structures of candidate target proteins and β-CIT interaction.</p>
Full article ">
16 pages, 14920 KiB  
Article
miR-9-5p and miR-221-3p Promote Human Mesenchymal Stem Cells to Alleviate Carbon Tetrachloride-Induced Liver Injury by Enhancing Human Mesenchymal Stem Cell Engraftment and Inhibiting Hepatic Stellate Cell Activation
by Lihong He, Jianwei Xu, Ping Huang, Yu Bai, Huanhuan Chen, Xiaojing Xu, Ya’nan Hu, Jinming Liu and Huanxiang Zhang
Int. J. Mol. Sci. 2024, 25(13), 7235; https://doi.org/10.3390/ijms25137235 - 30 Jun 2024
Viewed by 738
Abstract
Mesenchymal stem cells (MSCs) have shown great potential for the treatment of liver injuries, and the therapeutic efficacy greatly depends on their homing to the site of injury. In the present study, we detected significant upregulation of hepatocyte growth factor (HGF) in the [...] Read more.
Mesenchymal stem cells (MSCs) have shown great potential for the treatment of liver injuries, and the therapeutic efficacy greatly depends on their homing to the site of injury. In the present study, we detected significant upregulation of hepatocyte growth factor (HGF) in the serum and liver in mice with acute or chronic liver injury. In vitro study revealed that upregulation of miR-9-5p or miR-221-3p promoted the migration of human MSCs (hMSCs) toward HGF. Moreover, overexpression of miR-9-5p or miR-221-3p promoted hMSC homing to the injured liver and resulted in significantly higher engraftment upon peripheral infusion. hMSCs reduced hepatic necrosis and inflammatory infiltration but showed little effect on extracellular matrix (ECM) deposition. By contrast, hMSCs overexpressing miR-9-5p or miR-221-3p resulted in not only less centrilobular necrosis and venous congestion but also a significant reduction of ECM deposition, leading to obvious improvement of hepatocyte morphology and alleviation of fibrosis around central vein and portal triads. Further studies showed that hMSCs inhibited the activation of hepatic stellate cells (HSCs) but could not decrease the expression of TIMP-1 upon acute injury and the expression of MCP-1 and TIMP-1 upon chronic injury, while hMSCs overexpressing miR-9-5p or miR-221-3p led to further inactivation of HSCs and downregulation of all three fibrogenic and proinflammatory factors TGF-β, MCP-1, and TIMP-1 upon both acute and chronic injuries. Overexpression of miR-9-5p or miR-221-3p significantly downregulated the expression of α-SMA and Col-1α1 in activated human hepatic stellate cell line LX-2, suggesting that miR-9-5p and miR-221-3p may partially contribute to the alleviation of liver injury by preventing HSC activation and collagen expression, shedding light on improving the therapeutic efficacy of hMSCs via microRNA modification. Full article
Show Figures

Figure 1

Figure 1
<p>miR-9-5p and miR-221-3p promote the migration of hMSCs toward HGF in vitro. After infection with Ad, Ad-9, or Ad-221 or transfection with negative control (NC), miR-9-5p mimic, or miR-221-3p mimic, hMSCs were tested for migration. (<b>a</b>) Transfilter migration of hMSCs toward HGF using a Boyden chamber. HGF (50 ng/mL) was added to the lower compartment of the Boyden chamber. Cells were allowed to migrate toward HGF for 6 h at 37 °C in a 5% CO<sub>2</sub> humidified incubator. Left panel: Representative images of migratory cells/field on the membrane underside, scale bar = 250 μm. Right panel: Quantitative results of cell transfilter migration induced by HGF. Data are shown as mean ± SEM from three independent experiments (** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, compared with Ad or NC). (<b>b</b>) miR-9-5p and miR-221-3p promoted motility but not migration persistence of hMSCs in response to an HGF gradient using a Dunn chamber. Upper panel: hMSCs over the annular bridge between the inner and outer wells of the chamber observed under phase-contrast and fluorescence optics. Cell migration was recorded continuously by time-lapse frame grabbing for 6 h in humidified 5% CO<sub>2</sub> at 37 °C. Scale bar = 250 μm. Lower panel: Quantitative results of total distance (μm), migration velocity (μm/min), and the forward migration index (FMI). Data are shown as mean ± SEM from three independent experiments (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, compared with Ad).</p>
Full article ">Figure 2
<p>Engraftment of hMSCs after infusion in mice with acute or chronic liver injuries for 7 d. hMSCs that were infected with Ad, Ad-9, or Ad-221, or transfected with NC, miR-9-5p mimic, or miR-221-3p mimic and labeled with H33342 were transplanted to mice with acute (<b>a</b>) or chronic (<b>b</b>) liver injuries for 7 d (<span class="html-italic">n</span> = 5 per group). Representative images (scale bar = 250 μm) with GFP-positive or H33342-labeled transplanted cells are shown. Arrows show the engraftment of transplanted cells. GFP<sup>+</sup> cell number and fluorescence intensity of H33342 were quantified using ImageJ software 1.54. Data are shown as mean ± SEM from at least three independent experiments (** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, compared with Ad or NC).</p>
Full article ">Figure 3
<p>hMSCs overexpressing miR-9-5p or miR-221-3p improved hepatic microscopic architecture. Healthy mice or mice with CCl<sub>4</sub>-induced acute (<b>a</b>) and chronic (<b>b</b>) liver injury that received PBS or hMSCs for 7 d were cryosectioned and stained with hematoxylin and eosin (<span class="html-italic">n</span> = 5 per group). Asterisks show zonal centrilobular hepatocyte necrosis. Scale bar = 100 μm.</p>
Full article ">Figure 4
<p>hMSCs overexpressing miR-9-5p or miR-221-3p reduced ECM deposition after peripheral infusion for 7 d. Healthy mice and mice with chronic liver injury received PBS or hMSCs for 7 d (<span class="html-italic">n</span> = 5 per group). (<b>a</b>) Representative images of ECM deposition in different groups using Masson’s trichrome staining (scale bar = 100 μm). Arrows indicate the accumulation of ECM. (<b>b</b>) Quantification of ECM area using ImageJ software 1.54. Data are shown as mean ± SEM of at least three independent experiments (** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, compared with PBS).</p>
Full article ">Figure 5
<p>hMSCs overexpressing miR-9-5p or miR-221-3p suppress HSC activation after peripheral infusion for 7 d. Shown are representative images of α-SMA-positive cells (green, scale bar = 100 μm) based on immunofluorescence staining of α-SMA in acutely (<b>a</b>) and chronically (<b>b</b>) injured livers (<span class="html-italic">n</span> = 5 per group). High magnification views in insets show the expression of α-SMA in the perisinusoidal space of hepatic parenchyma. α-SMA expression was quantified using ImageJ software 1.54. Data are shown as mean ± SEM from at least three independent experiments (* <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001, compared with PBS; # <span class="html-italic">p</span> &lt; 0.05, ## <span class="html-italic">p</span> &lt; 0.01, compared with Ad).</p>
Full article ">Figure 6
<p>Effects of hMSC-conditioned medium and overexpression of miR-9-5p or miR-221-3p on HSC activation and procollagen expression. Activated human hepatic stellate cell line LX-2 at 60–70% confluence was incubated in the conditioned medium of hMSCs overexpressing miR-9-5p or miR-221-3p (<b>a</b>) or directly transfected with miR-9-5p mimic or miR-221-3p mimic for 48 h (<b>b</b>), and then the mRNA level of α-SMA and Col-1α1 were quantified using qRT-PCR. Data are shown as mean ± SEM from at least three independent experiments (* <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, compared with NC).</p>
Full article ">Figure 7
<p>Transcriptional expression of fibrogenic and proinflammatory factors in the liver after infusion of hMSCs for 7 d. (<b>a</b>) In acutely injured livers (<span class="html-italic">n</span> = 5 per group). (<b>b</b>) In chronically injured livers (<span class="html-italic">n</span> = 5 per group). Expression was normalized to <span class="html-italic">GAPDH</span> and expressed as a fold change of Healthy. Data presented are the mean ± SEM from at least three independent experiments (* <span class="html-italic">p &lt;</span> 0.05, *** <span class="html-italic">p &lt;</span> 0.001, compared with PBS; # <span class="html-italic">p &lt;</span> 0.05, ## <span class="html-italic">p &lt;</span> 0.01, ### <span class="html-italic">p &lt;</span> 0.001, compared with Ad).</p>
Full article ">
8 pages, 2059 KiB  
Brief Report
Rilpivirine Activates STAT1 in Non-Parenchymal Cells to Regulate Liver Injury in People Living with HIV and MASLD
by Ángela B. Moragrega, Carmen Busca, Nadezda Apostolova, Antonio Olveira, Luz Martín-Carbonero, Eulalia Valencia, Victoria Moreno, José I. Bernardino, Marta Abadía, Juan González-García, Juan V. Esplugues, María L. Montes and Ana Blas-García
Biomedicines 2024, 12(7), 1454; https://doi.org/10.3390/biomedicines12071454 - 29 Jun 2024
Viewed by 674
Abstract
Liver fibrosis is a key determinant of the progression of metabolic dysfunction-associated steatotic liver disease (MASLD). Its increasing prevalence and a lack of effective treatments make it a major health problem worldwide, particularly in people living with HIV, among whom the prevalence of [...] Read more.
Liver fibrosis is a key determinant of the progression of metabolic dysfunction-associated steatotic liver disease (MASLD). Its increasing prevalence and a lack of effective treatments make it a major health problem worldwide, particularly in people living with HIV, among whom the prevalence of advanced fibrosis is higher. We have published preclinical data showing that Rilpivirine (RPV), a widely used anti-HIV drug, selectively triggers hepatic stellate cell (HSC) inactivation and apoptosis through signal transducer and activator of transcription (STAT)1-mediated pathways, effects that clearly attenuate liver fibrosis and promote regeneration. We performed a retrospective, cross-sectional study of RPV-induced effects on steatosis, inflammation, and fibrosis in liver biopsies from well-controlled HIV-infected subjects diagnosed with MASLD. Patients on RPV exhibited similar levels of HIV-related parameters to those not receiving this drug, while showing a tendency toward improved liver function and lipid profile, as well as an enhanced activation of STAT1 in hepatic non-parenchymal cells in those with identified liver injury. This protective effect, promoting STAT1-dependent HSC inactivation, was observed at different stages of MASLD. Our results suggest that RPV-based therapy is especially indicated in HIV-infected patients with MASLD-derived liver injury and highlight the potential of RPV as a new therapeutic strategy for liver diseases. Full article
(This article belongs to the Special Issue Advances in the Pathogenesis and Treatment of AIDS)
Show Figures

Figure 1

Figure 1
<p>Nuclear STAT1 expression in non-parenchymal cells of liver biopsies from people living with HIV with diagnosed metabolic dysfunction-associated steatotic liver disease (different disease groups: steatosis &gt; 30%, steatohepatitis and fibrosis &gt; 0). Representative images of STAT1 immunohistochemistry in hepatic sections from patients with identified liver injury receiving RPV-free or RPV-based therapy. Black arrows indicate positive non-parenchymal cells. Scale bar = 0.1 mm.</p>
Full article ">
13 pages, 5051 KiB  
Article
Optimizing the Amino Acid Sequence Enhances the Productivity and Bioefficacy of the RBP-Albumin Fusion Protein
by Ji Hoon Park, Sohyun Kwon, So-Young Choi, Bongcheol Kim and Junseo Oh
Bioengineering 2024, 11(6), 617; https://doi.org/10.3390/bioengineering11060617 - 17 Jun 2024
Viewed by 840
Abstract
The significant growth of the global protein drug market, including fusion proteins, emphasizes the crucial role of optimizing amino acid sequences to enhance the productivity and bioefficacy. Among these fusion proteins, RBP-IIIA-IB, comprising retinol-binding protein in conjunction with the albumin domains, IIIA and [...] Read more.
The significant growth of the global protein drug market, including fusion proteins, emphasizes the crucial role of optimizing amino acid sequences to enhance the productivity and bioefficacy. Among these fusion proteins, RBP-IIIA-IB, comprising retinol-binding protein in conjunction with the albumin domains, IIIA and IB, has displayed efficacy in alleviating liver fibrosis by inhibiting the activation of hepatic stellate cells (HSCs). This study aimed to address the issue of the low productivity in RBP-IIIA-IB. To induce structural changes, the linking sequence, EVDD, between domain IIIA and IB in RBP-IIIA-IB was modified to DGPG, AAAA, and GGPA. Among these, RBP-IIIA-AAAA-IB demonstrated an increase in yield (>4-fold) and a heightened inhibition of HSC activation. Furthermore, we identified amino acid residues that could form disulfide bonds when substituted with cysteine. Through the mutation of N453S-V480S in RBP-IIIA-AAAA-IB, the productivity further increased by over 9-fold, accompanied by an increase in anti-fibrotic activity. Overall, there was a more than 30-fold increase in the fusion protein’s yield. These findings demonstrate the effectiveness of modifying linker sequences and introducing extra disulfide bonds to improve both the production yield and biological efficacy of fusion proteins. Full article
(This article belongs to the Section Biochemical Engineering)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Low productivity of RBP-IIIA-IB in ExpiCHO cells. Following transient transfection of ExpiCHO cells with the plasmid encoding RBP-IIIA-IB, both the culture supernatant and the Ni Sepharose eluate were analyzed via SDS-PAGE (<b>A</b>) and Western blotting (<b>B</b>). RBP-IIIA-IB is indicated by rectangular box or arrow. S: supernatant, E: Ni-Sepharose eluate, R: reducing condition, NR-non-reducing condition, M: molecular weight marker, P: positive control for protein mass (P: 2 μg, P1: 500 ng).</p>
Full article ">Figure 2
<p>The predicted structure of RBP-IIIA-IB as determined by AlphaFold2. Within this structure, the domains RBP and albumin IIIA and IB are discernible, with the connecting region between RBP and IIIA highlighted by a white arrow. The binding pockets for retinoic acid in domains IIIA and IB are illustrated by white closed circles, and the region of the modified linker sequence is indicated with a red circle.</p>
Full article ">Figure 3
<p>The expression of RBP-IIIA-linker-IB inactivated hepatic stellate cells (HSCs). HSCs after passage 1 were transiently transfected with a plasmid encoding a fusion protein containing different linker sequences, and the levels of alpha-smooth muscle actin (α-SMA) and collagen type I were evaluated using real-time PCR. The data represent the means ± standard deviation of three independent experiments. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, paired <span class="html-italic">t</span>-test (compared with control cells).</p>
Full article ">Figure 4
<p>The protein expression levels of RBP-IIIA-linker-IB and IIIA-linker-IB-RBP in Expi293 cells. Expi293 cells were transiently transfected with a plasmid encoding the fusion protein containing various linking sequences, and an equal amount of culture supernatant was subjected to Western blot analysis using an anti-His tag antibody. Lane 1: positive control for protein mass, 2: RBP-IIIA-DGPG-IB, 3: RBP-IIIA-GGPA-IB, 4: RBP-IIIA-EVDD-IB, 5: RBP-IIIA-AAAA-IB, 6: IIIA-DGPG-IB-RBP, 7: IIIA-GGPA-IB-RBP, 8: IIIA-EVDD-IB-RBP, 9: IIIA-AAAA-IB-RBP, M: molecular weight marker.</p>
Full article ">Figure 5
<p>The predicted structure of RBP-IIIA-EVDD-IB (yellow) and RBP-IIIA-AAAA-IB (cyan) as determined by AlphaFold2. RBP. It is notable that the RBP component in RBP-IIIA-AAAA-IB shifts away from the RBP of RBP-IIIA-EVDD-IB, indicated by the white arrow.</p>
Full article ">Figure 6
<p>The expression of RBP-IIIA-AAAA-IB with an additional disulfide bond induced phenotypic changes in hepatic stellate cells (HSCs). HSCs after passage 1 were transiently transfected with a plasmid encoding the fusion protein featuring cysteine substitutions at T446-L487, N453-V480, V457-Y476, or V144-A199, and their morphology was observed using a light microscope. Scale bar, 20 μm.</p>
Full article ">Figure 7
<p>The expression of RBP-IIIA-AAAA-IB with an additional disulfide bond inactivated hepatic stellate cells (HSCs). HSCs after passage 1 were transiently transfected with a plasmid encoding the fusion protein featuring cysteine substitutions at T446-L487, N453-V480, V457-Y476, or V144-A199, and the levels of alpha-smooth muscle actin (α-SMA) and collagen type I expression were assessed using real-time PCR. * <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, paired <span class="html-italic">t</span>-test (<span class="html-italic">n</span> = 3) (compared with control cells).</p>
Full article ">Figure 8
<p>The synthesis of RBP-IIIA-AAAA-IB and RBP-IIIA-AAAA-IB_C453-480 in Expi293 cells. Expi293 cells were transiently transfected with the plasmid encoding either RBP-IIIA-AAAA-IB (1) or RBP-IIIA-AAAA-IB_C453-480 (2), and the resulting fusion proteins were purified through Ni Sepharose and size exclusion chromatography. SDS-PAGE analysis was performed on the purified fusion protein samples (<b>A</b>), with the corresponding size exclusion chromatography profile displayed (<b>B</b>). M, molecular weight marker.</p>
Full article ">Figure 9
<p>The increased production of RBP-IIIA-AAAA-IB_C453-480 in ExpiCHO cells. After ExpiCHO cells were transiently transfected with the plasmid encoding RBP-IIIA-AAAA-IB_C453-480, both the culture supernatant (<b>A</b>) and the purified fusion protein (<b>B</b>) were analyzed using SDS-PAGE. M, molecular weight marker; R, reduction; NR, non-reduction.</p>
Full article ">Figure 10
<p>The anti-fibrotic effects of the fusion proteins on hepatic stellate cells (HSCs). HSCs after passage 1 were treated with purified fusion proteins (white), RBP-IIIA-EVDD-IB (black) (0.75 or 0.375 μM), RBP-IIIA-AAAA-IB (hatched), or RBP-IIIA-AAAA-IB_C453-480 (gray), for 16 h, and the levels of alpha-smooth muscle actin (α-SMA) (<b>A</b>) and collagen type I (<b>B</b>) expression were evaluated using real-time PCR. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, paired <span class="html-italic">t</span>-test (<span class="html-italic">n</span> = 3) (compared with control cells).</p>
Full article ">Figure 11
<p>The predicted structure of RBP-IIIA-EVDD-IB (yellow), RBP-IIIA-AAAA-IB (cyan), and RBP-IIIA-AAAA-IB_C453-480 (pink) as determined by AlphaFold2. RBP. It is notable that RBP components in RBP-IIIA-AAAA-IB and RBP-IIIA-AAAA-IB_C453-480 gradually diverge from the RBP of RBP-IIIA-IB, indicated by the yellow arrow.</p>
Full article ">
25 pages, 3042 KiB  
Review
Liver Cell Mitophagy in Metabolic Dysfunction-Associated Steatotic Liver Disease and Liver Fibrosis
by Jiaxin Chen, Linge Jian, Yangkun Guo, Chengwei Tang, Zhiyin Huang and Jinhang Gao
Antioxidants 2024, 13(6), 729; https://doi.org/10.3390/antiox13060729 - 15 Jun 2024
Cited by 1 | Viewed by 1529
Abstract
Metabolic dysfunction-associated steatotic liver disease (MASLD) affects approximately one-third of the global population. MASLD and its advanced-stage liver fibrosis and cirrhosis are the leading causes of liver failure and liver-related death worldwide. Mitochondria are crucial organelles in liver cells for energy generation and [...] Read more.
Metabolic dysfunction-associated steatotic liver disease (MASLD) affects approximately one-third of the global population. MASLD and its advanced-stage liver fibrosis and cirrhosis are the leading causes of liver failure and liver-related death worldwide. Mitochondria are crucial organelles in liver cells for energy generation and the oxidative metabolism of fatty acids and carbohydrates. Recently, mitochondrial dysfunction in liver cells has been shown to play a vital role in the pathogenesis of MASLD and liver fibrosis. Mitophagy, a selective form of autophagy, removes and recycles impaired mitochondria. Although significant advances have been made in understanding mitophagy in liver diseases, adequate summaries concerning the contribution of liver cell mitophagy to MASLD and liver fibrosis are lacking. This review will clarify the mechanism of liver cell mitophagy in the development of MASLD and liver fibrosis, including in hepatocytes, macrophages, hepatic stellate cells, and liver sinusoidal endothelial cells. In addition, therapeutic strategies or compounds related to hepatic mitophagy are also summarized. In conclusion, mitophagy-related therapeutic strategies or compounds might be translational for the clinical treatment of MASLD and liver fibrosis. Full article
Show Figures

Figure 1

Figure 1
<p>Processes of mitophagy. The process of mitophagy can be divided into five steps: (1) Initiation: ATG8 family proteins are anchored in the inner and outer membranes of the phagophores. (2) Recognition: SMRs or MMRs recognize impaired mitochondria and link the mitochondria to phagophores via the ATG8-LIR interaction. (3) Sequestration: Impaired mitochondria are sequestered, forming mitophagosomes. (4) Fusion: Mitophagosomes further fuse with lysosomes to form mitolysosomes. (5) Degradation: Impaired mitochondria are degraded and recycled in mitolysosomes. Abbreviations: Atg, autophagy-related gene; SMRs, soluble mitophagy receptors; MMRs, membrane-attached mitophagy receptors; UBDs, ubiquitin-binding domains; LIR, microtubule-associated protein 1A/1B light chain 3 (LC3)-interacting region (LIR); OMM, outer mitochondrial membrane.</p>
Full article ">Figure 2
<p>Mitophagy signaling pathways. The mitophagy signaling pathways can be classified as PINK1/Parkin-dependent (<b>a</b>), PINK1/Parkin-independent (<b>b</b>), or other mitophagy signaling pathways (<b>c</b>). (<b>a</b>) In healthy mitochondria, PINK1 translocates into mitochondria through TOMM and TIMM and is proteolytically cleaved by PARL and MPP. Upon mitochondrial damage, PINK1 fails to enter and accumulates on the OMM, which recruits and phosphorylates Parkin. PINK1 and Parkin jointly generate p-Ub chains on the OMM. SMRs (e.g., p62, OPTN, and NDP52) recognize p-Ub signals and link damaged mitochondria to phagophores via the LIR-ATG8 interaction. (<b>b</b>) MMRs directly interact with lipidated ATG8. BNIP3, BNIP3L, and FUNDC1 are common in response to hypoxia. PHB2 localizes to the IMM. Upon proteasome-mediated rupture of the OMM, PHB2 interacts with ATG8. (<b>c</b>) Several ubiquitin ligases, MUL1, SIAH1, and ARIH1, ubiquitinate the OMM to promote mitophagy. Notably, MUL1 has an LIR motif that interacts with ATG8. Damaged mitochondria can be translocated to other cells (e.g., macrophages) through exospheres or EVs for mitophagy. MDVs derived from mitochondria fuse with lysosomes for degradation. Cardiolipin translocates from the IMM to the OMM to interact with ATG8. Abbreviations: TOMM/TIMM, translocase complex of the outer/inner mitochondrial membrane (OMM/IMM); PARL, presenilin-associated rhomboid-like protein; MPP, mitochondrial processing peptidase; p-Ub, phosphorylated ubiquitin; MDVs, mitochondria-derived vesicles; EVs, extracellular vesicles.</p>
Full article ">Figure 3
<p>Hepatocyte mitophagy in MASLD and liver fibrosis. Many regulators and signaling pathways are involved in hepatocyte mitophagy in MASLD and liver fibrosis. In most cases, damaged hepatocyte mitophagy leads to the aggravation of inflammation, the accumulation of dysfunctional mitochondria, enhanced oxidative stress, lipid accumulation, and apoptosis. Notably, there is a crosstalk between hepatocytes and macrophages. A reduction in Miz1 results in increased levels of free PRDX6. PRDX6 interacts with Parkin and blocks Parkin autoubiquitination as well as downstream OMM ubiquitination. Impaired mitophagy caused hepatocyte inflammasome activation and stimulated macrophage TNF-α production. TNF-α further promotes Miz1 degradation in hepatocytes, forming a vicious feedback loop. Created with <a href="http://BioRender.com" target="_blank">BioRender.com</a> (accessed on 29 May 2024).</p>
Full article ">Figure 4
<p>HSC mitophagy in MASLD and liver fibrosis. In response to CCl<sub>4</sub>, downregulated SIRT3 fails to specifically deacetylate PINK1 and NIPSNAP1, resulting in impaired mitophagy. During the process of CCl<sub>4</sub> withdrawal, downregulation of BCL-B promotes HSC apoptosis by inducing Parkin-mediated mitophagy. Hepatocytes under nutritional stress release exosomal miR-27a to inhibit PINK1-mediated mitophagy in HSCs, leading to the activation of HSCs. Abbreviations: ECM, extracellular matrix. Created with <a href="http://BioRender.com" target="_blank">BioRender.com</a> (accessed on 30 March 2024).</p>
Full article ">Figure 5
<p>Macrophage mitophagy in MASLD and liver fibrosis. During MASH progression, elevated PTPROt expression exacerbates inflammation. However, PTPROt enhances mitophagy to partially restrict ROS production and inflammation. In liver fibrosis, XBP1 binds directly to the <span class="html-italic">Bnip3</span> promoter, suppressing the transcription of <span class="html-italic">Bnip3</span>. Impaired BNIP3-mediated mitophagy ultimately leads to increased release of pro-inflammatory and pro-fibrotic cytokines in macrophages. High TIM-4 levels contribute to increased levels of PINK1 and Parkin and the induction of TGF-β1. Created with <a href="http://BioRender.com" target="_blank">BioRender.com</a> (accessed on 29 May 2024).</p>
Full article ">
Back to TopTop