[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 (109)

Search Parameters:
Keywords = glomerulus

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
12 pages, 1160 KiB  
Article
Dehydration and Suboptimal Sleep Aggravate Early Renal Impairment in Children: Longitudinal Findings from the PROC Study
by Menglong Li, Huidi Xiao, Nubiya Amaerjiang, Bipin Thapa, Wen Shu, Yeerlin Asihaer, Mengying Guan, Sten H. Vermund, Zhiyong Zou, Dayong Huang and Yifei Hu
Nutrients 2024, 16(20), 3472; https://doi.org/10.3390/nu16203472 - 14 Oct 2024
Viewed by 462
Abstract
Background: While dehydration is associated with pediatric renal impairment, the regulation of hydration status can be affected by sleep. However, the interaction of hydration and sleep on kidney health remains unclear. Methods: We conducted a cohort study among 1914 healthy primary school children [...] Read more.
Background: While dehydration is associated with pediatric renal impairment, the regulation of hydration status can be affected by sleep. However, the interaction of hydration and sleep on kidney health remains unclear. Methods: We conducted a cohort study among 1914 healthy primary school children from October 2018 to November 2019 in Beijing, China. Four-wave urinary β2-microglobulin and microalbumin excretion were assayed to assess transient renal tubular and glomerular impairment, and specific gravity was measured to determine hydration status with contemporaneous assessment of sleep duration, other anthropometric, and lifestyle covariates. We used generalized linear mixed-effects models to assess longitudinal associations of sleep duration and hydration status with renal impairment. Results: We observed 1378 children with optimal sleep (9–<11 h/d, 72.0%), 472 with short sleep (<9 h/d), and 64 with long sleep (≥11 h/d, 3.3%). Over half (55.4%) of events determined across 6968 person-visits were transient dehydration, 19.4% were tubular, and 4.9% were glomerular impairment events. Taking optimal sleep + euhydration as the reference, the results of generalized linear mixed-effects models showed that children with long sleep + dehydration (odds ratio [OR]: 3.87 for tubular impairment [tubules] and 3.47 for glomerular impairment [glomerulus]), long sleep + euhydration (OR: 2.43 for tubules), optimal sleep + dehydration (OR: 2.35 for tubules and 3.00 for glomerulus), short sleep + dehydration (OR: 2.07 for tubules and 2.69 for glomerulus), or short sleep + euhydration (OR: 1.29 for tubules) were more likely to present transient renal impairment, adjusting for sex, age, body mass index z-score, systolic blood pressure z-score, screen time, physical activity, and Mediterranean diet adherence. Conclusions: Dehydration and suboptimal sleep aggravate transient renal impairment in children, suggesting its role in maintaining pediatric kidney health. Full article
(This article belongs to the Special Issue 2024 Collection: Dietary, Lifestyle and Children Health)
Show Figures

Figure 1

Figure 1
<p>Median and prevalence of urinary indicators across four waves of this study by sleep duration. β<sub>2</sub>-MG: β<sub>2</sub>-microglobulin; MA: microalbumin; SG: specific gravity; IQR: interquartile range.</p>
Full article ">Figure 2
<p>Longitudinal associations between the interaction of sleep duration and hydration status with renal tubular and glomerular impairment among children aged 6–9 years in Beijing, China. Model 1: unadjusted; model 2: adjusted for sex, age, and body mass index z-score; model 3: model 2 further adjusted for systolic blood pressure z-score, screen time, physical activity level, and Mediterranean diet adherence. cOR: crude odds ratio; aOR: adjusted odds ratio; CI: confidence interval. All models included one random effect: the weekday of the urine assay.</p>
Full article ">
11 pages, 1121 KiB  
Article
Impact of Serum Phosphate on Hemoglobin Level: A Longitudinal Analysis on a Large Cohort of Dialysis Patients
by Vincenzo Calabrese, Giovanni Luigi Tripepi, Domenico Santoro, Valeria Cernaro, Vincenzo Antonio Panuccio, Sabrina Mezzatesta, Francesco Mattace-Raso, Claudia Torino and on behalf of the Sicilian Registry of Nephrology, Dialysis and Transplantation
J. Clin. Med. 2024, 13(19), 5657; https://doi.org/10.3390/jcm13195657 - 24 Sep 2024
Viewed by 509
Abstract
Background/Objectives: Phosphate is a macro-element involved in all cellular energetic processes. As about 90% of the phosphate filtered by the glomerulus is excreted by kidneys, the impairment of renal function and the consequent over-secretion of parathyroid hormone and fibroblast growth factor 23 [...] Read more.
Background/Objectives: Phosphate is a macro-element involved in all cellular energetic processes. As about 90% of the phosphate filtered by the glomerulus is excreted by kidneys, the impairment of renal function and the consequent over-secretion of parathyroid hormone and fibroblast growth factor 23 results in the increase in the serum phosphate levels. The association between phosphate and hemoglobin is controversial, as both direct and indirect relationships have been reported. The present study aims to investigate the relationship between phosphate and hemoglobin in a large prospective, longitudinal cohort including dialysis patients from the Sicilian Registry of Nephrology, Dialysis, and Transplantation. Methods: In this prospective cohort study, we included 6263 hemodialysis patients to achieve a total of 120,462 repeated measurements of serum phosphate and hemoglobin over time. The longitudinal association between phosphate and hemoglobin was analyzed by univariate and multivariate Linear Mixed Models. Results: The mean age was 66 ± 16 years and the median dialysis vintage was 5 months [IQR: 2–16]. Mean and median values of hemoglobin and phosphate were 10.7 g/dL (SD 1.3 g/dL) and 4.6 mg/dL [IQR 3.9–5.5 mg/dL], respectively. The multivariate model, adjusted for potential confounders, confirmed the positive association between serum phosphate and hemoglobin [adjβ = 0.13, 95%CI 0.03–0.23, p = 0.01)]. These results were confirmed in analyses stratified for the use of phosphate binders. Conclusions: In our large cohort of dialysis patients, we found a linear, direct relationship between phosphate and hemoglobin levels. As a reduction in phosphate is associated with a parallel reduction in hemoglobin levels, hypophosphatemia can accentuate anemia in dialysis patients. Our results generate the hypothesis that monitoring serum phosphate in clinical practice might provide a better management of anemia. Full article
(This article belongs to the Special Issue Clinical Epidemiology in Chronic Kidney Disease)
Show Figures

Figure 1

Figure 1
<p>Scatterplot of the inter-relationship between serum phosphate and hemoglobin.</p>
Full article ">Figure 2
<p>Forest plot of the association between serum phosphate and hemoglobin for patients who took phosphate binders. Both of them were positively related to hemoglobin both in univariate (<b>A</b>) and multivariate (<b>B</b>) analysis. The square represents the coefficient, and the line represents the 95% confidence interval. Upper lines show the coefficients in patients who assumed phosphorous binders, whereas lower lines show the coefficients in patients who did not assume phosphorous binders.</p>
Full article ">Figure 3
<p>Forest plot of the association between serum phosphate and hemoglobin, stratified for deciles of age. Although no linear interaction was found, a trend was found in four deciles from 62 to 76.5 years. The square represents the coefficient, and the line represents the 95% confidence interval.</p>
Full article ">
24 pages, 3003 KiB  
Review
Fenestrated Endothelial Cells across Organs: Insights into Kidney Function and Disease
by Xingrui Mou, Sophia M. Leeman, Yasmin Roye, Carmen Miller and Samira Musah
Int. J. Mol. Sci. 2024, 25(16), 9107; https://doi.org/10.3390/ijms25169107 - 22 Aug 2024
Viewed by 1133
Abstract
In the human body, the vascular system plays an indispensable role in maintaining homeostasis by supplying oxygen and nutrients to cells and organs and facilitating the removal of metabolic waste and toxins. Blood vessels—the key constituents of the vascular system—are composed of a [...] Read more.
In the human body, the vascular system plays an indispensable role in maintaining homeostasis by supplying oxygen and nutrients to cells and organs and facilitating the removal of metabolic waste and toxins. Blood vessels—the key constituents of the vascular system—are composed of a layer of endothelial cells on their luminal surface. In most organs, tightly packed endothelial cells serve as a barrier separating blood and lymph from surrounding tissues. Intriguingly, endothelial cells in some tissues and organs (e.g., choroid plexus, liver sinusoids, small intestines, and kidney glomerulus) form transcellular pores called fenestrations that facilitate molecular and ionic transport across the vasculature and mediate immune responses through leukocyte transmigration. However, the development and unique functions of endothelial cell fenestrations across organs are yet to be fully uncovered. This review article provides an overview of fenestrated endothelial cells in multiple organs. We describe their development and organ-specific roles, with expanded discussions on their contributions to glomerular health and disease. We extend these discussions to highlight the dynamic changes in endothelial cell fenestrations in diabetic nephropathy, focal segmental glomerulosclerosis, Alport syndrome, and preeclampsia, and how these unique cellular features could be targeted for therapeutic development. Finally, we discuss emerging technologies for in vitro modeling of biological systems, and their relevance for advancing the current understanding of endothelial cell fenestrations in health and disease. Full article
(This article belongs to the Special Issue Molecular Mechanisms of Endothelial Dysfunction 3.0)
Show Figures

Figure 1

Figure 1
<p>Schematic illustration of endothelial fenestration development and their distribution in various human organs. (<b>A</b>) Types of endothelial cell fenestrations, including diaphragmed fenestrations with PLVAP-positive diaphragms (top), non-diaphragmed fenestrations (middle), and non-diaphragmed “discontinuations” with intercellular gaps (bottom). Created with BioRender (<a href="https://app.biorender.com/" target="_blank">https://app.biorender.com/</a>, accessed on 22 July 2024). (<b>B</b>) Developmental pathways implicated in endothelial fenestrations and downstream of VEGF signaling. NO: Nitric oxide. Created with BioRender (<a href="https://app.biorender.com/" target="_blank">https://app.biorender.com/</a>, accessed on 22 July 2024). (<b>C</b>) Hypothesis of structural changes in endothelial cells during fenestration development. Created with BioRender. (<b>D</b>) Schematic of a human body highlighting various organs that contain fenestrated endothelial cells. Endothelial cells are labeled in pink. Created with BioRender (<a href="https://app.biorender.com/" target="_blank">https://app.biorender.com/</a>, accessed on 22 July 2024).</p>
Full article ">Figure 2
<p>Illustration of glomerular endothelial cells in kidney glomerulus under healthy condition at the tissue level, with size-selective filtration function (<b>top left</b>) and at the molecular level with non-diaphragmed endothelial fenestrations (<b>top right</b>), compared to diseased condition at the tissue level with the loss of size-selective filtration function (<b>bottom left</b>) and at the molecular level with the presence of fenestration diaphragms (PLVAP-positive) along with increased VEGF signaling (<b>bottom right</b>). Glomerular endothelial cells are labeled in pink, podocytes are labeled in brown. Created with BioRender (<a href="https://app.biorender.com/" target="_blank">https://app.biorender.com/</a>, accessed on 14 May 2024).</p>
Full article ">Figure 3
<p>Engineered environments that induce tissue-specific fenestrations in ECs derived from unspecialized hiPSCs. Illustration of tissue-specific EC fenestration induction via in situ development including Glomeruloid, with influences by endogenous secretion of VEGF and other currently unidentified growth factors and cytokines (<b>top left</b>); Glomerulus Chip, with influences by VEGF, other growth factors, and cytokines, as well as mechanical factors (e.g., shear stress) and cell-laden ECM with physiologically-relevant structure (<b>bottom left</b>). Illustration of tissue-specific EC fenestration induction with defined variables including 3D-scaffolds (<b>top right</b>), chemically-defined medium (<b>middle right</b>), or mechanical stress (<b>bottom right</b>). Created with BioRender (<a href="https://app.biorender.com/" target="_blank">https://app.biorender.com/</a>, accessed on 13 June 2024).</p>
Full article ">
17 pages, 9325 KiB  
Article
Dysfunction of Mitochondrial Dynamics Induces Endocytosis Defect and Cell Damage in Drosophila Nephrocytes
by Jun-yi Zhu, Jianli Duan, Joyce van de Leemput and Zhe Han
Cells 2024, 13(15), 1253; https://doi.org/10.3390/cells13151253 - 25 Jul 2024
Viewed by 624
Abstract
Mitochondria are crucial for cellular ATP production. They are highly dynamic organelles, whose morphology and function are controlled through mitochondrial fusion and fission. The specific roles of mitochondria in podocytes, the highly specialized cells of the kidney glomerulus, remain less understood. Given the [...] Read more.
Mitochondria are crucial for cellular ATP production. They are highly dynamic organelles, whose morphology and function are controlled through mitochondrial fusion and fission. The specific roles of mitochondria in podocytes, the highly specialized cells of the kidney glomerulus, remain less understood. Given the significant structural, functional, and molecular similarities between mammalian podocytes and Drosophila nephrocytes, we employed fly nephrocytes to explore the roles of mitochondria in cellular function. Our study revealed that alterations in the Pink1–Park (mammalian PINK1–PRKN) pathway can disrupt mitochondrial dynamics in Drosophila nephrocytes. This disruption led to either fragmented or enlarged mitochondria, both of which impaired mitochondrial function. The mitochondrial dysfunction subsequently triggered defective intracellular endocytosis, protein aggregation, and cellular damage. These findings underscore the critical roles of mitochondria in nephrocyte functionality. Full article
(This article belongs to the Special Issue Drosophila Model in Molecular Mechanisms of Kidney Dysfunction)
Show Figures

Figure 1

Figure 1
<p>Mitochondria in <span class="html-italic">Drosophila</span> nephrocytes: (<b>A</b>) UAS-mito-GFP (green) induced by nephrocyte-specific driver <span class="html-italic">Dot</span>-Gal4 to label the mitochondria in <span class="html-italic">Drosophila</span> nephrocytes (<span class="html-italic">Dot</span> &gt; mito-GFP, female, 4-day-old). Phalloidin (red) labels the actin filaments and was used to visualize the fly heart. Scale bar = 20 µm. (<b>B</b>,<b>B’</b>) <span class="html-italic">Dot</span> &gt; mito-GFP (green) was to label the mitochondria in <span class="html-italic">Drosophila</span> nephrocytes (female, 4-day-old). Mitochondrial membrane potential was indicated by tetramethylrhodamine, methyl ester (TMRM; red). White dotted lines, outline the nephrocyte and the nucleus (N); purple dotted box, outlines the area magnified in (<b>B’</b>). Scale bar: (<b>B</b>) = 5 µm; (<b>B’</b>) = 1 µm. (<b>C</b>) Mosaic of transmission electron microscopy (TEM) images showing nephrocyte ultrastructure with nucleus (N) and mitochondria (M). Scale bars: (left) = 5 µm; (right, both) = 2 µm.</p>
Full article ">Figure 1 Cont.
<p>Mitochondria in <span class="html-italic">Drosophila</span> nephrocytes: (<b>A</b>) UAS-mito-GFP (green) induced by nephrocyte-specific driver <span class="html-italic">Dot</span>-Gal4 to label the mitochondria in <span class="html-italic">Drosophila</span> nephrocytes (<span class="html-italic">Dot</span> &gt; mito-GFP, female, 4-day-old). Phalloidin (red) labels the actin filaments and was used to visualize the fly heart. Scale bar = 20 µm. (<b>B</b>,<b>B’</b>) <span class="html-italic">Dot</span> &gt; mito-GFP (green) was to label the mitochondria in <span class="html-italic">Drosophila</span> nephrocytes (female, 4-day-old). Mitochondrial membrane potential was indicated by tetramethylrhodamine, methyl ester (TMRM; red). White dotted lines, outline the nephrocyte and the nucleus (N); purple dotted box, outlines the area magnified in (<b>B’</b>). Scale bar: (<b>B</b>) = 5 µm; (<b>B’</b>) = 1 µm. (<b>C</b>) Mosaic of transmission electron microscopy (TEM) images showing nephrocyte ultrastructure with nucleus (N) and mitochondria (M). Scale bars: (left) = 5 µm; (right, both) = 2 µm.</p>
Full article ">Figure 2
<p>Defective mitochondrial dynamics resulted in altered mitochondrial morphology and reduced mitochondrial function: (<b>A</b>–<b>D’</b>) Mitochondrial morphology in <span class="html-italic">Drosophila</span> nephrocytes (female, 4-day-old) overexpressing (OE) or inhibiting (RNAi; IR) <span class="html-italic">Pink1</span>, <span class="html-italic">park</span>, or <span class="html-italic">Marf</span>. Control, <span class="html-italic">Dot</span> &gt; mito-GFP. (<b>A</b>) <span class="html-italic">Dot</span> &gt; mito-GFP (green) was to label the mitochondria. Mitochondrial membrane potential was indicated by TMRM (red). Scale bar = 1 µm. (<b>B</b>) Quantitation of mitochondrial size in <span class="html-italic">Pink1</span>/<span class="html-italic">park</span>/<span class="html-italic">Marf</span>-OE/IR nephrocytes. (<b>C</b>) Quantitation of TMRM fluorescence in <span class="html-italic">Pink1</span>/<span class="html-italic">park</span>/<span class="html-italic">Marf</span>-OE/IR nephrocytes. (<b>D</b>,<b>D’</b>) Transmission electron microscopy (TEM) of mitochondria in nephrocytes. Red dotted box outlines the area magnified in (<b>D’</b>). Scale bar: (<b>D</b>) = 5 µm; (<b>D’</b>) = 2 µm. N = 30 mitochondria from 6 flies, per group per assay. Results presented as mean ± SD; statistical significance: *, <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 3
<p>Defective mitochondrial dynamics reduced nephrocyte function: (<b>A</b>–<b>D</b>) Function in <span class="html-italic">Drosophila</span> nephrocytes (female, 4-day-old) overexpressing (OE) or inhibiting (RNAi; IR) <span class="html-italic">Pink1</span>, <span class="html-italic">park</span>, or <span class="html-italic">Marf</span>. Control, <span class="html-italic">Dot</span> &gt; <span class="html-italic">w</span><sup>1118</sup>. (<b>A</b>) 10 kDa dextran (red) uptake by nephrocytes. Scale bar = 25 µm. (<b>B</b>) Quantitation of 10 kDa dextran fluorescence in <span class="html-italic">Pink1</span>/<span class="html-italic">park</span>/<span class="html-italic">Marf</span>-OE/IR nephrocytes, relative to control. (<b>C</b>) Albumin (green) uptake by nephrocytes. Scale bar = 25 µm. (<b>D</b>) Quantitation of albumin fluorescence in <span class="html-italic">Pink1</span>/<span class="html-italic">park</span>/<span class="html-italic">Marf</span>-OE/IR nephrocytes, relative to control. N = 30 nephrocytes from 6 flies, per group per assay. Results presented as mean ± SD; statistical significance: *, <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 4
<p>Defective mitochondrial dynamics in nephrocytes lead to a disrupted slit diaphragm structure: (<b>A</b>–<b>C</b>) Slit diaphragm in <span class="html-italic">Drosophila</span> nephrocytes (female, 4-day-old) overexpressing (OE) or inhibiting (RNAi; IR) <span class="html-italic">Pink1</span>, <span class="html-italic">park</span>, or <span class="html-italic">Marf</span>. Control, <span class="html-italic">Dot</span> &gt; <span class="html-italic">w</span><sup>1118</sup>. (<b>A</b>–<b>B’</b>) Distribution of slit diaphragm protein Polychaetoid (Pyd; red) by immunofluorescence in <span class="html-italic">Drosophila</span> nephrocytes. (<b>A</b>) Imaged at medial optical sections. Scale bar = 5 µm. (<b>B</b>,<b>B’</b>) Imaged at nephrocyte surface. While the dotted box outlines the area magnified in (<b>B’</b>). Scale bar: (<b>B</b>) = 5 µm; (<b>B’</b>) = 1 µm. (<b>C</b>) Transmission electron microscopy (TEM) displaying nephrocyte ultrastructure featuring slit diaphragms and lacunar channels around the circumference. Asterisk (*) indicates lacunar channel. Scale bar = 200 nm. (<b>D</b>) Quantitation of numbers of slit diaphragms in <span class="html-italic">Pink1</span>/<span class="html-italic">park</span>/<span class="html-italic">Marf</span>-OE/IR nephrocytes, relative to control. N = 8 flies, per group. Results presented as mean ± SD; statistical significance: *, <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 5
<p>Defective mitochondrial dynamics in nephrocytes impaired endocytic membrane trafficking: (<b>A</b>–<b>F</b>) Endocytic membrane trafficking in <span class="html-italic">Drosophila</span> nephrocytes (female, 4-day-old) overexpressing (OE) or inhibiting (RNAi; IR) <span class="html-italic">Pink1</span>, <span class="html-italic">park</span> or <span class="html-italic">Marf</span>. Control, (<b>A</b>) <span class="html-italic">Dot</span> &gt; Rab5-YFP, (<b>B</b>) <span class="html-italic">Dot</span> &gt; Rab7-YFP (<b>C</b>) <span class="html-italic">Dot</span> &gt; Rab11-YFP. (<b>A</b>) Early endosome visualized by Rab5-YFP (green) in nephrocytes. Scale bar = 5 µm. Boxed areas are shown magnified to the right (scale bar: 1 μm). (<b>B</b>) Quantitation of Rab5-YFP fluorescence in <span class="html-italic">Pink1</span>/<span class="html-italic">park</span>/<span class="html-italic">Marf</span>-OE/IR nephrocytes. (<b>C</b>) Late endosomes visualized by Rab7-YFP (green) in nephrocytes. Scale bar = 5 µm. Boxed areas are shown magnified to the right (scale bar: 1 μm). (<b>D</b>) Quantitation of Rab7-YFP fluorescence in <span class="html-italic">Pink1</span>/<span class="html-italic">park</span>/<span class="html-italic">Marf</span>-OE/IR nephrocytes. (<b>E</b>) Recycling endosomes visualized by Rab11-YFP (green) in nephrocytes. Scale bar = 5 µm. Boxed areas are shown magnified to the right (scale bar: 1 μm). (<b>F</b>) Quantitation of Rab11-YFP fluorescence in <span class="html-italic">Pink1</span>/<span class="html-italic">park</span>/<span class="html-italic">Marf</span>-OE/IR nephrocytes, relative to control. N = 30 nephrocytes from 6 flies, per group per assay. Results presented as mean ± SD; statistical significance: *, <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 6
<p>Defective mitochondrial dynamics in nephrocytes induced protein aggregation and increased ROS: (<b>A</b>–<b>C</b>) Reactive oxygen species (ROS) and protein aggregation in <span class="html-italic">Drosophila</span> nephrocytes (female, 4-day-old) overexpressing (OE) or inhibiting (RNAi; IR) <span class="html-italic">Pink1</span>, <span class="html-italic">park</span>, or <span class="html-italic">Marf</span>. Control, <span class="html-italic">Dot</span> &gt; <span class="html-italic">w</span><sup>1118</sup>. (<b>A</b>) Dihydroethidium (DHE; red) labels reactive oxygen species (ROS); and, DAPI (blue) was used to visualize the nuclei in nephrocytes. Scale bar = 25 µm. (<b>B</b>) Quantitation of DHE fluorescence in <span class="html-italic">Pink1</span>/<span class="html-italic">park</span>/<span class="html-italic">Marf</span>-OE/IR nephrocytes. (<b>C</b>) Anti-ubiquitinylated proteins antibody, clone FK2 (green) was used to label protein aggregates in nephrocytes. Scale bar = 5 µm. Boxed areas are shown magnified to the right (scale bar: 1 μm). (<b>D</b>) Quantitation of FK2 fluorescence in <span class="html-italic">Pink1</span>/<span class="html-italic">park</span>/<span class="html-italic">Marf</span>-OE/IR nephrocytes, relative to control. N = 30 nephrocytes from 6 flies, per group per assay. Results are presented as mean ± SD. Statistical significance (*) is defined as <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 7
<p>Defective endocytosis reduced nephrocyte function, induced protein aggregation, and increased ROS: (<b>A</b>–<b>H</b>) Function, protein aggregation, and ROS in <span class="html-italic">Drosophila</span> nephrocytes (female, 4-day-old) inhibiting (RNAi; IR) <span class="html-italic">Rab5</span>. Control, <span class="html-italic">Dot</span> &gt; <span class="html-italic">w</span><sup>1118</sup>. (<b>A</b>) 10 kDa dextran (red) and albumin (green) uptake by nephrocytes. Scale bar = 25 µm. (<b>B</b>) Quantitation of 10 kDa dextran fluorescence in <span class="html-italic">Rab4</span>-IR nephrocytes. (<b>C</b>) Quantitation of albumin fluorescence in <span class="html-italic">Rab5</span>-IR nephrocyte. (<b>D</b>) Localization of slit diaphragm protein Polychaetoid (Pyd; red) by immunofluorescence in <span class="html-italic">Drosophila</span> nephrocytes. Imaged at nephrocyte surface. Scale bar = 1 µm. (<b>E</b>) Dihydroethidium (DHE; red) labels reactive oxygen species (ROS); and, DAPI (blue) was used to visualize the nuclei in nephrocytes. Scale bar = 25 µm. (<b>F</b>) Quantitation of DHE fluorescence in <span class="html-italic">Rab5</span>-IR nephrocytes. (<b>G</b>) Anti-ubiquitinylated proteins antibody, clone FK2 (green) was used to label protein aggregates in nephrocytes. Scale bar = 5 µm. Boxed areas are shown magnified to the right (scale bar: 1 μm). (<b>H</b>) Quantitation of FK2 fluorescence in <span class="html-italic">Rab5</span>-IR nephrocytes, relative to control. N = 30 nephrocytes from 6 flies, per group per assay. Results are presented as mean ± SD. Statistical significance (*) is defined as <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 8
<p>Model of nephrocyte damage mediated by mitochondrial dynamics Schematic representation of the regulation of mitochondrial dynamics. Mitochondrial dynamics can be disrupted by the genetic modification of the Pink1–Park pathway (<span class="html-italic">Pink1</span>/<span class="html-italic">park</span>/<span class="html-italic">Marf</span> overexpression or silencing; red upward arrow indicates increase; blue downward arrow indicates decrease). The resulting changes in mitochondrial morphology lead to mitochondrial dysfunction, which impairs the endocytic trafficking pathway, resulting in protein aggregation, and increased reactive oxygen species (ROS), culminating in further tissue damage. Marf, Mitochondrial assembly regulatory factor; Park, Parkin; Pink1, PTEN-induced putative kinase 1.</p>
Full article ">
15 pages, 1234 KiB  
Article
Identification of Pathogenic Pathways for Recurrence of Focal Segmental Glomerulosclerosis after Kidney Transplantation
by Sahra Pajenda, Daniela Gerges, Ludwig Wagner, David O’Connell, Monika Aiad, Richard Imre, Karl Mechtler, Alexander Zimprich, Alice Schmidt, Guerkan Sengoelge and Wolfgang Winnicki
Diagnostics 2024, 14(15), 1591; https://doi.org/10.3390/diagnostics14151591 - 24 Jul 2024
Viewed by 814
Abstract
Primary focal segmental glomerulosclerosis (FSGS) is a disease of the podocytes and glomerulus, leading to nephrotic syndrome and progressive loss of renal function. One of the most serious aspects is its recurrence of disease in over 30% of patients following allogeneic kidney transplantation, [...] Read more.
Primary focal segmental glomerulosclerosis (FSGS) is a disease of the podocytes and glomerulus, leading to nephrotic syndrome and progressive loss of renal function. One of the most serious aspects is its recurrence of disease in over 30% of patients following allogeneic kidney transplantation, leading to early graft loss. This research investigates the individual genetic predispositions and differences in the immune responses leading to recurrence of FSGS after transplantation. We performed exome sequencing on six patients with recurrent FSGS to identify variants in fifty-one genes and found significant variations in the alpha-2-macroglobulin (A2M). Immunoblotting was used to investigate effects of specific gene variants at the protein level. Further expression analysis identified A2M, exophilin 5 (EXPH5) and plectin (PLEC) as specific proteins linked to podocytes, endothelial cells, and the glomerulus. Subsequent protein array screening revealed the presence of non-HLA-specific antibodies, including TRIM21, after transplantation. Using Metascape for pathway and process enrichment analysis, we focused on the IL-17 signaling and chemotaxis pathways. ELISA measurements showed significantly elevated IL-17 levels in patients with recurrent FSGS (32.30 ± 9.12 pg/mL) compared to individuals with other glomerular diseases (23.16 ± 2.49 pg/mL; p < 0.01) and healthy subjects (22.28 ± 0.94 pg/mL; p < 0.01), with no significant difference in plasma CCL2/MCP-1 levels between groups. This study explores the molecular dynamics underlying recurrence of FSGS after transplantation, offering insights into potential biomarkers and therapeutic targets for the future development of individualized treatments for transplant patients. Full article
(This article belongs to the Special Issue Nephrology: Diagnosis and Management)
Show Figures

Figure 1

Figure 1
<p>Serum Electrophoresis of Patient with Heterozygous Leu18Arg Missense Mutation in A2M. The morphology of the double peak formation in the alpha2 fraction (marked as circle in <b>a</b>) changed over time and course of disease (<b>a</b>–<b>d</b>). For comparison, the serum electrophoresis of a healthy subject (<b>e</b>) and of a patient with nephrotic syndrome (<b>f</b>). Data were recorded by capillary electrophoresis; the <span class="html-italic">x</span>-axis represents capillary mobility and the <span class="html-italic">y</span>-axis abundance of protein.</p>
Full article ">Figure 2
<p>Metascape enrichment analysis of 51 altered genes linked to recurrence of FSGS in kidney allografts. The analysis identified 13 clusters, each represented by significant enrichment terms.</p>
Full article ">Figure 3
<p>Plasma concentration of IL-17 in patients with FSGS recurrence after kidney transplantation compared to patients with other glomerular disease and healthy individuals.</p>
Full article ">Figure 4
<p>Plasma concentration of CCL2/MCP-1 in patients with FSGS recurrence after kidney transplantation compared to patients with glomerular disease and healthy individuals.</p>
Full article ">
9 pages, 1440 KiB  
Communication
Lack of VEGFA/KDR Signaling in Conventional Renal Cell Carcinoma Explains the Low Efficacy of Target Therapy and Frequent Adverse Events
by Lehel Peterfi, Maria V. Yusenko, Gyula Kovacs and Tamas Beothe
Int. J. Mol. Sci. 2024, 25(13), 7359; https://doi.org/10.3390/ijms25137359 - 4 Jul 2024
Viewed by 747
Abstract
It is acknowledged that conventional renal cell carcinoma (cRCC), which makes up 85% of renal malignancies, is a highly vascular tumor. Humanized monoclonal antibodies were developed to inhibit tumor neo-angiogenesis, which is driven by VEGFA/KDR signaling. The results largely met our expectations, and [...] Read more.
It is acknowledged that conventional renal cell carcinoma (cRCC), which makes up 85% of renal malignancies, is a highly vascular tumor. Humanized monoclonal antibodies were developed to inhibit tumor neo-angiogenesis, which is driven by VEGFA/KDR signaling. The results largely met our expectations, and in several cases, adverse events occurred. Our study aimed to analyze the expression of VEGFA and its receptor KDR by immunohistochemistry in tissue multi-array containing 811 cRCC and find a correlation between VEGFA/KDR signaling and new vessel formation. None of the 811 cRCC displayed VEGFA-positive immunostaining. However, each glomerulus in normal kidney showed VEGFA-positive endothelial cells. KDR expression in endothelial meshwork was found in only 9% of cRCC, whereas 2% of the cRCC displayed positive KDR reaction in the cytoplasm of tumor cells. Our results disclose the involvement of VEGFA/KDR signaling in the neo-vascularization of cRCC and explain the frequent resistance to drugs targeting the VEGFA/KDR signaling and the high frequency of adverse events. Full article
(This article belongs to the Special Issue Molecular Mechanisms of Angiogenesis and Cancer)
Show Figures

Figure 1

Figure 1
<p>Expression of KDR and VEGF mRNA in distinct types of kidney tumors. KDR mRNA is expressed exclusively in cRCC without progression, whereas VEGF is expressed in nearly all conventional RCCs (marked in red). Abbreviations: FK, fetal kidney; WT, Wilms’ tumor; CCSK, clear cell sarcoma of the kidney; RTK, rhabdoid tumor of the kidney; MTSCC, mucinous tubular and spindle cell carcinoma; pRCC, papillary renal cell carcinoma; cRCC, conventional renal cell carcinoma; chRCC, chromophobe RCC; RO, renal oncocytoma; CDC, collecting duct carcinoma.</p>
Full article ">Figure 2
<p>Expression of KDR and VEGFA in fetal and adult kidneys. (<b>A</b>) KDR is expressed in emerging stromal fibroblasts in fetal kidney. (<b>B</b>) No VEGFA expression was detected in fetal kidney. (<b>C</b>) In adult kidney, larger blood vessels displayed positive KDR staining in myo-fibroblasts of arterial walls, whereas endothelial cells were negative (arrows). (<b>D</b>) VEGFA expressed exclusively in endothelial cells of an arteria of similar caliber (arrows). (<b>E</b>) In some cortical areas of adult kidney, single endothelial cells were positive with KDR (arrows). (<b>F</b>) VEGFA expressed exclusively in endothelial cells of glomeruli (arrows). Scale bar: 40 μm.</p>
Full article ">Figure 3
<p>Expression of CD31 and KDR in conventional RCC. (<b>A</b>) Conventional RCC showing a capillary meshwork of CD31-positive endothelial cells. (<b>B</b>) No KDR expression is seen in identical TMA-biopsy specimen. (<b>C</b>) Strong CD31 expression in endothelial cells of a tubular–papillary growing conventional RCC (arrows). (<b>D</b>) In the same tumor biopsy, only few endothelial cells display positive staining with KDR antibody (arrows). (<b>E</b>) Diffuse regular meshwork of KDR-positive capillaries in a conventional RCC. (<b>F</b>) Positive KDR staining in the cytoplasm of a cRCC of rhabdoid histology. Scale bar: 40 μm.</p>
Full article ">
23 pages, 1747 KiB  
Review
The Role of Mitochondrial Sirtuins (SIRT3, SIRT4 and SIRT5) in Renal Cell Metabolism: Implication for Kidney Diseases
by Florian Juszczak, Thierry Arnould and Anne-Emilie Declèves
Int. J. Mol. Sci. 2024, 25(13), 6936; https://doi.org/10.3390/ijms25136936 - 25 Jun 2024
Cited by 1 | Viewed by 1200
Abstract
Kidney diseases, including chronic kidney disease (CKD), diabetic nephropathy, and acute kidney injury (AKI), represent a significant global health burden. The kidneys are metabolically very active organs demanding a large amount of ATP. They are composed of highly specialized cell types in the [...] Read more.
Kidney diseases, including chronic kidney disease (CKD), diabetic nephropathy, and acute kidney injury (AKI), represent a significant global health burden. The kidneys are metabolically very active organs demanding a large amount of ATP. They are composed of highly specialized cell types in the glomerulus and subsequent tubular compartments which fine-tune metabolism to meet their numerous and diverse functions. Defective renal cell metabolism, including altered fatty acid oxidation or glycolysis, has been linked to both AKI and CKD. Mitochondria play a vital role in renal metabolism, and emerging research has identified mitochondrial sirtuins (SIRT3, SIRT4 and SIRT5) as key regulators of renal cell metabolic adaptation, especially SIRT3. Sirtuins belong to an evolutionarily conserved family of mainly NAD+-dependent deacetylases, deacylases, and ADP-ribosyl transferases. Their dependence on NAD+, used as a co-substrate, directly links their enzymatic activity to the metabolic status of the cell. In the kidney, SIRT3 has been described to play crucial roles in the regulation of mitochondrial function, and the antioxidative and antifibrotic response. SIRT3 has been found to be constantly downregulated in renal diseases. Genetic or pharmacologic upregulation of SIRT3 has also been associated with beneficial renal outcomes. Importantly, experimental pieces of evidence suggest that SIRT3 may act as an important energy sensor in renal cells by regulating the activity of key enzymes involved in metabolic adaptation. Activation of SIRT3 may thus represent an interesting strategy to ameliorate renal cell energetics. In this review, we discuss the roles of SIRT3 in lipid and glucose metabolism and in mediating a metabolic switch in a physiological and pathological context. Moreover, we highlight the emerging significance of other mitochondrial sirtuins, SIRT4 and SIRT5, in renal metabolism. Understanding the role of mitochondrial sirtuins in kidney diseases may also open new avenues for innovative and efficient therapeutic interventions and ultimately improve the management of renal injuries. Full article
(This article belongs to the Special Issue Sirtuins as Players in Cell Metabolism and Functions)
Show Figures

Figure 1

Figure 1
<p>SIRT3-mediated regulation of renal lipid metabolism. SIRT3 directly deacetylates LCAD which promotes FAO and LKB1. LKB1 is an upstream kinase that activates AMPK, leading to the phosphorylation of ACC and suppression of malonyl-CoA production, which leads to the suppression of inhibition of CPT-1 activity and increased FAO. In parallel, AMPK induces the stimulation of PGC1α transcriptional activity via CREB, enhancing SIRT3 expression. AMPK, AMP-activated protein kinase; ACC, Acetyl-CoA carboxylase; LCAD, long-chain acyl-CoA dehydrogenase; CREB, cAMP response element-binding protein; PGC1α, co-activator peroxisome proliferator-activated receptor-γ co-activator-1α; LKB1, serine/threonine liver kinase B1; CPT1, carnitine palmitoyltransferase I; FAO, fatty acid oxidation.</p>
Full article ">Figure 2
<p>SIRT3 deficiency leads to a metabolic switch from FAO to glycolysis which drives the epithelial–mesenchymal transition (EMT) in renal cells. SIRT3 plays a central role in renal fibrosis through its regulation of cellular metabolism and various signaling pathways. SIRT3 deficiency leads to a metabolic switch from FAO to glycolysis, which drives epithelial–mesenchymal transition (EMT) in renal cells. In the context of acute kidney injury (AKI) and chronic kidney disease (CKD), SIRT3 inhibition (red arrows) exacerbates fibrosis and EMT due to increased glycolysis. This process is mainly mediated by Smad3, HIF-1α, and the dimerization of PKM2. SIRT3 inhibits TGFβ-Smad3 signaling, thus preventing the Smad3-induced fibrotic response. SIRT3 suppresses HIF-1α by inhibiting mitochondrial ROS production. SIRT3 also controls the PKM2 tetramer-to-dimer interconversion, preventing EMT. Additionally, SIRT3 inhibition results in reduced pyruvate dehydrogenase complex (PDC) activity and increased lactate accumulation, further promoting a glycolytic phenotype. These changes contribute to a profibrotic environment, highlighting the critical role of SIRT3 in maintaining metabolic homeostasis and preventing renal fibrosis. TGFβ, transforming growth factor β1; Smad3, SMAD family member 3; PKM2, pyruvate kinase isozymes M2; HIF-1α, Hypoxia-inducible factor 1α; TCA, tricarboxylic cycle; FAO, fatty acid oxidation; PDC, pyruvate dehydrogenase complex.</p>
Full article ">Figure 3
<p>The pivotal roles of mitochondrial SIRT3, SIRT4, and SIRT5 in renal cell metabolism. Both AKI, CKD, and diabetic nephropathy are linked to decreased NAD<sup>+</sup> content and subsequent SIRT activity inhibition. SIRT3 is linked to the inhibition of glycolysis while it activates FAO. SIRT3 is inhibited in renal diseases. Activity of SIRT4 is also decreased in renal diseases and plays a putative role in regulating FAO in renal cells. Activity of SIRT5 expression is increased in renal diseases and plays a role in regulating peroxisomal FAO as well as glycolysis in renal cells. SIRT5 inhibition (in SIRT5<sup>−/−</sup> kidney) blocks mitochondrial FAO and leads to a compensatory shift to peroxisomal FAO (red arrows). FAO, fatty acid oxidation; NAD, nicotinamide adenine dinucleotide.</p>
Full article ">
14 pages, 3219 KiB  
Article
Molecular Basis of CO2 Sensing in Hyphantria cunea
by Jian Zhang, Shiwen Duan, Wenlong Wang, Duo Liu and Yinliang Wang
Int. J. Mol. Sci. 2024, 25(11), 5987; https://doi.org/10.3390/ijms25115987 - 30 May 2024
Viewed by 583
Abstract
Carbon dioxide (CO2) released by plants can serve as a cue for regulating insect behaviors. Hyphantria cunea is a widely distributed forestry pest that may use CO2 as a cue for foraging and oviposition. However, the molecular mechanism underlying its [...] Read more.
Carbon dioxide (CO2) released by plants can serve as a cue for regulating insect behaviors. Hyphantria cunea is a widely distributed forestry pest that may use CO2 as a cue for foraging and oviposition. However, the molecular mechanism underlying its ability to sense CO2 has not been elucidated. Our initial study showed that CO2 is significantly attractive to H. cunea adults. Subsequently, 44 H. cunea gustatory receptors (GRs) were identified using transcriptome data, and 3 candidate CO2 receptors that are specifically expressed in the labial palps were identified. In vivo electrophysiological assays revealed that the labial palp is the primary organ for CO2 perception in H. cunea, which is similar to findings in other lepidopteran species. By using the Xenopus oocyte expression system, we showed that the HcunGR1 and HcunGR3 co-expressions produced a robust response to CO2, but HcunGR2 had an inhibitory effect on CO2 perception. Finally, immunohistochemical staining revealed sexual dimorphism in the CO2-sensitive labial pit organ glomerulus (LPOG). Taken together, our results clarified the mechanism by which H. cunea sense CO2, laying the foundation for further investigations into the role of CO2 in the rapid spread of H. cunea. Full article
(This article belongs to the Special Issue Plant Response to Insects and Microbes 2.0)
Show Figures

Figure 1

Figure 1
<p>The number of female <span class="html-italic">Hyphantria cunea</span> attracted to different concentrations of CO<sub>2</sub>. Statistical differences were evaluated via the Chi−square test. ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, ns: no significance.</p>
Full article ">Figure 2
<p>The maximum likelihood tree of candidate gustatory receptors (<span class="html-italic">GRs</span>). Bootstrap replications up to 1000.</p>
Full article ">Figure 3
<p>Expression profiles of <span class="html-italic">HcunGRs</span> in different body parts of adult <span class="html-italic">H. cunea</span>. (<b>a</b>) Characteristic expression patterns of 44 <span class="html-italic">HcunGRs</span> in different body parts based on FPKM (normalization by row). FPKM: Fragments per kilobase of transcript per million fragments mapped. (<b>b</b>) Expression patterns of three candidate CO<sub>2</sub> <span class="html-italic">GRs</span> in different body parts of adult <span class="html-italic">H. cunea</span>. A different lowercase indicates a significant difference based on one-way ANOVA followed by Tukey’s multiple comparison test (<span class="html-italic">p</span> &lt; 0.05). The data are presented as the means ± standard errors of the means (SEMs), <span class="html-italic">N</span> = 3.</p>
Full article ">Figure 4
<p>Electrolabialpalpography (ELPG) and electroantennogram (EAG) nonlinear regression curve for CO<sub>2</sub>. (<b>a</b>) Response of the labial palps of the <span class="html-italic">H. cunea</span> to CO<sub>2</sub>; (<b>b</b>) response of the antennae of the <span class="html-italic">H. cunea</span> to CO<sub>2</sub>. Statistical differences were evaluated by the Wilcoxon signed-rank test. * <span class="html-italic">p</span> &lt; 0.05, ns: no significance. The data are presented as the means ± SEMs, <span class="html-italic">N</span> = 3 biological replicates (<a href="#app1-ijms-25-05987" class="html-app">Tables S4 and S5</a>).</p>
Full article ">Figure 5
<p>Two-electrode voltage clamp recording (TEVC) responses of <span class="html-italic">HcunGR1, HcunGR2,</span> and <span class="html-italic">HcunGR3</span> alone and in combination with different concentrations of NaCl and NaHCO<sub>3</sub>. (<b>a</b>) Response of <span class="html-italic">HcunGR1+HcunGR3</span>; (<b>b</b>) response of <span class="html-italic">HcunGR1+HcunGR2+HcunGR3</span>; (<b>c</b>) response of <span class="html-italic">HcunGR1</span>; (<b>d</b>) response of <span class="html-italic">HcunGR2</span>; (<b>e</b>) response of <span class="html-italic">HcunGR3</span>; (<b>f</b>) response of <span class="html-italic">HcunGR1+HcunGR2</span>; (<b>g</b>) response of <span class="html-italic">HcunGR2+HcunGR3</span>; green traces represent the response in NaCl solution; red traces represent the response in NaHCO<sub>3</sub> solution; the number involved in (<b>a</b>–<b>g</b>) indicates concentration NaCl and NaHCO<sub>3</sub>; (<b>h</b>) nonlinear regression curve of <span class="html-italic">HcunGR1, HcunGR2,</span> and <span class="html-italic">HcunGR3</span> alone and in combination after excluding the influence of Na<sup>+</sup>. <span class="html-italic">N</span> = 3 biological replicates (<a href="#app1-ijms-25-05987" class="html-app">Table S6</a>); The data are presented as the mean ± SEM.</p>
Full article ">Figure 6
<p>Anterograde dye-filling of labial palps and two-dimensional reconstructions of the antennal lobe (AL) in the DP1 region of <span class="html-italic">H. cunea</span>. (<b>a</b>–<b>c</b>) The central projections of female labial pit organ sensory neurons passing through the gnathal ganglion (GNG) to DP1; (<b>d</b>–<b>f</b>) the central projections of male labial pit organ sensory neurons passing through the gnathal ganglion (GNG) to DP1; (<b>g</b>,<b>h</b>) confocal images of the male <span class="html-italic">H. cunea</span> AL glomeruli seen from the ventral view. The sections are from anterior to posterior at a depth of 156 μm. Scale bar = 50 μm; (<b>i</b>,<b>j</b>) confocal images of female <span class="html-italic">H. cunea</span> AL glomeruli taken from the ventral view. The sections are from anterior to posterior at a depth of 142 μm. Scale bar = 50 μm; (<b>k</b>) the volume compares male and female DP1s. The data are presented as the mean ± SEM, <span class="html-italic">N</span> ≥ 4. Statistical differences were evaluated by unpaired <span class="html-italic">t</span> tests. * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">
15 pages, 1039 KiB  
Review
Research Progress of Drug Delivery Systems Targeting the Kidneys
by Li-Feng Huang, Qiao-Ru Ye, Xiao-Cui Chen, Xiao-Rong Huang, Qiao-Fei Zhang, Chun-Yu Wu, Hua-Feng Liu and Chen Yang
Pharmaceuticals 2024, 17(5), 625; https://doi.org/10.3390/ph17050625 - 13 May 2024
Cited by 1 | Viewed by 1493
Abstract
Chronic kidney disease (CKD) affects more than 10% of the global population, and its incidence is increasing, partially due to an increase in the prevalence of disease risk factors. Acute kidney injury (AKI) is an independent risk factor for CKD and end-stage renal [...] Read more.
Chronic kidney disease (CKD) affects more than 10% of the global population, and its incidence is increasing, partially due to an increase in the prevalence of disease risk factors. Acute kidney injury (AKI) is an independent risk factor for CKD and end-stage renal disease (ESRD). The pathogenic mechanisms of CKD provide several potential targets for its treatment. However, due to off-target effects, conventional drugs for CKD typically require high doses to achieve adequate therapeutic effects, leading to long-term organ toxicity. Therefore, ideal treatments that completely cure the different types of kidney disease are rarely available. Several approaches for the drug targeting of the kidneys have been explored in drug delivery system research. Nanotechnology-based drug delivery systems have multiple merits, including good biocompatibility, suitable degradability, the ability to target lesion sites, and fewer non-specific systemic effects. In this review, the development, potential, and limitations of low-molecular-weight protein–lysozymes, polymer nanomaterials, and lipid-based nanocarriers as drug delivery platforms for treating AKI and CKD are summarized. Full article
(This article belongs to the Section Pharmaceutical Technology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Targeted drug delivery systems: (<b>A</b>) deliverable medications; (<b>B</b>) different types of carriers; (<b>C</b>) different target ligands; (<b>D</b>) different strategies of drug delivery systems; and (<b>E</b>) targeting kidney. Abbreviations: DSPE-PEG-GLU: phospholipid–polyethylene glycol–glucose; LMWPs: low-molecular-weight proteins; PLGA: poly lactic-co-glycolic acid; PAMAMs: polyamidoamine dendrimers; and TRX-20: 3,5-dipentadecyloxybenzamidine hydrochloride. The figure was drawn by Figdraw (<a href="http://www.figdraw.com" target="_blank">www.figdraw.com</a>), accessed on 25 April 2024.</p>
Full article ">Figure 2
<p>Drugs absorbed via different receptors: (<b>A</b>) the structure of the glomeruli in the kidneys is shown; (<b>B</b>) nanocomplex absorbed via VCAM-1, ICAM-1, megalin/cubilin, and the KIM-1 receptor in renal tubular epithelial cells; (<b>C</b>) nanocomplex absorbed via VCAM-1 receptors in endothelial cells; (<b>D</b>) nanocomplex absorbed via VCAM-1 receptors in podocytes.; and (<b>E</b>) nanocomplexes absorbed by GLUT1 and chondroitin sulfate proteoglycan receptors in mesangial cells. Abbreviations: ICAM-1: intercellular adhesion molecule-1; VCAM-1: vascular cell adhesion molecule-1. The figure was drawn by Figdraw (<a href="http://www.figdraw.com" target="_blank">www.figdraw.com</a>), accessed on 25 April 2024.</p>
Full article ">
18 pages, 9633 KiB  
Article
Characterizing Glomerular Barrier Dysfunction with Patient-Derived Serum in Glomerulus-on-a-Chip Models: Unveiling New Insights into Glomerulonephritis
by Shin Young Kim, Yun Yeong Choi, Eun Jeong Kwon, Seungwan Seo, Wan Young Kim, Sung Hyuk Park, Seokwoo Park, Ho Jun Chin, Ki Young Na and Sejoong Kim
Int. J. Mol. Sci. 2024, 25(10), 5121; https://doi.org/10.3390/ijms25105121 - 8 May 2024
Viewed by 1147
Abstract
Glomerulonephritis (GN) is characterized by podocyte injury or glomerular filtration dysfunction, which results in proteinuria and eventual loss of kidney function. Progress in studying the mechanism of GN, and developing an effective therapy, has been limited by the absence of suitable in vitro [...] Read more.
Glomerulonephritis (GN) is characterized by podocyte injury or glomerular filtration dysfunction, which results in proteinuria and eventual loss of kidney function. Progress in studying the mechanism of GN, and developing an effective therapy, has been limited by the absence of suitable in vitro models that can closely recapitulate human physiological responses. We developed a microfluidic glomerulus-on-a-chip device that can recapitulate the physiological environment to construct a functional filtration barrier, with which we investigated biological changes in podocytes and dynamic alterations in the permeability of the glomerular filtration barrier (GFB) on a chip. We also evaluated the potential of GN-mimicking devices as a model for predicting responses to human GN. Glomerular endothelial cells and podocytes successfully formed intact monolayers on opposite sides of the membrane in our chip device. Permselectivity analysis confirmed that the chip was constituted by a functional GFB that could accurately perform differential clearance of albumin and dextran. Reduction in cell viability resulting from damage was observed in all serum-induced GN models. The expression of podocyte-specific marker WT1 was also decreased. Albumin permeability was increased in most models of serum-induced IgA nephropathy (IgAN) and membranous nephropathy (MN). However, sera from patients with minimal change disease (MCD) or lupus nephritis (LN) did not induce a loss of permeability. This glomerulus-on-a-chip system may provide a platform of glomerular cell culture for in vitro GFB in formation of a functional three-dimensional glomerular structure. Establishing a disease model of GN on a chip could accelerate our understanding of pathophysiological mechanisms of glomerulopathy. Full article
(This article belongs to the Special Issue Molecular Study of Renal Diseases)
Show Figures

Figure 1

Figure 1
<p>Representative histologic features via hematoxylin and eosin (H&amp;E) staining (<b>A</b>–<b>D</b>) and electronic microscopy (<b>E</b>–<b>G</b>) of renal biopsy tissue derived from IgAN3 (<b>A</b>,<b>E</b>), MN4, (<b>B</b>,<b>F</b>), MCD5 (<b>C</b>,<b>G</b>), and LN5 (<b>D</b>). Scale bars: (<b>A</b>–<b>D</b>) 20 μm; (<b>E</b>,<b>G</b>) 5 μm; (<b>F</b>) 2 μm. Magnifications: (<b>A</b>–<b>D</b>) ×400; (<b>E</b>) ×6000; (<b>F</b>) ×8000; (<b>G</b>) ×4000.</p>
Full article ">Figure 2
<p>Morphology of human CIHP-1 at each stage of differentiation. Representative images demonstrating the morphology of CIHP-1 before differentiation at permissive temperature 33 °C (<b>A</b>) and after differentiation at non-permissive temperature 37 °C (<b>B</b>–<b>D</b>). Phase-contrast microscope images showing undifferentiated cobblestone-like cells without processes (<b>A</b>) and differentiated cell indicating short rounded projections (asterisk), long spindle-like projections (white arrow), and cell–cell contacts (black arrow and magnified insert) with interdigitations between cells (<b>B</b>–<b>D</b>). Scale bars: 50 μm. Magnifications: ×200.</p>
Full article ">Figure 3
<p>Formation and quantification of filopodia length and density of in differentiated CIHP-1 cells during the differentiation process. (<b>A</b>) Schematic diagram showing the formation of actin-rich lamellipodia and filopodia in an attempt to invade through the surrounding extracellular matrix (ECM) and migrate towards a specific direction. (<b>B</b>–<b>D</b>) Representative images of progressively extended lamellipodia and filopodia (yellow arrow) protrusions during podocyte differentiation and development. (<b>E</b>) Quantification of the length and density of filopodia (×200). Scale bars: (<b>B</b>–<b>D</b> upper) 50 μm; (<b>B</b>–<b>D</b> lower) 20 μm. Magnifications: (<b>B</b>–<b>D</b> upper) ×200; (<b>B</b>–<b>D</b> lower) ×400. Data are presented as the median (interquartile range) or mean ± standard deviation. * <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>Immunofluorescence and Western blot analysis of podocyte-specific markers in undifferentiated and differentiated CIHP-1 cells. Representative immunofluorescence staining for (<b>A</b>) podocin (green), (<b>B</b>) nephrin (green), (<b>C</b>) WT1 (purple), F-actin (red or green), and nuclei (DAPI, blue) in undifferentiated and differentiated CIHP-1 cells. Scale bars: 50 μm. Magnifications: ×200. (<b>D</b>) Western blotting of synaptopodin (99 kDa, left panel) and WT1 (55 kDa, right panel) using total protein extracts from undifferentiated and differentiated CIHP-1 cells. GAPDH (37 kDa) was used as a loading control.</p>
Full article ">Figure 5
<p>Modeling the glomerular structural and functional filtration barrier in a microfluid glomerulus-on-a-chip. (<b>A</b>) Three-dimensional visualization of Z-stack images showing the formation of glomerular endothelial cell (RFP-HGMVEC) and podocyte (CIHP-1) monolayers in the chip. (<b>B</b>) TEER measurements showing the barrier integrity of co-culture devices under baseline conditions. (<b>C</b>) Representative fluorescence images for glomerular endothelial cells (RFP-HGMVECs, red) and podocytes (CIHP-1, nephrin-FITC, green) grown on inner and outer membranes of the insert in the chip, respectively, after co-culture for 8 days. (<b>D</b>) Quantification of filtration barrier permeability to albumin-FITC (66 kDa, green) and dextran-rhodamine B (10 kDa, red) at 2 and 6 h. Data are presented as mean ± standard deviation. *** <span class="html-italic">p</span> &lt; 0.001. F-actin (purple), nuclei (DAPI, blue). Scale bars: 50 μm. Magnifications: ×200.</p>
Full article ">Figure 6
<p>Viability of CIHP-1 cells on a chip after exposure to serum for 24 h. Quantitative assessment of cell viability of patients with IgAN (<b>A</b>), MN (<b>B</b>), MCD (<b>C</b>), or LN (<b>D</b>) using the CCK-8 assay. (<b>E</b>) Representative images of CIHP-1 cells stained with calcein-AM (green, live cell). (<b>F</b>) Quantification of fluorescence intensity of calcein-AM stained CIHP-1 cells. Data are presented as mean ± standard deviation. Scale bars: 200 μm. Magnifications: ×50. * <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 7
<p>WT-1 expression in CIHP-1 cells subjected to glomerulonephritis-derived serum in a glomerular chip device. Confocal microscopic images of WT-1 staining in CIHP-1 cells after exposure to serum from patient with IgAN3, MN1, MCD5, or LN5 for 24 h. Scale bars: 50 μm. Magnifications: ×200.</p>
Full article ">Figure 8
<p>Change in the glomerular barrier permeability after exposure to GN serum in the glomerular chip system. Barrier permeability to fluorescence-labeled albumin under exposure to sera from patients with IgAN (<b>A</b>), MN (<b>B</b>), MCD (<b>C</b>), or LN (<b>D</b>). Data are presented as mean ± standard deviation. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 9
<p>Schematic drawing of a three-dimensional microfluidic glomerulus-on-a-chip device to recapitulate the structure and function of the kidney glomerular filtration barrier. (<b>A</b>) Illustration (left panel) and an electron micrograph (right panel) of the glomerular filtration barrier (GFB). The GFB comprises podocytes with their foot processes, glomerular basement membrane, and fenestrated endothelial cells. Foot processes are interconnected by slit diaphragms (arrowheads). (<b>B</b>) Architecture of the glomerular chip. The chip consists of three cell culture chambers, four microfluidic channels, and two media reservoirs, in which podocytes (CIHP-1) and glomerular endothelial cells (RFP-HGMVECs) are co-cultured on the collagen type I-coated extracellular matrix of the insert to form the GFB. (<b>C</b>) Real image of the microfluidic chip in a cell culture dish (100 mm). (<b>D</b>) Workflow for evaluating the integrity and reproducibility of GFB and for developing a glomerulonephritis (GN) model into a microfluidic chip. CIHP-1 cells were first seeded on the bottom surface of a cell culture insert. Then, RFP-HGMVECs were seeded on the inside surface of the insert. Patient serum was added to both the chamber and the inside of the insert on day 8 to induce the GN model.</p>
Full article ">
21 pages, 9712 KiB  
Article
Renal Pathological Image Classification Based on Contrastive and Transfer Learning
by Xinkai Liu, Xin Zhu, Xingjian Tian, Tsuyoshi Iwasaki, Atsuya Sato and Junichiro James Kazama
Electronics 2024, 13(7), 1403; https://doi.org/10.3390/electronics13071403 - 8 Apr 2024
Cited by 1 | Viewed by 958
Abstract
Following recent advancements in medical laboratory technology, the analysis of high-resolution renal pathological images has become increasingly important in the diagnosis and prognosis prediction of chronic nephritis. In particular, deep learning has been widely applied to computer-aided diagnosis, with an increasing number of [...] Read more.
Following recent advancements in medical laboratory technology, the analysis of high-resolution renal pathological images has become increasingly important in the diagnosis and prognosis prediction of chronic nephritis. In particular, deep learning has been widely applied to computer-aided diagnosis, with an increasing number of models being used for the analysis of renal pathological images. The diversity of renal pathological images and the imbalance between data acquisition and annotation have placed a significant burden on pathologists trying to perform reliable and timely analysis. Transfer learning based on contrastive pretraining is emerging as a viable solution to this dilemma. By incorporating unlabeled positive pretraining images and a small number of labeled target images, a transfer learning model is proposed for high-accuracy renal pathological image classification tasks. The pretraining dataset used in this study includes 5000 mouse kidney pathological images from the Open TG-GATEs pathological image dataset (produced by the Toxicogenomics Informatics Project of the National Institutes of Biomedical Innovation, Health, and Nutrition in Japan). The transfer training dataset comprises 313 human immunoglobulin A (IgA) chronic nephritis images collected at Fukushima Medical University Hospital. The self-supervised contrastive learning algorithm “Bootstrap Your Own Latent” was adopted for pretraining a residual-network (ResNet)-50 backbone network to extract glomerulus feature expressions from the mouse kidney pathological images. The self-supervised pretrained weights were then used for transfer training on the labeled images of human IgA chronic nephritis pathology, culminating in a binary classification model for supervised learning. In four cross-validation experiments, the proposed model achieved an average classification accuracy of 92.2%, surpassing the 86.8% accuracy of the original RenNet-50 model. In conclusion, this approach successfully applied transfer learning through mouse renal pathological images to achieve high classification performance with human IgA renal pathological images. Full article
Show Figures

Figure 1

Figure 1
<p>An original WSI of a mouse kidney with an actual size of 75,695 × 22,500.</p>
Full article ">Figure 2
<p>An original WSI of a patient with IgA nephritis with an actual size of 35,856 × 23,388.</p>
Full article ">Figure 3
<p>WSI patches containing human (<b>left</b>) and mouse (<b>right</b>) glomerulus.</p>
Full article ">Figure 4
<p>The structure of the identity block.</p>
Full article ">Figure 5
<p>The structure of the convolutional block.</p>
Full article ">Figure 6
<p>The ResNet backbone with BYOL algorithm.</p>
Full article ">Figure 7
<p>Learning rate reduction strategy for contrastive learning.</p>
Full article ">Figure 8
<p>Learning rate reduction strategy for transfer learning.</p>
Full article ">Figure 9
<p>Data and experimental flow in this study. The blue stream represents semi-supervised learning training. The red and green streams represent transfer learning training with pretraining by mouse kidney images from different animal drug experiments. The orange stream represents supervised learning training used for comparison and evaluation.</p>
Full article ">Figure 10
<p>The training loss curves for contrastive pretraining. The red curve represents the training process using mouse glomerulus images, and the black curve represents the training process using human IgA nephritis glomerulus images.</p>
Full article ">Figure 11
<p>Training loss curves for semi-supervised learning (red), supervised learning (black), and transfer learning with mouse dataset A (blue).</p>
Full article ">Figure 12
<p>Accurate predictions of a positive image containing glomeruli. (<b>a</b>) is the original image patch. (<b>b</b>) is the Grad-CAM of the semi-supervised learning model. The model prediction is at the top, which comprises a binary group representing the positive and negative prediction scores, as is the prediction label. The first and second terms of the binary group represent the probability of a positive prediction (containing glomeruli) and a negative prediction (without glomeruli), respectively. (<b>c</b>) is the Grad-CAM of the transfer learning model. The model prediction scores and labels are also listed at the top.</p>
Full article ">Figure 13
<p>A correct prediction of a positive image containing glomeruli (true positive). (<b>a</b>) is the original image patch and labeled as a positive image. (<b>b</b>) has a positive prediction score of 1.00. (<b>c</b>) is the Grad-CAM of the prediction.</p>
Full article ">Figure 14
<p>An incorrect prediction of a positive image containing glomeruli. (<b>a</b>) is the original image patch and labeled as a positive image. (<b>b</b>) has a positive prediction score of 0.26. (<b>c</b>) is the Grad-CAM of the positive prediction.</p>
Full article ">Figure 15
<p>A correct prediction of a negative image without glomeruli. (<b>a</b>) is the original image patch and labeled as a negative image. (<b>b</b>) has a negative prediction score of 0.99. (<b>c</b>) is the Grad-CAM of the prediction.</p>
Full article ">Figure 16
<p>An incorrect prediction of a negative image containing glomeruli. (<b>a</b>) is the original image patch and labeled as a positive image. (<b>b</b>) has a negative prediction score of 0.05. (<b>c</b>) is the Grad-CAM of the prediction.</p>
Full article ">Figure 17
<p>Confusion matrices of two transfer learning models (<b>a</b>,<b>b</b>), the supervised learning model (<b>c</b>), and the semi-supervised learning model (<b>d</b>), respectively.</p>
Full article ">Figure 18
<p>The ROC and AUROC of the 3 different models. The orange curve represents the transfer learning model of mouse dataset A with an AUROC of 0.973. The black curve represents the supervised learning model with an AUROC of 0.958. The red curve represents the semi-supervised learning model with an AUROC of 0.925. The dotted line represents a random classifier.</p>
Full article ">
18 pages, 1178 KiB  
Review
Type I IFN in Glomerular Disease: Scarring beyond the STING
by Alexis Paulina Jimenez-Uribe, Steve Mangos and Eunsil Hahm
Int. J. Mol. Sci. 2024, 25(5), 2497; https://doi.org/10.3390/ijms25052497 - 21 Feb 2024
Viewed by 1765
Abstract
The field of nephrology has recently directed a considerable amount of attention towards the stimulator of interferon genes (STING) molecule since it appears to be a potent driver of chronic kidney disease (CKD). STING and its activator, the cyclic GMP-AMP synthase (cGAS), along [...] Read more.
The field of nephrology has recently directed a considerable amount of attention towards the stimulator of interferon genes (STING) molecule since it appears to be a potent driver of chronic kidney disease (CKD). STING and its activator, the cyclic GMP-AMP synthase (cGAS), along with intracellular RIG-like receptors (RLRs) and toll-like receptors (TLRs), are potent inducers of type I interferon (IFN-I) expression. These cytokines have been long recognized as part of the mechanism used by the innate immune system to battle viral infections; however, their involvement in sterile inflammation remains unclear. Mounting evidence pointing to the involvement of the IFN-I pathway in sterile kidney inflammation provides potential insights into the complex interplay between the innate immune system and damage to the most sensitive segment of the nephron, the glomerulus. The STING pathway is often cited as one cause of renal disease not attributed to viral infections. Instead, this pathway can recognize and signal in response to host-derived nucleic acids, which are also recognized by RLRs and TLRs. It is still unclear, however, whether the development of renal diseases depends on subsequent IFN-I induction or other processes involved. This review aims to explore the main endogenous inducers of IFN-I in glomerular cells, to discuss what effects autocrine and paracrine signaling have on IFN-I induction, and to identify the pathways that are implicated in the development of glomerular damage. Full article
(This article belongs to the Special Issue Molecular Pathology, Diagnostics and Therapeutics of Nephropathy 3.0)
Show Figures

Figure 1

Figure 1
<p>Origins and intracellular sensors of self-nucleic acids promoting the IFN-I signaling pathway. During cellular stress, self-derived nucleic acids (DNA or RNA) are released from damaged mitochondria and the nucleus, consequently accumulating in the cytosol. In addition, extracellular nucleic acids liberated from neighboring dying or damaged cells are internalized by endocytosis. These intracellular (cytosolic or endosomal) nucleic acids are recognized by diverse intracellular nucleic acid sensors, triggering the activation of the signaling pathways that produce IFN-I and proinflammatory cytokines. Specifically, sensors for endosomal nucleic acids include TLR3 (for double-stranded DNA), TLR7 (for single-stranded RNA), and TLR9 (for single-stranded DNA). Cytosolic double-stranded RNA is detected by RIG-I or MDA5, while cytosolic double-stranded DNA is recognized by members of the cGAS-STING pathway. Activation of these intracellular nucleic acid sensors stimulates TBK1 activation, prompting the translocation of IRFs and NF-kB into the nucleus. There, they orchestrate the expression of IFN-I and proinflammatory cytokine genes. Subsequently, binding of IFN-I to IFNAR1/IFNAR2 triggers the activation of the JAK-STAT pathway, culminating in the induction of ISGs. This cascade of events illustrates the intricate molecular mechanisms involved in the recognition of self-nucleic acids, the subsequent activation of signaling pathways leading to the expression and secretion of IFN-I and proinflammatory cytokines, and the subsequent induction of ISGs via the JAK-STAT pathway. This figure was created with BioRender.com. IFN-I, type I interferons; TLR, toll-like receptor; RIG-I, retinoic acid-inducible gene I; MDA5, melanoma differentiation-associated protein 5; cGAS, cyclic GMP-AMP (cGAMP) synthase; STING, stimulator of interferon genes; TBK1, TANK-binding kinase 1; IFNAR, type I interferon receptor; ISGs, interferon-stimulated genes.</p>
Full article ">Figure 2
<p>The potential deleterious effects of IFN-I on different glomerular cell types. IFN-I can exhibit a direct influence on diverse cell types within the glomerular compartment, altering their cellular functions. In addition, increased IFN-I levels induce the expression of interferon-stimulated genes (ISGs). This could intensify inflammatory processes by recruiting inflammatory immune cells to the kidney and increasing the production of autoantibodies. This figure was created with BioRender.com. APOL-1, apolipoprotein L1; VE-cadherin, vascular endothelial-cadherin; NO, nitric oxide. Red arrows indicate increased responses.</p>
Full article ">
18 pages, 993 KiB  
Review
An Update on Current Therapeutic Options in IgA Nephropathy
by Regina Shaoying Lim, See Cheng Yeo, Jonathan Barratt and Dana V. Rizk
J. Clin. Med. 2024, 13(4), 947; https://doi.org/10.3390/jcm13040947 - 7 Feb 2024
Cited by 4 | Viewed by 9142
Abstract
Immunoglobulin A nephropathy (IgAN) remains the leading cause of primary glomerular disease worldwide. Outcomes are poor with high rates of progressive chronic kidney disease and kidney failure, which contributes to global healthcare costs. Although this disease entity has been described, there were no [...] Read more.
Immunoglobulin A nephropathy (IgAN) remains the leading cause of primary glomerular disease worldwide. Outcomes are poor with high rates of progressive chronic kidney disease and kidney failure, which contributes to global healthcare costs. Although this disease entity has been described, there were no disease-specific treatments until recently, with the current standard of care focusing on optimal supportive measures including lifestyle modifications and optimization of the renin-angiotensin-aldosterone blockade. However, with significant advances in the understanding of the pathogenesis of IgAN in the past decade, and the acceptance of surrogate outcomes for accelerated drug approval, there have been many new investigational agents tested to target this disease. As these agents become available, we envision a multi-pronged treatment strategy that simultaneously targets the consequences of ongoing nephron loss, stopping any glomerular inflammation, inhibiting pro-fibrotic signals in the glomerulus and tubulo-interstitium, and inhibiting the production of pathogenic IgA molecules. This review is an update on a previous review published in 2021, and we aim to summarize the developments and updates in therapeutic strategies in IgAN and highlight the promising discoveries that are likely to add to our armamentarium. Full article
Show Figures

Figure 1

Figure 1
<p>Proposed four-hit hypothesis of IgAN pathogenesis and current therapeutic options. (1) B cell priming and activation in the mucosa-associated lymphoid tissue including Peyer’s patches concentrated in the terminal ileum, resulting in the production of Galactose-deficient IgA1 (Gd-IgA1). (2) Formation of autoantibodies against IgA1 (anti-Gd-IgA1 antibodies). (3) Formation of circulating immune complexes as IgG anti-IgA1 antibodies bind to the hinge region of Gd-IgA1. (4) Deposition of circulating immune complexes in the mesangium through mesangial trapping, which triggers downstream complement activation, tissue injury, and damage. BAFF: B Cell Activating Factor; APRIL: A Proliferation-inducing Ligand; SGLT2 inhibitors: Sodium-glucose cotransporter-2 inhibitors.</p>
Full article ">Figure 2
<p>Pillars of IgAN Treatment. The management of patients with IgAN requires a multi-pronged approach to target the predominant pathogenic pathways specific to each patient at each disease timepoint, on top of a cornerstone of optimal supportive care. (1) Mitigation of the consequences of ongoing nephron loss through optimal RAAS blockade, SGLT2 inhibitors, endothelin receptor antagonists, and mineralocorticoid antagonists. (2) Halting glomerular inflammation through the use of systemic glucocorticoid therapy has been long studied for use in IgAN but offers short-term efficacy and comes with significant side effects. With a better understanding of IgAN being an immune mediated disease with activation of the alternative and lectin pathways mediating inflammation, complement pathway inhibitors may be an alternative to limit glomerular inflammation and injury in IgAN. (3) B cells are central to the pathogenesis of IgAN through the production of pathogenic Gd-IgA1. The use of TRF-budesonide can inhibit mucosal IgA production within Peyer’s patches with reduced systemic side effects of glucocorticoid therapy. B cell modulating therapies that inhibit BAFF and APRIL to reduce B cell proliferation and survival, as well as B cell depleting therapies such as borteozomib or felzartamab, may be useful in targeting B cell dysregulation. (4) There are no approved therapies for stopping pro-fibrotic signals in the kidney on the horizon, but we are hopeful for new therapeutic strategies to develop in the future. Gd-IgA1: Galactose-deficient IgA1; SGLT2 inhibitors: RAAS: renin-angiotensin-aldosterone system; Sodium-glucose cotransporter-2 inhibitors; TRF-budesonide: Targeted Release Formulation of Budesonide; BAFF: B Cell Activating Factor; APRIL: A Proliferation-inducing Ligand; MMF: Mycophenolate Mofetil; BP: blood pressure.</p>
Full article ">
8 pages, 723 KiB  
Review
Middle Molecular Uremic Toxin and Blood Purification Therapy
by Hideki Kawanishi
J. Clin. Med. 2024, 13(3), 647; https://doi.org/10.3390/jcm13030647 - 23 Jan 2024
Viewed by 2022
Abstract
The purpose of blood purification therapy is to remove uremic toxins, and middle molecules (MMs) are a specific target. An MM is defined as a solute that passes through the glomerulus with a molecular weight in the range of 0.5–58 kDa, and new [...] Read more.
The purpose of blood purification therapy is to remove uremic toxins, and middle molecules (MMs) are a specific target. An MM is defined as a solute that passes through the glomerulus with a molecular weight in the range of 0.5–58 kDa, and new classifications of “small-middle 0.5–15 kDa,” “medium-middle 15–25 kDa,” and “large-middle 25–58 kDa” were proposed. In Japan, the removal of α1-microglobulin (αMG) in the large-middle range has been the focus, but a new theory of removal has been developed, emphasizing the antioxidant effect of αMG as a physiological function. Clinical proof of this mechanism will lead to further development of blood purification therapies. Full article
(This article belongs to the Special Issue Application of Hemodialysis in the Treatment of Kidney Diseases)
Show Figures

Figure 1

Figure 1
<p>Selection of blood purification modality on middle molecules. Definition of middle molecules (MM)s: MWs 0.5 to &lt;58 KD. CV: convection volume, βMG: β2 microglobulin, FLC: free light chain, ALB: albumin.</p>
Full article ">
16 pages, 1251 KiB  
Review
Kidney Fibrosis and Oxidative Stress: From Molecular Pathways to New Pharmacological Opportunities
by Francesco Patera, Leonardo Gatticchi, Barbara Cellini, Davide Chiasserini and Gianpaolo Reboldi
Biomolecules 2024, 14(1), 137; https://doi.org/10.3390/biom14010137 - 22 Jan 2024
Cited by 8 | Viewed by 2559
Abstract
Kidney fibrosis, diffused into the interstitium, vessels, and glomerulus, is the main pathologic feature associated with loss of renal function and chronic kidney disease (CKD). Fibrosis may be triggered in kidney diseases by different genetic and molecular insults. However, several studies have shown [...] Read more.
Kidney fibrosis, diffused into the interstitium, vessels, and glomerulus, is the main pathologic feature associated with loss of renal function and chronic kidney disease (CKD). Fibrosis may be triggered in kidney diseases by different genetic and molecular insults. However, several studies have shown that fibrosis can be linked to oxidative stress and mitochondrial dysfunction in CKD. In this review, we will focus on three pathways that link oxidative stress and kidney fibrosis, namely: (i) hyperglycemia and mitochondrial energy imbalance, (ii) the mineralocorticoid signaling pathway, and (iii) the hypoxia-inducible factor (HIF) pathway. We selected these pathways because they are targeted by available medications capable of reducing kidney fibrosis, such as sodium-glucose cotransporter-2 (SGLT2) inhibitors, non-steroidal mineralocorticoid receptor antagonists (MRAs), and HIF-1alpha-prolyl hydroxylase inhibitors. These drugs have shown a reduction in oxidative stress in the kidney and a reduced collagen deposition across different CKD subtypes. However, there is still a long and winding road to a clear understanding of the anti-fibrotic effects of these compounds in humans, due to the inherent practical and ethical difficulties in obtaining sequential kidney biopsies and the lack of specific fibrosis biomarkers measurable in easily accessible matrices like urine. In this narrative review, we will describe these three pathways, their interconnections, and their link to and activity in oxidative stress and kidney fibrosis. Full article
(This article belongs to the Special Issue Redox Imbalance and Mitochondrial Abnormalities in Kidney Disease II)
Show Figures

Figure 1

Figure 1
<p>Pathomechanisms driving chronic kidney disease and novel disease-modifying treatment strategies. Hypertension and hyperglycemia are the main causes of CKD, leading to progressive kidney damage, as depicted by the formation of fibrotic tissue. Abnormal activity of RAAS triggers oxidative stress, inflammatory pathways, and fibrinogenesis. Under hyperglycemic conditions, proximal tubule cells have enhanced energy requirements to sustain the incessant reuptake of glucose through the SGLT2 transporters, resulting in altered mitochondrial activity and ROS generation and finally triggering the hypoxia pathway. Many therapeutic strategies under development aim to ameliorate oxidative stress, to dampen the associated inflammatory response, and to slow down the fibrotic tissue deposition with the use of drugs targeting different pathways, such as MR antagonists, GLP-1R agonists, SGLT2 transporter inhibitors, and HIF-PHD inhibitors. Ang II, angiotensin II; CKD, chronic kidney disease; GLP-1R, glucagon-like peptide 1 receptor; GLP-1RA, glucagon-like peptide 1 receptor agonist; HIF-1α, hypoxia-inducible factor 1α; HIF-PHI, HIF-PHD inhibitor; MR, mineralocorticoid receptor; MRAs, mineralocorticoid receptor antagonists. RAAS, renin–angiotensin–aldosterone system; ROS, reactive oxygen species; SGLT2i, sodium glucose transporter 2-inhibitors; TCA, tricarboxylic acid cycle.</p>
Full article ">
Back to TopTop