[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
Next Issue
Volume 22, March-2
Previous Issue
Volume 22, February-2
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
ijms-logo

Journal Browser

Journal Browser

Int. J. Mol. Sci., Volume 22, Issue 5 (March-1 2021) – 555 articles

Cover Story (view full-size image): Polycystic ovary syndrome (PCOS) is the most common endocrine disorder in reproductive-age women. PCOS is characterized by hyperandrogenism and ovulatory dysfunction. Women with PCOS have a high prevalence of obesity, insulin resistance (IR), increased blood pressure (BP), and activation of the renin angiotensin system (RAS). We hypothesized that hyperandrogenemia upregulates renal sodium-glucose cotransporter-2 (SGLT2) expression and that the SGLT2 inhibitor Empagliflozin ameliorates cardiometabolic complications in a hyperandrogenemic PCOS model. Androgens upregulated renal SGLT2 expression. Empagliflozin attenuated androgen-mediated increases in fat mass, plasma leptin, intra-renal RAS, and BP but failed to decrease plasma insulin, HbA1c, or albuminuria. In summary, SGLT2 inhibition proved beneficial in adiposity and BP reduction in the PCOS model; however, additional therapies may be needed to [...] Read more.
Order results
Result details
Section
Select all
Export citation of selected articles as:
13 pages, 11062 KiB  
Article
Hyaluronic Acid and a Short Peptide Improve the Performance of a PCL Electrospun Fibrous Scaffold Designed for Bone Tissue Engineering Applications
by Dana Rachmiel, Inbar Anconina, Safra Rudnick-Glick, Michal Halperin-Sternfeld, Lihi Adler-Abramovich and Amit Sitt
Int. J. Mol. Sci. 2021, 22(5), 2425; https://doi.org/10.3390/ijms22052425 - 12 Mar 2021
Cited by 26 | Viewed by 5307
Abstract
Bone tissue engineering is a rapidly developing, minimally invasive technique for regenerating lost bone with the aid of biomaterial scaffolds that mimic the structure and function of the extracellular matrix (ECM). Recently, scaffolds made of electrospun fibers have aroused interest due to their [...] Read more.
Bone tissue engineering is a rapidly developing, minimally invasive technique for regenerating lost bone with the aid of biomaterial scaffolds that mimic the structure and function of the extracellular matrix (ECM). Recently, scaffolds made of electrospun fibers have aroused interest due to their similarity to the ECM, and high porosity. Hyaluronic acid (HA) is an abundant component of the ECM and an attractive material for use in regenerative medicine; however, its processability by electrospinning is poor, and it must be used in combination with another polymer. Here, we used electrospinning to fabricate a composite scaffold with a core/shell morphology composed of polycaprolactone (PCL) polymer and HA and incorporating a short self-assembling peptide. The peptide includes the arginine-glycine-aspartic acid (RGD) motif and supports cellular attachment based on molecular recognition. Electron microscopy imaging demonstrated that the fibrous network of the scaffold resembles the ECM structure. In vitro biocompatibility assays revealed that MC3T3-E1 preosteoblasts adhered well to the scaffold and proliferated, with significant osteogenic differentiation and calcium mineralization. Our work emphasizes the potential of this multi-component approach by which electrospinning, molecular self-assembly, and molecular recognition motifs are combined, to generate a leading candidate to serve as a scaffold for bone tissue engineering. Full article
(This article belongs to the Special Issue Molecular Recognition in Biological and Bioengineered Systems)
Show Figures

Figure 1

Figure 1
<p>Schematic illustration of the electrospinning system used to fabricate the nanofibrous scaffolds, and the molecular structures of the different components used.</p>
Full article ">Figure 2
<p>SEM micrographs (top), zoom in on the fibers (middle), and fiber diameter distribution (bottom) for electrospun meshes of (<b>a</b>) polycaprolactone (PCL), (<b>b</b>) PCL + Fmoc-phenylalanine-arginine-glycine-aspartic acid (FmocFRGD), (<b>c</b>) hyaluronic acid (HA)/PCL core/shell, and (<b>d</b>) HA + FmocFRGD/PCL core/shell fibers.</p>
Full article ">Figure 3
<p>Confocal microscopy images of a dry mesh of HA + FmocFRGD/PCL core/shell fibers. (<b>a</b>) The PCL component (blue). (<b>b</b>) The HA + FmocFRGD component (green). (<b>c</b>) The overlay of the two components. (<b>d</b>) The same sample after immersion in water, displaying the transition of the HA + FmocFRGD (green) to the surface of the fibers, leaving the PCL (blue) at the center of the fibers.</p>
Full article ">Figure 4
<p>Characterization and surface properties of HA, FmocFRGD, and PCL electrospun fibers. (<b>a</b>–<b>d</b>) Images of a water droplet on (<b>a</b>) PCL, (<b>b</b>) FmocFRGD + PCL, (<b>c</b>) HA/PCL core/shell, and (<b>d</b>) HA + FmocFRGD/PCL core/shell scaffolds. (<b>e</b>–<b>h</b>) Images of a water droplet after plasma treatment on (<b>e</b>) PCL, (<b>f</b>) FmocFRGD + PCL, (<b>g</b>) HA/PCL core/shell, and (<b>h</b>) HA + FmocFRGD/PCL core/shell scaffolds. (<b>i</b>–<b>l</b>) Images of the electrospun scaffolds (<b>i</b>) PCL, (<b>j</b>) FmocFRGD + PCL, (<b>k</b>) HA/PCL core/shell, and (<b>l</b>) HA + FmocFRGD/PCL core/shell. (<b>m</b>–<b>p</b>) Images of the electrospun scaffolds after 24 h immersion in cell culture media: (<b>m</b>) PCL, (<b>n</b>) FmocFRGD + PCL, (<b>o</b>) HA/PCL core/shell, and (<b>p</b>) HA + FmocFRGD/PCL core/shell. Scale bar = 7 mm.</p>
Full article ">Figure 5
<p>MC3T3-E1 cell viability and spreading on HA + FmocFRGD/PCL core/shell electrospun fibers. (<b>a</b>) Number of viable MC3T3-E1 cells at different time points evaluated by Alamar blue. (<b>b</b>–<b>e</b>) MC3T3-E1 cells cultured for three days on an HA + FmocFRGD/PCL core/shell scaffold and stained with fluorescein diacetate and propidium iodine. (<b>f</b>–<b>g</b>) 3D reconstruction of an HA + FmocFRGD/PCL core/shell scaffold. (<b>h</b>) SEM image of MC3T3-E1 cells on an HA + FmocFRGD/PCL core/shell scaffold. Data were analyzed by using two-way ANOVA. * <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 6
<p>Osteogenesis on HA + FmocFRGD/PCL core/shell electrospun fibers. (<b>a</b>) Quantification of Alkaline Phosphatase (ALP) activity of MC3T3-E1 cells, 16 days after seeding with osteogenic differentiation media. (<b>b</b>) Quantification of calcification by Alizarin red (AR) staining, 16 days after seeding in osteogenic differentiation media. (<b>c</b>, <b>d</b>) Optic microscope images of MC3T3-E1 preosteoblast cells stained with Alizarin red. (<b>c</b>) MC3T3-E1 cells 16 days after seeding on PCL fibers. (<b>d</b>) MC3T3-E1 cells 16 days after seeding on HA + FmocFRGD/PCL core/shell fibers. Data were analyzed by using a one-way ANOVA. * <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 ">
27 pages, 2435 KiB  
Review
TRP Channels as Cellular Targets of Particulate Matter
by Alina Milici and Karel Talavera
Int. J. Mol. Sci. 2021, 22(5), 2783; https://doi.org/10.3390/ijms22052783 - 9 Mar 2021
Cited by 21 | Viewed by 5553
Abstract
Particulate matter (PM) is constituted by particles with sizes in the nanometer to micrometer scales. PM can be generated from natural sources such as sandstorms and wildfires, and from human activities, including combustion of fuels, manufacturing and construction or specially engineered for applications [...] Read more.
Particulate matter (PM) is constituted by particles with sizes in the nanometer to micrometer scales. PM can be generated from natural sources such as sandstorms and wildfires, and from human activities, including combustion of fuels, manufacturing and construction or specially engineered for applications in biotechnology, food industry, cosmetics, electronics, etc. Due to their small size PM can penetrate biological tissues, interact with cellular components and induce noxious effects such as disruptions of the cytoskeleton and membranes and the generation of reactive oxygen species. Here, we provide an overview on the actions of PM on transient receptor potential (TRP) proteins, a superfamily of cation-permeable channels with crucial roles in cell signaling. Their expression in epithelial cells and sensory innervation and their high sensitivity to chemical, thermal and mechanical stimuli makes TRP channels prime targets in the major entry routes of noxious PM, which may result in respiratory, metabolic and cardiovascular disorders. On the other hand, the interactions between TRP channel and engineered nanoparticles may be used for targeted drug delivery. We emphasize in that much further research is required to fully characterize the mechanisms underlying PM-TRP channel interactions and their relevance for PM toxicology and biomedical applications. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic of the structure of (<b>A</b>) TRPA1, (<b>B</b>) TRPV1, (<b>C</b>) TRPM8 and (<b>D</b>) TRPV4 and the representative sites involved in the activation by PM. Modified from [<a href="#B51-ijms-22-02783" class="html-bibr">51</a>] (with permission).</p>
Full article ">Figure 2
<p>Amorphous silica nanoparticles inhibit the chemical activation of TRPV4. Acute response elicited in cultured human airway epithelial (16HBE) cells by different concentration of 10 nm amorphous silica nanoparticles (SiNPs) and the corresponding TRPV4 inhibition ((<b>a</b>): 0 μg/mL, (<b>b</b>): 10 μg/mL, (<b>c</b>): 100 μg/mL, (<b>d</b>): 1000 μg/mL). (<b>e</b>) The influx in Ca<sup>2+</sup> upon SiNPs challenging is concentration-dependent. (<b>f</b>) The inhibition of TRPV4 activation by agonist in the presence of SiNPs is concentration-dependent. Reproduced with permission from [<a href="#B24-ijms-22-02783" class="html-bibr">24</a>].</p>
Full article ">Figure 3
<p>Changes in interleukin 8 (IL-8) mRNA expression as an indicator of inflammation induced by DEP originating from a “black smoker” diesel truck, a diesel exhaust filter regeneration machine, an emission station and NIST SRM 2975. TRPA1 inhibitor HC-030031 reduces the inflammatory effect. The symbols * and # indicate significant induction relative to control and significant inhibition by HC-030031, respectively. Reproduced with permission from [<a href="#B98-ijms-22-02783" class="html-bibr">98</a>].</p>
Full article ">Figure 4
<p>Graphical summary of the activation of TRPA1, TRPC4, TRPM2, TRPM8 and TRPV3 channels by C<sub>60</sub> fullerenes, cigarette smoke (CS), wood smoke (WS), diesel exhaust (DE), ultrafine ambient particles (UFP), coal fly ash (CFA), zinc NPs (ZnNP), lanthanide NPs (LnNP), silica NPs (SiNP) and DOX-coated gold nanorods (AuNP).</p>
Full article ">Figure 5
<p>Graphical summary of the activation of TRPV1, TRPV2 and TRPV4 channels by titanium NPs (TiNP), carbon NPs (CNP), coal fly ash (CFA), silica NPs (SiNP), ultrafine ambient particles (UFP), cigarette smoke (CS), diesel exhaust (DE) and magnetic particles (mag NP).</p>
Full article ">
26 pages, 5689 KiB  
Article
The Leukotriene Receptor Antagonist Montelukast Attenuates Neuroinflammation and Affects Cognition in Transgenic 5xFAD Mice
by Johanna Michael, Julia Zirknitzer, Michael Stefan Unger, Rodolphe Poupardin, Tanja Rieß, Nadine Paiement, Horst Zerbe, Birgit Hutter-Paier, Herbert Reitsamer and Ludwig Aigner
Int. J. Mol. Sci. 2021, 22(5), 2782; https://doi.org/10.3390/ijms22052782 - 9 Mar 2021
Cited by 25 | Viewed by 5648
Abstract
Alzheimer’s disease (AD) is the most common form of dementia. In particular, neuroinflammation, mediated by microglia cells but also through CD8+ T-cells, actively contributes to disease pathology. Leukotrienes are involved in neuroinflammation and in the pathological hallmarks of AD. In consequence, leukotriene signaling—more [...] Read more.
Alzheimer’s disease (AD) is the most common form of dementia. In particular, neuroinflammation, mediated by microglia cells but also through CD8+ T-cells, actively contributes to disease pathology. Leukotrienes are involved in neuroinflammation and in the pathological hallmarks of AD. In consequence, leukotriene signaling—more specifically, the leukotriene receptors—has been recognized as a potential drug target to ameliorate AD pathology. Here, we analyzed the effects of the leukotriene receptor antagonist montelukast (MTK) on hippocampal gene expression in 5xFAD mice, a commonly used transgenic AD mouse model. We identified glial activation and neuroinflammation as the main pathways modulated by MTK. The treatment increased the number of Tmem119+ microglia and downregulated genes related to AD-associated microglia and to lipid droplet-accumulating microglia, suggesting that the MTK treatment targets and modulates microglia phenotypes in the disease model compared to the vehicle. MTK treatment further reduced infiltration of CD8+T-cells into the brain parenchyma. Finally, MTK treatment resulted in improved cognitive functions. In summary, we provide a proof of concept for MTK to be a potential drug candidate for AD and provide novel modes of action via modulation of microglia and CD8+ T-cells. Of note, 5xFAD females showed a more severe pathology, and in consequence, MTK treatment had a more pronounced effect in the females compared to the males. The effects on neuroinflammation, i.e., microglia and CD8+ T-cells, as well as the effects on cognitive outcome, were dose-dependent, therefore arguing for the use of higher doses of MTK in AD clinical trials compared to the approved asthma dose. Full article
(This article belongs to the Special Issue Molecular, Cellular and Systemic Signature of Microglia)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) The leukotriene signaling pathway. Leukotrienes arise from arachidonic acid (AA), which is converted to LTA4 by the enzyme 5-Lipoxygenase (5-Lox) and its activating protein (FLAP). LTA4 is either metabolized to LTB4 by LTA4 hydrolase or to LTC4 by LTC4 synthase. From LTC4, the other cysteinyl leukotrienes, i.e., LTD4 and LTE4, arise. Cysteinyl leukotrienes bind to the receptors CysLTR1, CysLTR2 and GPR17. Leukotriene signaling can be inhibited by targeting the receptors for cysteinyl leukotrienes with the leukotriene receptor antagonist montelukast. (<b>B</b>) Experimental setup. Five-month-old 5xFAD transgenic mice were treated daily with vehicle or montelukast in two different doses for 89 days. Behavioral tests were performed between day 72 and day 89. On the last day of the behavioral tests, the mice were perfused, and tissue samples were collected.</p>
Full article ">Figure 2
<p>MTK alters the gene expression profile in 5xFAD mice. Gene expression data from hippocampal tissue (<span class="html-italic">n</span> = 10/group). Only tissues from the vehicle and high-dose (10 mg/kg/d) groups were used in this analysis. (<b>A</b>) Volcano plot showing differentially expressed genes (DEGs) in high dose- versus vehicle-treated groups. Significantly upregulated DEGs are depicted in red and significantly downregulated DEGs are depicted in blue. Gene names associated with the terms innate immunity (yellow), adaptive immunity (green) and Alzheimer’s disease (AD) (purple) are highlighted. (<b>B</b>) Heatmap showing the top 50 significantly up- and downregulated genes. (<b>C</b>) Genes from GO annotations that fit into the broad terms inflammation and synaptic activity are the top downregulated biological processes. Here, the respective genes are shown in a heatmap.</p>
Full article ">Figure 3
<p>Bioinformatic and immunohistochemical analysis of microglia with and without MTK treatment in the hippocampus. (<b>A</b>) Overlap between genes downregulated with MTK treatment (green) and genes upregulated in lipid droplet-accumulating microglia (red), disease-associated microglia (blue) and AD microglia (yellow). (<b>B</b>) Representative images of microglia positive for Iba1 and Tmem119; scale bar: 20 µm. (<b>C</b>) We observed an increase in numbers of Iba1+ microglia with MTK, which was significant between vehicle and high dose. (<b>D</b>) No significant difference in numbers of Iba1+/Tmem119− cells between groups. (<b>E</b>) We observed a significant increase in numbers of Iba1+/Tmem119+ microglia with MTK. (<b>F</b>) The increase in Iba1+ cells comes from female mice, which have significantly more microglia than male mice of the same age. (<b>G</b>) Numbers of cells separated by sex. Female mice have significantly more Iba1+/Tmem119− cells than male mice. (<b>H</b>) The increase in the subpopulation comes from females, which have less of these cells in the vehicle group but increase to more of these cells in both MTK treatment groups. Data are shown as mean +/− SEM. One-way ANOVA was performed with Tukey’s post hoc test. <span class="html-italic">P</span>-values of <span class="html-italic">p</span> &lt; 0.0001 and <span class="html-italic">p</span> &lt; 0.001 were considered extremely significant (***), <span class="html-italic">p</span> &lt; 0.01 very significant (**) and <span class="html-italic">p</span> &lt; 0.05 significant (*).</p>
Full article ">Figure 4
<p>Analysis of microglia activation. (<b>A</b>) Representative image of Iba1+ microglia positive and negative for Tmem119; scale bar: 5 µm. (<b>B</b>) MTK reduces soma sizes in the Iba1+/Tmem119+ subpopulation in the hippocampus. (<b>C</b>) Sex-specific soma size analysis: effects of MTK on soma size are seen only in females in the hippocampus. (<b>D</b>) Representative image of Iba1+ microglia positive for 5-Lox; scale bar: 5 µm. (<b>E</b>) Quantification of 5-Lox-positive cells. (<b>F</b>) Percentage of 5-Lox+ cells in the Iba1+ microglia population. There was no significant difference between male and female mice in this analysis. Data are shown as mean +/− SEM. One-way ANOVA was performed with Tukey’s post hoc test. <span class="html-italic">P</span>-values of <span class="html-italic">p</span> &lt; 0.01 were considered very significant (**) and <span class="html-italic">p</span> &lt; 0.05 significant (*).</p>
Full article ">Figure 5
<p>Immunohistochemical analysis of CD8+ T-cells in the hippocampus. (<b>A</b>) Representative images of parenchymal CD8+ T-cells and ColIV in the hippocampus. Scale bar: 10 µm (<b>B</b>) Representative image of vessel-associated CD8+ T-cells. Scale bar: 5 µm (<b>C</b>) Number of CD8+ T-cells per mm2 with and without MTK. (<b>D</b>) Distribution of CD8+ T-cells between vessel and parenchyma among all groups. Data are shown as mean +/− SEM. One-way ANOVA was performed with Tukey´s post hoc test. <span class="html-italic">P</span>-values of <span class="html-italic">p</span> &lt; 0.0001 and <span class="html-italic">p</span> &lt; 0.001 were considered extremely significant (****) and <span class="html-italic">p</span> &lt; 0.05 significant (*). Images were created with Zen blue (<b>A</b>) and Imaris (<b>B</b>).</p>
Full article ">Figure 6
<p>Learning behavior in the Barnes maze (BM). Mice were tested in the BM (<span class="html-italic">n</span> = 7–8/group). (<b>A</b>) During the learning phase of the test, mice treated with vehicle performed worse than mice treated with the high dose of MTK. Mice treated with the high dose performed significantly better than the vehicle group on days 3 and 4. (<b>B</b>) Mean area under curve (AUC) of latency on days 3–4 of the BM shows a trend to a dose-dependent decrease (<span class="html-italic">p</span> = 0.07). (<b>C</b>) Mice from the high-dose treatment group had higher frequencies of target contact, which was significant on day 3 between high dose- and vehicle-treated groups. (<b>D</b>) AUC of frequency on days 3–4 of the BM shows a dose-dependent increase, which is significant between vehicle and high-dose treatment groups. (<b>E</b>,<b>F</b>) During the memory test, no significant difference between groups was detected in the parameters of latency and frequency. However, mice had huge motivational problems regardless of their group. (<b>G</b>,<b>H</b>) Mice treated with MTK did show significant longer distances and higher velocities compared to vehicle-treated animals. Data are shown as mean +/− SEM. Two-way ANOVA (<b>A</b>,<b>C</b>) or one-way ANOVA (<b>B</b>,<b>D</b>,<b>E</b>–<b>H</b>) was performed with Tukey´s post hoc test. <span class="html-italic">P</span>-values <span class="html-italic">p</span> &lt; 0.01 were considered very significant (**) and <span class="html-italic">p</span> &lt; 0.05 significant (*).</p>
Full article ">Figure 7
<p>Learning and memory in the Morris water maze (MWM). Mice were tested in the MWM (<span class="html-italic">n</span> = 14–15/group). (<b>A</b>–<b>C</b>) Latency to target contact in the learning phase did not reveal a significant difference between groups (<b>A</b>) when tested regardless of gender. When female mice were analyzed separately (<span class="html-italic">n</span> = 7–8/group), a trend (<span class="html-italic">p</span> = 0.06) to better learning was detected on day 3 between vehicle and high-dose treatment groups (<b>B</b>). Separate analysis of male mice revealed no significant difference between groups amongst males (<b>C</b>). (<b>D</b>–<b>F</b>) Analysis of distance revealed no significant difference between groups (<b>D</b>) when tested regardless of gender. When female mice were analyzed separately, they showed shorter distances, which were significant on day 3 (<b>E</b>). Separate analysis of male mice revealed no significant difference between groups amongst males (<b>F</b>). Memory tests on day 5 revealed no significant difference between groups regardless of gender (pink = female, blue = male) (<b>G</b>,<b>H</b>). Data are shown as mean +/− SEM. Two-way ANOVA (<b>A</b>–<b>F</b>) or one-way ANOVA (<b>G</b>,<b>H</b>) was performed with Tukey’s post hoc test. <span class="html-italic">p</span>-values of <span class="html-italic">p</span> &lt; 0.05 were considered significant (*) and <span class="html-italic">p</span> &lt; 0.07 were considered as trend.</p>
Full article ">Figure 8
<p>MTK has pleiotropic effects on the diseased brain. MTK treatment with 10 mg/kg/d has pleiotropic effects in the brain, i.e., reduced microglia activation and CD8+ T-cell invasion and improvement in cognition. Furthermore, MTK affects gene expression of genes related to synaptic activity.</p>
Full article ">
24 pages, 3128 KiB  
Article
Vitamin D Compounds PRI-2191 and PRI-2205 Enhance Anastrozole Activity in Human Breast Cancer Models
by Beata Filip-Psurska, Mateusz Psurski, Artur Anisiewicz, Patrycja Libako, Ewa Zbrojewicz, Magdalena Maciejewska, Michał Chodyński, Andrzej Kutner and Joanna Wietrzyk
Int. J. Mol. Sci. 2021, 22(5), 2781; https://doi.org/10.3390/ijms22052781 - 9 Mar 2021
Cited by 10 | Viewed by 4208
Abstract
1,25-Dihydroxycholecalciferol, the hormonally active vitamin D3 metabolite, is known to exhibit therapeutic effects against breast cancer, mainly by lowering the expression of estrogen receptors and aromatase activity. Previously, the safety of the vitamin D active metabolite (24R)-1,24-dihydroxycholecalciferol (PRI-2191) and 1,25(OH) [...] Read more.
1,25-Dihydroxycholecalciferol, the hormonally active vitamin D3 metabolite, is known to exhibit therapeutic effects against breast cancer, mainly by lowering the expression of estrogen receptors and aromatase activity. Previously, the safety of the vitamin D active metabolite (24R)-1,24-dihydroxycholecalciferol (PRI-2191) and 1,25(OH)2D3 analog PRI-2205 was tested, and the in vitro activity of these analogs against different cancer cell lines was studied. We determined the effect of the two vitamin D compounds on anastrozole (An) activity against breast cancer based on antiproliferative activity, ELISA, flow cytometry, enzyme inhibition potency, PCR, and xenograft study. Both the vitamin D active metabolite and synthetic analog regulated the growth of not only estrogen receptor-positive cells (T47D and MCF-7, in vitro and in vivo), but also hormone-independent cancer cells such as SKBR-3 (HER-2-positive) and MDA-MB-231 (triple-negative), despite their relatively low VDR expression. Combined with An, PRI-2191 and PRI-2205 significantly inhibited the tumor growth of MCF-7 cells. Potentiation of the antitumor activity in combined treatment of MCF-7 tumor-bearing mice is related to the reduced activity of aromatase by both An (enzyme inhibition) and vitamin D compounds (switched off/decreased aromatase gene expression, decreased expression of other genes related to estrogen signaling) and by regulation of the expression of the estrogen receptor ERα and VDR. Full article
Show Figures

Figure 1

Figure 1
<p>Structures of vitamin D compounds.</p>
Full article ">Figure 2
<p>The proliferation inhibition (<b>A</b>–<b>D</b>) of human breast cancer cell lines after treatment with analogs of 1,25(OH)<sub>2</sub>D<sub>3</sub> alone or in combination with anastrozole. As indicated (<b>A</b>) MCF-7 cell line, (<b>B</b>) T47D cell line, (<b>C</b>) SKBR-3 cell line, and (<b>D</b>) MDA-MB-231 cell line after 120 h of treatment with calcitriol or PRI-2191 and PRI-2205 and anastrozole. Calcitriol and PRI-2191 and PRI-2005 were used at the dose of 100 nM (<b>A</b>,<b>C</b>,<b>D</b>) or 10 nM (<b>B</b>). Anastrozole was used at the dose of 0.1 mg/mL. The tests were performed 3 to 6 times, each in triplicate. The % (mean ± SD) of proliferation inhibition of cancer cells is shown on the graphs. The nonparametric Kruskal-Wallis ANOVA for multiple comparisons was performed. <span class="html-italic">p</span> &lt; 0.05 was considered to be statistically significant; a—compared to anastrozole, b—compared to PRI-2205 alone.</p>
Full article ">Figure 3
<p>Aromatase inhibition by calcitriol and vitamin D compounds PRI-2191 and PRI-2205 (Corning CYP19 inhibition test). The concentrations of the compounds were used according to the manufacturers’ protocol by referring to the concentrations of the positive control used (ketoconazole). Ethanol was used as a solvent for the compounds. The test was performed three times, each time in triplicates. Statistical analysis was performed using the nonparametric Kruskal-Wallis test. <span class="html-italic">p</span> &lt; 0.05 was considered to be statistically significant; * compared to calcitriol.</p>
Full article ">Figure 4
<p>Expression of the estrogen receptors ERα and ERβ in the MCF-7 breast cancer cell line after 72 h of treatment with calcitriol, PRI-2191 and PRI-2205, and anastrozole as measured by flow cytometry. Calcitriol, PRI-2191, and PRI-2205 were used at the concentration of 100 nM and anastrozole at the concentration of 0.1 mg/mL. The test was performed three times, each time in duplicates.</p>
Full article ">Figure 5
<p>Secretion of 17-β estradiol by MCF-7 cells after in vitro treatment with vitamin D analogs and anastrozole Calcitriol, PRI-2191, and PRI-2205 were tested at 100 nM, while anastrozole was tested at 0.1 mg/mL concentration. (<b>A</b>) After 48 h of treatment, the culture medium was replaced with fresh, colorless DMEM medium without the tested compounds. The culture media were collected after additional 24 h. Total time of cell culture = 72 h. The test was performed two times, each time in triplicates. (<b>B</b>) After 96 h of treatment, the medium was replaced with fresh, colorless DMEM medium without the tested compounds. The culture media were collected after additional 24 h. Total time of cell culture = 120 h. The test was performed three times, each time in triplicates. The nonparametric Kruskal-Wallis ANOVA was used for multiple comparisons. <span class="html-italic">p</span> &lt; 0.05 was considered to be statistically significant; * compared to untreated control.</p>
Full article ">Figure 6
<p>Changes in the mRNA profile of MCF-7 cells after treatment with calcitriol or vitamin D analogs with/without anastrozole. The figures presented above represent genes encoding molecules involved in estrogen metabolism. (<b>A</b>) BCAR3—breast cancer antiestrogen resistance protein 3; (<b>B</b>,<b>C</b>) ERα, ERβ—estrogen receptors α and β, respectively; (<b>D</b>) CYP19—aromatase; (<b>E</b>,<b>F</b>) ESRRA, ESRRG—Estrogen Related Receptor Alpha and Gamma, respectively; (<b>G</b>) GNRH1—Gonadotropin-Releasing Hormone 1; (<b>H</b>) SULT1E1—sulfotransferase family 1E, estrogen-preferring member 1; (<b>I</b>) HSD17B2—hydroxysteroid dehydrogenase 17 beta 2. PCR analysis was performed for two independent samples (each sample was tested in triplicate).</p>
Full article ">Figure 7
<p>The antitumor effect of vitamin D analogs when used alone or in combination with anastrozole in MCF-7 tumor-bearing mice. Five days after tumor inoculation (tumor volume was measured), the mice received subcutaneous (s.c.) injections of PRI-2191 (1 μg/kg body weight (b.w.) in each injection) or PRI-2205 (10 μg/kg b.w. in each injection) and/or anastrozole (200 μg/mouse in each injection). PRI-2191 and PRI-2205 were administered three times a week, and the total dose was 17 μg/kg and 170 μg/kg respectively. Anastrozole was administered five times a week, at the total dose of 5.2 mg/mouse. (<b>A</b>,<b>B</b>) tumor growth kinetics after treatment with PRI-2191 (<b>A</b>) or PRI-2205 (<b>B</b>) alone or combined with anastrozole. (<b>C</b>,<b>D</b>) present the tumor growth inhibition (TGI) in % calculated according to the formula given in this article. (<b>C</b>) Red bars indicate TGI for combined PRI-2191 and anastrozole treatment. □—PRI-2191 alone, ▼—anastrozole alone treatment. (<b>D</b>) Green bars indicate TGI for combined PRI-2205 and anastrozole treatment, ○—PRI-2205 alone, ▼—anastrozole alone treatment. (<b>E</b>,<b>F</b>) body weight (b.w.) changes during the treatment in all the experimental groups. (<b>G</b>) The serum calcium level at the end of the experiment. The serum calcium levels were measured at the end of the experiment by using a Roche Diagnostic Kit. Data presented in mmol/L are expressed as mean with SD for each group, except for calcitriol, which was administered subcutaneously to a single NOD/SCID mouse at the dose of 1 µg/kg b.w. (sacrificed after 48 h). To demonstrated the safety of using both vitamin D analogs in comparison to calcitriol, the serum calcium level was evaluated in one mouse. (<b>H</b>) The 17β-estradiol serum levels after treatment. The 17-β estradiol levels were measured using the IBL ELISA kit, which is used for in vitro diagnostic quantitative determination of 17β-estradiol. Statistical analysis was performed using Kruskal–Wallis ANOVA, For calcium and 17β-estradiol level in serum * <span class="html-italic">p</span> &lt; 0.05 compared to the control group.</p>
Full article ">Figure 8
<p>The expression of selected proteins in MCF-7 tumor tissue. NOD/SCID mice were treated with PRI-2191, PRI-2205, and/or anastrozole. Primary antibodies were used in the following concentrations: anti-VDR (sc-1008) 1:400, anti-CYP27B1 (sc-67261) 1:500, and anti-ERα (sc-542) and anti-ERβ (sc-8974) 1:500 (all from Santa Cruz Biotechnology Inc., Santa Cruz, CA, USA) and anti-β-actin (Sigma-Aldrich, Poznań, Poland) 1:500. ECF Western Blotting Reagent Pack (Amersham, GE Healthcare, Little Chalfont, Buckinghamshire, UK) was used for the analysis. Densitometric analysis was performed using ImageJ 1.43u (Wayne Rasband, National Institutes of Health, USA, <a href="https://imagej.nih.gov/ij/" target="_blank">https://imagej.nih.gov/ij/</a>, accessed on 8 March 2021).</p>
Full article ">
21 pages, 4452 KiB  
Article
Microglial Heterogeneity and Its Potential Role in Driving Phenotypic Diversity of Alzheimer’s Disease
by Stefano Sorrentino, Roberto Ascari, Emanuela Maderna, Marcella Catania, Bernardino Ghetti, Fabrizio Tagliavini, Giorgio Giaccone and Giuseppe Di Fede
Int. J. Mol. Sci. 2021, 22(5), 2780; https://doi.org/10.3390/ijms22052780 - 9 Mar 2021
Cited by 12 | Viewed by 4113
Abstract
Alzheimer’s disease (AD) is increasingly recognized as a highly heterogeneous disorder occurring under distinct clinical and neuropathological phenotypes. Despite the molecular determinants of such variability not being well defined yet, microglial cells may play a key role in this process by releasing distinct [...] Read more.
Alzheimer’s disease (AD) is increasingly recognized as a highly heterogeneous disorder occurring under distinct clinical and neuropathological phenotypes. Despite the molecular determinants of such variability not being well defined yet, microglial cells may play a key role in this process by releasing distinct pro- and/or anti-inflammatory cytokines, potentially affecting the expression of the disease. We carried out a neuropathological and biochemical analysis on a series of AD brain samples, gathering evidence about the heterogeneous involvement of microglia in AD. The neuropathological studies showed differences concerning morphology, density and distribution of microglial cells among AD brains. Biochemical investigations showed increased brain levels of IL-4, IL-6, IL-13, CCL17, MMP-7 and CXCL13 in AD in comparison with control subjects. The molecular profiling achieved by measuring the brain levels of 25 inflammatory factors known to be involved in neuroinflammation allowed a stratification of the AD patients in three distinct “neuroinflammatory clusters”. These findings strengthen the relevance of neuroinflammation in AD pathogenesis suggesting, in particular, that the differential involvement of neuroinflammatory molecules released by microglial cells during the development of the disease may contribute to modulate the characteristics and the severity of the neuropathological changes, driving—at least in part—the AD phenotypic diversity. Full article
(This article belongs to the Special Issue Microglia Heterogeneity and Its Relevance for Translational Research)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Microglial characterization in fAD cases. (<b>a</b>,<b>d</b>,<b>g</b>,<b>j</b>,<b>m</b>) APPA673V (fAD1); (<b>b</b>,<b>e</b>,<b>h</b>,<b>k</b>,<b>n</b>) APPA713V (fAD2); (<b>c</b>,<b>f</b>,<b>i</b>,<b>l</b>,<b>o</b>) PS1P117A (fAD3). Scale bar 1 mm (<b>a</b>–<b>c</b>) 200 μm (<b>d</b>–<b>f</b>); 50 μm (<b>g</b>–<b>i</b>). (<b>a</b>–<b>i</b>) Frontal cortex sections immunostained with the IBA1 antibody. (<b>j</b>–<b>l</b>) Morphological shape extracted and edited by imageJ software analysis. (<b>m</b>–<b>o</b>) Quantification of IBA1 immunoreactivity based on the intensity and number of pixels (data were normalized respect to the higher signal obtained among Alzheimer’s disease (AD) samples).</p>
Full article ">Figure 2
<p>Microglial characterization in sAD cases. (<b>a</b>,<b>d</b>,<b>g</b>,<b>j</b>,<b>m</b>) sAD6; (<b>b</b>,<b>e</b>,<b>h</b>,<b>k</b>,<b>n</b>) sAD19; (<b>c</b>,<b>f</b>,<b>i</b>,<b>l</b>,<b>o</b>) sAD21. Scale bar 1 mm (<b>a</b>–<b>c</b>) 200 μm (<b>d</b>–<b>f</b>); 50 μm (<b>g</b>–<b>i</b>). (<b>a</b>–<b>i</b>) Frontal cortex sections immunostained with the IBA1 antibody. Arrows in (a) highlight the bilayer distribution of microglial cells across the frontal cortex. (<b>j</b>–<b>l</b>) Morphological shape extracted and edited by imageJ software analysis. (<b>m</b>–<b>o</b>) Quantification of IBA1 immunoreactivity based on the intensity and number of pixels (data were normalized respect to the higher signal obtained among AD samples).</p>
Full article ">Figure 3
<p>Comparison between AD and control samples. Boxplots of the most significant analytes, IL-4, IL-6, IL-13, CCL17, MMP-7, CXCL13. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.005, *** <span class="html-italic">p</span> &lt; 0.001. For IL-6 the most extreme observations have been omitted for a better graphical representation.</p>
Full article ">Figure 4
<p>Comparison between pro- and anti-inflammatory cytokines in AD and control samples. Boxplot of the pro-inflammatory (IFN-γ, IL-1ɑ, IL-2, IL-6, IL-12 p70, IL-18) and the anti-inflammatory (IL-1rn, IL-4, IL-13) cytokines are shown for AD (<span class="html-italic">p</span>-value &lt; 0.0001) (<b>a</b>) and control (<span class="html-italic">p</span>-value 0.29) (<b>b</b>) group. * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001. Dots indicate outlier values. Natural logarithm scale has been used for a better graphical representation.</p>
Full article ">Figure 5
<p>Families of neuroinflammatory factors. (<b>a</b>) STRING network highlights the direct (physical) and indirect (functional) connection of elements belongs to each family. Known interactions: from curated databases (light blue), experimentally determined (purple); predicted interactions: gene neighborhood (green), gene fusions (red), gene co-occurrence (blue); other interactions: text mining (yellow), co-expression (black), protein homology (grey). (<b>b</b>) Pie graph indicates the relative distribution of cytokines (red), chemokines (green), MMPs (pink) and IIFs (blue) within the AD group. (<b>c</b>) The heat map shows correlations between the four neuroinflammatory classes and clinical, neuropathological and biological data. Aβ42 levels in the insoluble fraction of brain homogenates (Aβ42i); Aβ40 levels in the insoluble fraction of brain homogenates (Aβ40i); Aβ42 levels in the soluble fraction of brain homogenates (Aβ42s); Aβ40 levels in the soluble fraction of brain homogenates (Aβ40s).</p>
Full article ">Figure 6
<p>(<b>a</b>) Hierarchical Cluster Analysis (HCA) shows the different subgroups of AD patients characterized by different expression of neuroinflammatory molecular factors. In abscissa, there are the AD patients and the ordinate corresponds to the complete linkage measured by Euclidean distance. The dashed red line represents the cut off for three clusters (CL): AD-CL1 in yellow; AD-CL2 in red; AD-CL3 in light blue. (<b>b</b>) Histogram concerning the relative abundance of each inflammatory family (cytokines in red, chemokines in green, matrix-metalloproteinases (MMPs) in cream and Innate immunity factors in blue) within each cluster.</p>
Full article ">
19 pages, 2328 KiB  
Article
Plumbagin, a Biomolecule with (Anti)Osteoclastic Properties
by Sevinj Sultanli, Soni Ghumnani, Richa Ashma and Katharina F. Kubatzky
Int. J. Mol. Sci. 2021, 22(5), 2779; https://doi.org/10.3390/ijms22052779 - 9 Mar 2021
Cited by 14 | Viewed by 4217
Abstract
Plumbagin is a plant-derived naphthoquinone that is widely used in traditional Asian medicine due to its anti-inflammatory and anti-microbial properties. Additionally, plumbagin is cytotoxic for cancer cells due to its ability to trigger reactive oxygen species (ROS) formation and subsequent apoptosis. Since it [...] Read more.
Plumbagin is a plant-derived naphthoquinone that is widely used in traditional Asian medicine due to its anti-inflammatory and anti-microbial properties. Additionally, plumbagin is cytotoxic for cancer cells due to its ability to trigger reactive oxygen species (ROS) formation and subsequent apoptosis. Since it was reported that plumbagin may inhibit the differentiation of bone resorbing osteoclasts in cancer-related models, we wanted to elucidate whether plumbagin interferes with cytokine-induced osteoclastogenesis. Using C57BL/6 mice, we unexpectedly found that plumbagin treatment enhanced osteoclast formation and that this effect was most pronounced when cells were pre-treated for 24 h with plumbagin before subsequent M-CSF/RANKL stimulation. Plumbagin caused a fast induction of NFATc1 signalling and mTOR-dependent activation of p70S6 kinase which resulted in the initiation of protein translation. In line with this finding, we observed an increase in RANK surface expression after Plumbagin stimulation that enhanced the responsiveness for subsequent RANKL treatment. However, in Balb/c mice and Balb/c-derived RAW264.7 macrophages, these findings could not be corroborated and osteoclastogenesis was inhibited. Our results suggest that the effects of plumbagin depend on the model system used and can therefore either trigger or inhibit osteoclast formation. Full article
Show Figures

Figure 1

Figure 1
<p>Plumbagin does not change the phenotype of bone marrow-derived macrophages: (<b>A</b>) the influence of increasing concentrations of plumbagin in the presence of 25 ng/mL M-CSF was tested after 48 h by measuring cell viability (<span class="html-italic">n</span> = 3, with triplicates); (<b>B</b>) cell death triggered by increasing plumbagin concentrations was measured after the incubation of the cells with SYTOX after 48 h of treatment in the presence of M-CSF (<span class="html-italic">n</span> = 4); statistical analysis was performed comparing the results to the positive control (M-CSF) in (<b>A</b>) using Friedman, and in (<b>B</b>) using Mann–Whitney test; (<b>C</b>) the effect of 2 µM plumbagin was tested on the expression of the macrophage marker molecules F4/80, CD11b, CD80, CD86 and CD115, respectively. The graphs represent one typical example (<span class="html-italic">n</span> = 3). *: <span class="html-italic">p</span> ≤ 0.05; **: <span class="html-italic">p</span> ≤ 0.01; ****: <span class="html-italic">p</span> ≤ 0.0001.</p>
Full article ">Figure 2
<p>Plumbagin pre-treatment of bone marrow-derived macrophages (BMDMs) facilitates osteoclastogenesis: (<b>A</b>,<b>B</b>) BMDMs were seeded in a 96-well plate and treated with M-CSF or 2 µM plumbagin on day 0. After 24 h, M-CSF/RANKL was added to the wells of M/R and PB-M/R samples. On day 3, the cells were stimulated with M-CSF, M-CSF/RANKL or plumbagin/M-CSF/RANKL, respectively, until the osteoclasts were fully differentiated; (<b>A</b>) shows representative pictures (magnification 100×) and (<b>B</b>) presents the quantification of TRAP stains (<span class="html-italic">n</span> = 6, five of them with duplicates); (<b>C</b>,<b>D</b>) BMDMs were stimulated as described in (<b>A</b>) on 24-well osteo assay surface plates before quantifying the resorbed area on day 10 (<span class="html-italic">n</span> = 3, with duplicates); statistical analysis was carried out comparing the results to the positive control (M-CSF/RANKL) using Friedman test in (<b>B</b>) using a Friedman test and a Mann–Whitney test in (<b>D</b>). *: <span class="html-italic">p</span> ≤ 0.05; ns: <span class="html-italic">p</span> &gt; 0.05.</p>
Full article ">Figure 3
<p>Plumbagin induces the rapid activation of cellular signalling in BMDMs: (<b>A</b>,<b>B</b>) Western blot analysis to test typical osteoclast signalling molecules was performed using antibodies against NFATc1 (<b>A</b>) and molecules of the NF-κB pathway (<b>B</b>). GAPDH and actin were used as loading controls. The blots are typical examples out of three independent experiments; (<b>C</b>) differentiating osteoclasts were analysed on day 3 by RT-PCR to determine the expression of the osteoclastic marker genes <span class="html-italic">Acp5, Ctsk, Ocstamp, Dcstamp and Oscar</span>, respectively (<span class="html-italic">n</span> = 4). Statistical analysis comparing the results to the positive control (M-CSF/RANKL) was performed using Mann–Whitney test; (<b>D</b>,<b>E</b>) BMDM were either pre-stimulated with plumbagin for 24 h before stimulation with 100 ng/mL LPS for 6 h or stimulated directly with LPS. (<b>D</b>) RT-PCR was performed to determine the expression of the osteoclastogenic, pro-inflammatory cytokine genes <span class="html-italic">Il6, Il1 and Tnfa</span>, respectively (<span class="html-italic">n</span> = 4); (<b>E</b>) the supernatants of the samples were subjected to ELISA to determine the levels of produced cytokines (<span class="html-italic">n</span> = 5). Statistical analysis comparing the results to the positive control (LPS) was carried out using a Mann–Whitney test in (<b>D</b>) and (<b>E</b>). *: <span class="html-italic">p</span> ≤ 0.05; **: <span class="html-italic">p</span> ≤ 0.01.</p>
Full article ">Figure 4
<p>Plumbagin inhibits osteoclastogenesis in RAW264.7 macrophages. (<b>A</b>) RAW264.7 cells were treated with increasing concentrations of plumbagin for 24 h and cell viability was measured after incubation with SYTOX. Statistical analysis was performed comparing the results to the control in using a Mann–Whitney test (<span class="html-italic">n</span> = 4); (<b>B</b>,<b>C</b>) and (<b>D</b>) RAW264.7 cells were seeded in a 24 well plate and treated with RANKL or plumbagin (1 µM) /RANKL. On day 3, cells were restimulated with RANKL, plumbagin/RANKL, respectively, until the osteoclasts were fully differentiated; (<b>B</b>) shows representative pictures and (<b>C</b>) the quantification of TRAP stains (<span class="html-italic">n</span> = 3, with duplicates); (<b>D</b>) one day before the TRAP staining was performed, light microscopic pictures (magnification 100×) were taken to document cell viability and cell numbers. *: <span class="html-italic">p</span> ≤ 0.05; **: <span class="html-italic">p</span> ≤ 0.01.</p>
Full article ">Figure 5
<p>Plumbagin impacts ROS production in the presence of M-CSF/RANKL in BMDMs: (<b>A</b>) cells were incubated for 48 h with plumbagin or treated with M-CSF/RANKL before ROS production was analysed in a (iso)luminol-based assay (<span class="html-italic">n</span> = 3); (<b>B</b>) cells were incubated for 24 h with M-CSF/RANKL, plumbagin or pre-treated with plumbagin before the addition of M-CSF/RANKL and ROS production was analysed in a (iso)luminol based assay on day 2 (<span class="html-italic">n</span> = 3); (<b>C</b>) BMDMs were incubated for 6 h, as indicated. Mitochondrial activity was measured using Mitotracker deep red stain for 30 min before FACS analysis (<span class="html-italic">n</span> = 4); (<b>D</b>) BMDMs were pre-treated for 24 h with plumbagin before the addition of M-CSF/RANKL, or treated with only M-CSF and M-CSF/RANKL, respectively. Mitochondrial activity was measured using Mitotracker deep red stain for 30 min before FACS analysis (<span class="html-italic">n</span> = 4); (<b>E</b>) BMDMs were differentiated as indicated and the mitochondrial copy number was determined by comparing mitochondrial and nuclear DNA levels (<span class="html-italic">n</span> = 5); (<b>F</b>) the expression of OxPhos components was analysed in the mitochondrial extracts from cells treated with M-CSF/RANKL or plumbagin pre-stimulated (PB-M/R) cells by Western blot analysis on day 3. The blot is one typical example out of 3 experiments. Statistical analysis was performed comparing the results to the untreated control for (<b>C</b>) and for (<b>F</b>) to the day 0 or positive control (M-CSF/RANKL) using a Mann–Whitney test. *: <span class="html-italic">p</span> ≤ 0.05; **: <span class="html-italic">p</span> ≤ 0.01. ns. <span class="html-italic">p</span> &gt; 0.05.</p>
Full article ">Figure 6
<p>Plumbagin pre-treatment of BMDM facilitates osteoclastogenesis: (<b>A</b>) the expression of the metabolic signalling molecules pAMPK, p4E-BP1 and p70S6K was analysed from whole cell lysates at the indicated time-points of plumbagin stimulation. The blots show one typical example out of 3 experiments; (<b>B</b>) translational activity was measured using a Click-IT assay and the subsequent measurement of the incorporated, Alexa488 click-labelled HPG-containing proteins. Cells were treated for 6 h as indicated before chasing for 30 min with HPG. Untreated unstained and untreated stained cells were used as negative controls for background fluorescence and background translational activity, respectively. Fixed cells and Alexa488 were then measured by FACS analysis (<span class="html-italic">n</span> = 4). Statistical analysis was performed by comparing results to the untreated control using a Friedman test; (<b>C</b>) RANK expression was measured by FACS analysis using a phycoerythrin (PE)-labelled antibody; (<b>D</b>) cells were pre-treated with plumbagin or 10 µM cycloheximide to block protein translation (<span class="html-italic">n</span> = 3). Statistical analysis was implemented by comparing the results to the positive control (M-CSF/RANKL) using Kruskal–Wallis test. * <span class="html-italic">p</span> &gt; 0.05.</p>
Full article ">
23 pages, 1663 KiB  
Review
Prominent Role of Histone Modifications in the Regulation of Tumor Metastasis
by Mariam Markouli, Dimitrios Strepkos, Efthimia K. Basdra, Athanasios G. Papavassiliou and Christina Piperi
Int. J. Mol. Sci. 2021, 22(5), 2778; https://doi.org/10.3390/ijms22052778 - 9 Mar 2021
Cited by 21 | Viewed by 4962
Abstract
Tumor aggressiveness and progression is highly dependent on the process of metastasis, regulated by the coordinated interplay of genetic and epigenetic mechanisms. Metastasis involves several steps of epithelial to mesenchymal transition (EMT), anoikis resistance, intra- and extravasation, and new tissue colonization. EMT is [...] Read more.
Tumor aggressiveness and progression is highly dependent on the process of metastasis, regulated by the coordinated interplay of genetic and epigenetic mechanisms. Metastasis involves several steps of epithelial to mesenchymal transition (EMT), anoikis resistance, intra- and extravasation, and new tissue colonization. EMT is considered as the most critical process allowing cancer cells to switch their epithelial characteristics and acquire mesenchymal properties. Emerging evidence demonstrates that epigenetics mechanisms, DNA methylation, histone modifications, and non-coding RNAs participate in the widespread changes of gene expression that characterize the metastatic phenotype. At the chromatin level, active and repressive histone post-translational modifications (PTM) in association with pleiotropic transcription factors regulate pivotal genes involved in the initiation of the EMT process as well as in intravasation and anoikis resistance, playing a central role in the progression of tumors. Herein, we discuss the main epigenetic mechanisms associated with the different steps of metastatic process, focusing in particular on the prominent role of histone modifications and the modifying enzymes that mediate transcriptional regulation of genes associated with tumor progression. We further discuss the development of novel treatment strategies targeting the reversibility of histone modifications and highlight their importance in the future of cancer therapy. Full article
Show Figures

Figure 1

Figure 1
<p>Epigenetic mechanisms regulate critical checkpoints of metastasis. Metastasis checkpoints are regulated by highly coordinated epigenetic mechanisms. EMT is mainly characterized by loss of E-cadherin expression. This can be achieved through regulation of transcription factors ZEB1, 2, SNAIL, and TWIST1. The p300/CBP complex acetylates the promoter of ZEB1, which further activates BRG1, in order to suppress E-cadherin via histone methylation. Additionally, ZEB1/ZEB2 have the ability tο recruit the CtBP complex, which enables HDAC1 to suppress the E-cadherin gene. At the same time, TGF-β, which is activated by TGFBR2-mediated induction of SNAI1/2, can activate both JMJ3 demethylase and lncRNA-HIT. JMJ3 is part of a loop along with TGF-β and SNAI1, regulating each other’s induction. Histone methyltransferase G9 targets SNAI1 which recruits SIN3A to suppress E-cadherin via deacetylation. Additionally, SET8 monomethylates H4K20 and cooperates with TWIST1 to increase N-cadherin and suppress E-cadherin. The methyltransferase EZH2, part of the PRC2 complex and the MYC-activated miR-9 can also inhibit E-cadherin expression by histone methylation and targeting of the E-cadherin mRNA. During the second step of intravasation and anoikis resistance, cancer cells express anti-apoptotic factors, such as BCL2 as well as mesenchymal markers. FZD7 induces activation of TWIST1 which increases BCL2 and FLIP inhibits anoikis by suppressing caspase-8 production. Moreover, JMJD2B interacts with β-catenin to upregulate the mesenchymal marker vimentin and PHF8 is involved in elevation of integrins expression. In the final step of extravasation and colonization, the cell differentiation <span class="html-italic">NDRG1</span> gene plays a crucial role in reversing EMT and promoting metastasis.</p>
Full article ">Figure 2
<p>Histone acetylation regulates genes involved in cancer progression. Histone acetylation is an activating histone modification that favors gene transcription. H3K4Ac marks are present on the promoters of several genes, <span class="html-italic">GLI1</span>, <span class="html-italic">SMO</span>, <span class="html-italic">FOXF1</span>, <span class="html-italic">Bmi1</span>, and <span class="html-italic">SIRT2</span>, which are upregulated in cancer, favoring tumor progression. Upon SNAI1/2 activation, increased H3K9 acetylation has been observed on the promoter of <span class="html-italic">TGFBR2</span>, causing its upregulation. Critical acetyltransferases TIP60, GCN5, PCAF, and p300 catalyze histone acetylation of metastasis-promoting genes. hMOF, a H4K16 histone acetyltransferase normally maintains the expression of EMT-related tumor suppressor genes, such as <span class="html-italic">TMS1, CDH1</span>, and <span class="html-italic">ESR1</span>, but is often downregulated in cancer. TIP60 is also commonly downregulated in various tumors, where it prevents the activity of anti-tumor DDR and p53 pathways, indicating how reduction in histone acetylation promotes cancer progression. HDAC3 can also be detected on the promoters of activated mesenchymal genes, even though most HDACs are associated with gene repression. Upon Notch signaling activation, H3K4Ac is removed by HDAC3, allowing for Notch-mediated expression of EMT-related genes. Profilin-2 interacts with HDAC1 and inhibits its binding to the promoters of <span class="html-italic">Smad2</span> and <span class="html-italic">Smad3</span>, causing Smad protein activation and subsequently enhancing TGF-β-induced EMT and angiogenesis. HDACs also catalyze H3K4/56 deacetylation at the <span class="html-italic">CDH1</span> promoter, thus repressing E-cadherin. Similarly, HDAC1 and 2 are recruited by ZEB1 to <span class="html-italic">CDH1</span> promoter, inducing its repression. In the same context, ZEB1 recruits SIRT1, a nicotinamide adenine dinucleotide (NAD)-dependent deacetylase, associated with a global H3K27 deacetylation and reduced H3K9Ac and H3K27Ac levels on the promoters of <span class="html-italic">CDH1</span> and other epithelial genes, such as <span class="html-italic">EPCAM</span>, <span class="html-italic">ST14, ESRP1</span>, and <span class="html-italic">RAB25</span>, promoting EMT and metastasis. Finally, overexpression of FLIP leads to the upregulation of the anti-apoptotic Bcl-XL, inhibiting the anoikis apoptotic pathway. HDAC inhibition prevents this outcome by reducing FLIP levels and inducing apoptosis of cancer cells, suggesting that HDACs participate in anoikis resistance of cancer cells through yet unknown mechanisms.</p>
Full article ">Figure 3
<p>Role of histone methylation in gene regulation during tumor progression. Trimethylation of H3K27 and H3K4 has been associated with suppression of epithelial genes or activation of mesenchymal ones, respectively. In this process, the PRMT5-MEP50 complex suppresses <span class="html-italic">CDH1</span> and <span class="html-italic">GAS1</span> by mediating H3K4 and H4R3 methylation. E-cadherin is suppressed by SNAI1-mediated PRC2 trimethylation of H3K27. PRC2 also suppresses <span class="html-italic">KLF2</span> expression. JARID2 recruits both PRC2 and G9a to promote H3K27 and H3K9 methylation at the promoters of <span class="html-italic">CDH1</span> and <span class="html-italic">miR-200</span>. G9a also silences the <span class="html-italic">ep-CAM</span> gene, promoting metastasis. The TRIM33/Smad2/3 complex inhibits the binding of HP1 on the DNA, leading to expression of <span class="html-italic">GSC</span> and <span class="html-italic">MIXL1</span> mesenchymal genes. The PRMT5-MEP50 complex recruits WDR5 and causes H3R2 methylation, activating the EMT-promoting genes <span class="html-italic">VIM</span> and <span class="html-italic">SNAI1/2</span>. Over-expression of the histone demethylase, JARID1A leads to the activation of cyclin D1/E1 and ITGB1 expression, promoting tumor progression. The H3K27 demethylase, UTX, activates several pro-metastatic genes, such as <span class="html-italic">MMP-9/11</span> and <span class="html-italic">SIX1</span>, after interacting with the MLL4 complex, which includes a H3K4 methyltransferase, thus causing enhancement of EMT and metastasis. The PHF8 demethylase upregulates the expression of Vimentin, Integrin, and Rho-associated protein kinase (ROCK) kinase by removing H3K9 methylation marks, thus favoring metastasis. JMJD2B, a H3K9 demethylase, induces the expression of Vimentin, after interacting with β-catenin and also demethylates the Integrin <span class="html-italic">(ITGB2, ITGAM, ITGA9, ITGAB2)</span> gene promoters, further promoting EMT and metastasis. JMJ3, a H3K27 histone demethylase is induced by TGF-β and activates SNAI1 expression to facilitate EMT. EZH2 establishes repressive H3K27me3 marks and downregulates <span class="html-italic">FOXC,</span> inducing metastasis. It also silences the <span class="html-italic">CDH1,</span> <span class="html-italic">AXIN2, NKD1</span>, <span class="html-italic">PPP2R2B</span>, <span class="html-italic">PRICKLE1</span>, <span class="html-italic">SFRP5</span>, <span class="html-italic">RKIP</span>, promoting EMT and cancer metastasis. SETDB1 methyltransferase amplification is responsible for cancer cell growth by methylating and stabilizing the oncogenic p53 mutants. In this context, PRMT6 methyltransferase reduces the expression of p21, promoting tumor cell migration.</p>
Full article ">
17 pages, 1757 KiB  
Review
Bcl-xL: A Focus on Melanoma Pathobiology
by Anna Maria Lucianò, Ana B. Pérez-Oliva, Victoriano Mulero and Donatella Del Bufalo
Int. J. Mol. Sci. 2021, 22(5), 2777; https://doi.org/10.3390/ijms22052777 - 9 Mar 2021
Cited by 21 | Viewed by 4485
Abstract
Apoptosis is the main mechanism by which multicellular organisms eliminate damaged or unwanted cells. To regulate this process, a balance between pro-survival and pro-apoptotic proteins is necessary in order to avoid impaired apoptosis, which is the cause of several pathologies, including cancer. Among [...] Read more.
Apoptosis is the main mechanism by which multicellular organisms eliminate damaged or unwanted cells. To regulate this process, a balance between pro-survival and pro-apoptotic proteins is necessary in order to avoid impaired apoptosis, which is the cause of several pathologies, including cancer. Among the anti-apoptotic proteins, Bcl-xL exhibits a high conformational flexibility, whose regulation is strictly controlled by alternative splicing and post-transcriptional regulation mediated by transcription factors or microRNAs. It shows relevant functions in different forms of cancer, including melanoma. In melanoma, Bcl-xL contributes to both canonical roles, such as pro-survival, protection from apoptosis and induction of drug resistance, and non-canonical functions, including promotion of cell migration and invasion, and angiogenesis. Growing evidence indicates that Bcl-xL inhibition can be helpful for cancer patients, but at present, effective and safe therapies targeting Bcl-xL are lacking due to toxicity to platelets. In this review, we summarized findings describing the mechanisms of Bcl-xL regulation, and the role that Bcl-xL plays in melanoma pathobiology and response to therapy. From these findings, it emerged that even if Bcl-xL plays a crucial role in melanoma pathobiology, we need further studies aimed at evaluating the involvement of Bcl-xL and other members of the Bcl-2 family in the progression of melanoma and at identifying new non-toxic Bcl-xL inhibitors. Full article
(This article belongs to the Special Issue Advances in Bcl-xL Research 2.0)
Show Figures

Figure 1

Figure 1
<p>Schematic representation of the anti-apoptotic Bcl-2 family members. Bcl-2 homology (BH) and transmembrane domains are represented in distinct colours. The hydrophobic pocket is marked in grey.</p>
Full article ">Figure 2
<p>Schematic representation of the alternative spliced variants Bcl-xL, Bcl-xS and Bcl-xβ. Bcl-2 homology (BH) and transmembrane domains are represented in green. Alternative sequence present in Bcl-x(Beta)(Bcl-xβ) is reported in violet. No defined domains are shown in pink.</p>
Full article ">Figure 3
<p>Transcriptional factors and effectors involved in the Bcl-xL regulation.</p>
Full article ">Figure 4
<p>Non-canonical role played by Bcl-xL on melanoma. Bcl-xL is involved in metastasis, angiogenesis, chemoresistance against different chemotherapeutic drugs, and autophagy.</p>
Full article ">
18 pages, 1921 KiB  
Review
Interactions of Lipid Droplets with the Intracellular Transport Machinery
by Selma Yilmaz Dejgaard and John F. Presley
Int. J. Mol. Sci. 2021, 22(5), 2776; https://doi.org/10.3390/ijms22052776 - 9 Mar 2021
Cited by 16 | Viewed by 7533
Abstract
Historically, studies of intracellular membrane trafficking have focused on the secretory and endocytic pathways and their major organelles. However, these pathways are also directly implicated in the biogenesis and function of other important intracellular organelles, the best studied of which are peroxisomes and [...] Read more.
Historically, studies of intracellular membrane trafficking have focused on the secretory and endocytic pathways and their major organelles. However, these pathways are also directly implicated in the biogenesis and function of other important intracellular organelles, the best studied of which are peroxisomes and lipid droplets. There is a large recent body of work on these organelles, which have resulted in the introduction of new paradigms regarding the roles of membrane trafficking organelles. In this review, we discuss the roles of membrane trafficking in the life cycle of lipid droplets. This includes the complementary roles of lipid phase separation and proteins in the biogenesis of lipid droplets from endoplasmic reticulum (ER) membranes, and the attachment of mature lipid droplets to membranes by lipidic bridges and by more conventional protein tethers. We also discuss the catabolism of neutral lipids, which in part results from the interaction of lipid droplets with cytosolic molecules, but with important roles for both macroautophagy and microautophagy. Finally, we address their eventual demise, which involves interactions with the autophagocytotic machinery. We pay particular attention to the roles of small GTPases, particularly Rab18, in these processes. Full article
(This article belongs to the Special Issue Intracellular Membrane Transport: Models and Machines)
Show Figures

Figure 1

Figure 1
<p>Comparison of the formation of coatomer protein II (COPII) vesicles with biogenesis of LDs. (<b>A</b>) COPII polymerizes laterally to form coated buds, with cargo proteins being sorted into the vesicle through interactions of their cytoplasmic domains with the COPII coat. Coat assembly and sorting take place primarily through protein–protein interactions. (<b>B</b>) Lipid droplet (LD) formation begins with the synthesis of neutral lipids: triglycerides by diacylglycerol O-acetyltransferase 1 (DGAT1) or cholesterol by acetyl-CoA acetyltransferase (ACAT)1/2. Neutral lipids condense into lenses in ER membranes, which then bud to form LDs. DGAT2 and other LD-resident proteins with short hydrophobic loops that are stable in a lipid monolayer can move to LDs while still attached to ER. These processes are driven primarily by phase separation and biophysical processes of lipids, but with proteins playing important modulatory roles.</p>
Full article ">Figure 2
<p>Interaction of LDs with other membranes. (<b>A</b>) Illustration of a direct connection. LDs can remain attached to the ER by direct membranous connections stabilized by seipin rings. These connections permit exchange of proteins such as lipid droplet assembly factor 1 (LDAF1) between LDs and ER membranes. (<b>B</b>) Examples of protein complexes which may tether LDs to ER in the absence of direct membranous connections, including SNX14 and the Rab18/NRZ complex. While Rab18 is required for recruitment of fatty acid synthase to LDs, the link representing binding between fatty acid synthetase (FAS) and Rab18 is speculative. (<b>C</b>) Examples of proteins and protein complexes tethering LDs to mitochondria. At least two perilipins (PLIN1 and PLIN5) are involved. The mitoguardin 2 (MIGA2) complex may be involved in forming three-way complexes involving mitochondria and LDs linked to ER via VAP-B.</p>
Full article ">Figure 3
<p>Triglyceride catabolism by cytoplasmic lipases on the LD surface. ATGL catalyzes release of the first fatty acid, followed by HSL and MAG in sequence. While ATGL and MAG seem to have narrow specificity as shown here, HSL has some activity towards TAG.</p>
Full article ">Figure 4
<p>Catabolism of LDs by autophagy. (<b>A</b>) Macroautophagy is a process by which the entire LD can be enveloped by a phagophore or isolation membrane which seals to become an autophagosome. The autophagosome then undergoes a maturation process and fuses with a lysosome, leading to digestion of the LD and release of free fatty acid, glycerol and cholesterol. (<b>B</b>) Microautophagy of LDs has also been reported. While this process is not currently well understood, it may involve the internalization of a small portion of the LD into an internal vesicle in the lysosome. This process may allow autophagy of lipid from LDs which are too large for macroautophagy.</p>
Full article ">
15 pages, 2842 KiB  
Article
Deletion of Irf4 in T Cells Suppressed Autoimmune Uveitis and Dysregulated Transcriptional Programs Linked to CD4+ T Cell Differentiation and Metabolism
by Minkyung Kang, Hyun-Su Lee, Jin Kyeong Choi, Cheng-Rong Yu and Charles E. Egwuagu
Int. J. Mol. Sci. 2021, 22(5), 2775; https://doi.org/10.3390/ijms22052775 - 9 Mar 2021
Cited by 3 | Viewed by 3312
Abstract
Interferon regulatory factor-4 (IRF4) and IRF8 regulate differentiation, growth and functions of lymphoid and myeloid cells. Targeted deletion of irf8 in T cells (CD4-IRF8KO) has been shown to exacerbate colitis and experimental autoimmune uveitis (EAU), a mouse model of human uveitis. We therefore [...] Read more.
Interferon regulatory factor-4 (IRF4) and IRF8 regulate differentiation, growth and functions of lymphoid and myeloid cells. Targeted deletion of irf8 in T cells (CD4-IRF8KO) has been shown to exacerbate colitis and experimental autoimmune uveitis (EAU), a mouse model of human uveitis. We therefore generated mice lacking irf4 in T cells (CD4-IRF4KO) and investigated whether expression of IRF4 by T cells is also required for regulating T cells that suppress autoimmune diseases. Surprisingly, we found that CD4-IRF4KO mice are resistant to EAU. Suppression of EAU derived in part from inhibiting pathogenic responses of Th17 cells while inducing expansion of regulatory lymphocytes that secrete IL-10 and/or IL-35 in the eye and peripheral lymphoid tissues. Furthermore, CD4-IRF4KO T cells exhibit alterations in cell metabolism and are defective in the expression of two Ikaros zinc-finger (IKZF) transcription factors (Ikaros, Aiolos) that are required for lymphocyte differentiation, metabolism and cell-fate decisions. Thus, synergistic effects of IRF4 and IkZFs might induce metabolic reprogramming of differentiating lymphocytes and thereby dynamically regulate relative abundance of T and B lymphocyte subsets that mediate immunopathogenic mechanisms during uveitis. Moreover, the diametrically opposite effects of IRF4 and IRF8 during EAU suggests that intrinsic function of IRF4 in T cells might be activating proinflammatory responses while IRF8 promotes expansion of immune-suppressive mechanisms. Full article
(This article belongs to the Section Molecular Immunology)
Show Figures

Figure 1

Figure 1
<p>Generation and characterization of mice with targeted deletion of <span class="html-italic">irf4</span> in T cell. (<b>A</b>) <span class="html-italic">Irf4</span><sup>fl/fl</sup> mice were crossed with CD4-Cre mice to generate mice deficient of IRF4 in CD4<sup>+</sup> T cells (CD4-IRF4KO). The homozygote KO mice (CD4<sup>cre</sup>IRF4<sup>fl/fl</sup>) were identified by PCR analysis of mouse tail genomic DNA. (<b>B</b>,<b>C</b>) Purified CD4<sup>+</sup> T cells or CD19<sup>+</sup> B cells from lymph nodes or spleen of wild-type (WT) or CD4-IRFK4O mice were analyzed for IRF4 expression by Western blot (<b>B</b>) or RT-PCR (<b>C</b>) analysis. Images of the full-length Western blot gels are provided (<a href="#app1-ijms-22-02775" class="html-app">Supplementary Figure S1</a>). (<b>D</b>) CD4<sup>+</sup> T cells from LN and spleen of WT or CD4-IRF4KO mice were stimulated with anti-CD3/anti-CD28 for 3 days and analyzed by RT-PCR. (<b>E</b>) CD4<sup>+</sup> T cells from LN/spleen and splenic CD19<sup>+</sup> B cells of WT or CD4-IRF4KO mice were isolated 21 days after immunization with IRBP/CFA and analyzed by intracellular cytokine staining assay. Results represent more than 3 independent studies **** <span class="html-italic">p</span> &lt; 0.0001. N.S.: Not significant</p>
Full article ">Figure 2
<p>Inducible deletion of IRF4 in CD4<sup>+</sup> T cells confers protection from experimental autoimmune uveitis (EAU). EAU was induced by immunization of C57BL/6J or CD4-IRF4KO mice with the uveitogenic peptide, IRBP<sup>651−670</sup> in complete Freund’s adjuvant (CFA) (n = 12). Disease progression was monitored and assessed by fundoscopy, histology, optical coherence tomography (OCT) and electroretinography (ERG). (<b>A</b>) Fundus images of the retina at day 21 after EAU induction were taken using an otoendoscopic imaging system. Black arrow indicates inflammation with blurred optic disc margins and enlarge juxtapapillary areas; blue arrows indicate retinal vasculitis; white arrows indicate yellow-whitish retinal and choroidal infiltrates. Severe uveitis was observed in the WT compared to the CD4-IRF4KO mouse eyes as indicated by the clinical EAU scores. (<b>B</b>) Histologic sections are of eyes harvested 21 days after immunization with IRBP peptide and H&amp;E staining reveal substantial numbers of inflammatory cells in the vitreous of WT compared to the CD4-IRF4KO eyes. EAU in the WT mice is characterized by the development massive retinal in-folding (*), a hallmark feature of severe uveitis. V, vitreous; GCL, ganglion cell layer; INL, inner nuclear layer; ONL, outer nuclear layer; RPE retinal pigmented epithelial layer. Black arrows, lymphocytes; Asterisks, retinal folds. (<b>C</b>) Representative OCT images show marked increase of inflammatory cells (white arrows) in the vitreous and optic nerve (OPN). (<b>D</b>) ERG analysis of the retina on day 20 after EAU induction. The averages of light- or dark-adapted ERG a-wave or b-wave amplitudes are plotted as a function of flash luminance. Data presented as the mean ± SEM of four mice in each group of at least three independent experiments. (<b>E</b>) Cells from lymph nodes and spleen of IRBP<sub>651-670</sub> immunized WT or CD4-IRF4KO mice were re-stimulated ex vivo and 1 × 10<sup>7</sup> cells were adoptively transferred to naïve WT mice. Disease scores determined by masked investigators indicate reduced EAU symptoms in mice that received CD4-IR4KO cells. Results represent 3 independent studies. * <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; 0.0001.</p>
Full article ">Figure 3
<p>Th1 and Th17 cells are reduced in CD4-IRF4KO mice during EAU. CD4<sup>+</sup> T cells from the eye, lymph nodes or spleen of WT and CD4-IRF4KO mice immunized with IRBP/CFA were sorted and analyzed by the intracellular cytokine staining assay. Numbers in quadrants represent percentage CD4<sup>+</sup> T cells expressing IL-17 and/or IFN-γ (<b>A</b>), ROR-γt (<b>B</b>) or ROR-γt and IL-17 (<b>C</b>). Data represent at least three independent experiments and were analyzed using Student’s <span class="html-italic">t</span>-test (two-tailed). * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001. ****: <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 4
<p>Regulatory T and B cells that produce IL-10 or IL-35 are expanded during EAU in CD4-IRF4KO mice. (<b>A</b>) Lymphocytes in the eyes of WT or CD4-IRF4KO mice immunized with IRBP/CFA were stained with CD4 or B220 Abs and numbers in quadrants represent percent B cells and T cells. (<b>B</b>–<b>D</b>) CD4<sup>+</sup> T cells in eyes or lymph nodes (<b>B</b>,<b>C</b>) or B220<sup>+</sup> B cells in the spleen of EAU mice (<b>D</b>) were analyzed by intracellular cytokine staining assay. Numbers in quadrants represent percent CD4<sup>+</sup> T cells expressing IL-10 (<b>B</b>) or IL-35 (<b>C</b>) or B cells secreting IL-35 (<b>D</b>). Data represent three independent experiments. Analysis was performed by Student’s <span class="html-italic">t</span>-test (two-tailed). * <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; 0.0001.</p>
Full article ">Figure 5
<p>Mice lacking IRF4 in CD4<sup>+</sup> T cells exhibit altered expression of Ikaros Zinc Finger (IKZF) and Th17 signature genes. RNA was isolated from control (WT) and CD4-IRF4KO CD4<sup>+</sup> T cells at day 21 post-immunization with IRBP/CFA, and subjected to RNA-Seq analyses. (<b>A</b>,<b>B</b>) Heatmap derived from global RNA-Seq analysis identified upregulated and downregulated genes in IRF4-deficient T cells. (<b>C</b>) KEGG pathway analysis shows effects of IRF4 on the transcription of genes that regulate Th1, Th2 or Th17 differentiation. (<b>D</b>,<b>E</b>) Heatmaps reveal differential expression of Th17 signature (<b>D</b>) and IKZF (<b>E</b>) genes by WT and CD4-IRF4KO T cells. (<b>F</b>) RNA was isolated from control (WT) and CD4-IRF4KO T cells at day 21 post immunization with IRBP/CFA and subjected to qPCR analyses. (<b>G</b>) A Gene Ontology enrichment analysis of the differentially expressed genes was performed to evaluate enriched biological processes. Data represent at least three independent experiments. ** <span class="html-italic">p</span> &lt; 0.01; **** <span class="html-italic">p</span> &lt; 0.0001. ns: Not significant</p>
Full article ">Figure 6
<p>Loss of IRF4 in CD4<sup>+</sup> T cells induced changes in cell metabolism during EAU. (<b>A</b>,<b>B</b>) CD4<sup>+</sup> T cells derived from WT and CD4-IRF4KO mice during EAU were subjected to KEGG pathway analysis to determine whether loss of IRF4 expression in CD4<sup>+</sup> T cells correlates with changes in the expression of metabolic gene. (<b>B</b>) Representative heatmap showing alterations in expression of metabolic pathway genes of WT and CD4-IRF4KO T cells. (<b>C</b>,<b>D</b>) Purified CD4<sup>+</sup> T cells were isolated from WT or CD4-IRF4KO EAU mice (<b>C</b>) Images of the full-length Western blot gels are provided (<a href="#app1-ijms-22-02775" class="html-app">Supplementary Figure S2</a>). Glycolic assays were performed using the Seahorse Glycolytic Rate assay (<b>D</b>). Results are presented as time-dependent changes in oxygen consumption rate (OCR). Data represent three independent experiments.</p>
Full article ">
25 pages, 8426 KiB  
Article
Synthesis and Pharmacological In Vitro Investigations of Novel Shikonin Derivatives with a Special Focus on Cyclopropane Bearing Derivatives
by Nadine Kretschmer, Antje Hufner, Christin Durchschein, Katrin Popodi, Beate Rinner, Birgit Lohberger and Rudolf Bauer
Int. J. Mol. Sci. 2021, 22(5), 2774; https://doi.org/10.3390/ijms22052774 - 9 Mar 2021
Cited by 9 | Viewed by 3301
Abstract
Melanoma is the deadliest form of skin cancer and accounts for about three quarters of all skin cancer deaths. Especially at an advanced stage, its treatment is challenging, and survival rates are very low. In previous studies, we showed that the constituents of [...] Read more.
Melanoma is the deadliest form of skin cancer and accounts for about three quarters of all skin cancer deaths. Especially at an advanced stage, its treatment is challenging, and survival rates are very low. In previous studies, we showed that the constituents of the roots of Onosma paniculata as well as a synthetic derivative of the most active constituent showed promising results in metastatic melanoma cell lines. In the current study, we address the question whether we can generate further derivatives with optimized activity by synthesis. Therefore, we prepared 31, mainly novel shikonin derivatives and screened them in different melanoma cell lines (WM9, WM164, and MUG-Mel2 cells) using the XTT viability assay. We identified (R)-1-(1,4-dihydro-5,8-dihydroxy-1,4-dioxonaphthalen-2-yl)-4-methylpent-3-enyl 2-cyclopropyl-2-oxoacetate as a novel derivative with even higher activity. Furthermore, pharmacological investigations including the ApoToxGloTM Triplex assay, LDH assay, and cell cycle measurements revealed that this compound induced apoptosis and reduced cells in the G1 phase accompanied by an increase of cells in the G2/M phase. Moreover, it showed hardly any effects on the cell membrane integrity. However, it also exhibited cytotoxicity against non-tumorigenic cells. Nevertheless, in summary, we could show that shikonin derivatives might be promising drug leads in the treatment of melanoma. Full article
(This article belongs to the Special Issue Skin Inflammation Aging and Diseases)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Acylation of shikonin (<b>1</b>).</p>
Full article ">Figure 2
<p>Synthesis of the bicyclic acetic acid <b>p1</b>.</p>
Full article ">Figure 3
<p>Synthesis of 3-cyclopropylpropanoic acid (<b>p2</b>).</p>
Full article ">Figure 4
<p>Synthesis of precursor acids <b>p3</b> and <b>p4</b>.</p>
Full article ">Figure 5
<p>Results of the XTT assay. For clarity reasons, only the results of the treatment with 5.0 µM for 72 h are shown. The complete results can be found in the <a href="#app1-ijms-22-02774" class="html-app">Supplementary Material (Figures S1 and S2)</a>. <b>2</b> was tested as a reference at 5.0 µM. The strongest cytotoxicity was found for <b>5</b> (<span class="html-italic">n</span> = 6, mean ± sem). Vinblastine was used as positive control. At a concentration of 0.01 nM, it reduced the cell viability compared to control cells to: WM9 cells: 23.8 ± 1.5%, WM164 cells: 59.4 ± 4.4 %, and MUG-Mel2 cells: 65.0 ± 6.9 % (<span class="html-italic">n</span> = 6, mean ± sem, 72 h of treatment).</p>
Full article ">Figure 6
<p>Results of the ApoToxGlo™ Triplex Assay. WM9 and WM164 cells were treated with 1 µM, 5 µM, 10 µM, and 20 µM of <b>5</b> for 4 h, 24 h, or 48 h (<span class="html-italic">n</span> = 6, mean ± sem). (<b>A</b>) Viability of the cells measured as relative fluorescence of control cells. (<b>B</b>) Cytotoxicity of <b>5</b> towards the cells measured as relative fluorescence of control cells. (<b>C</b>) Activity of caspases 3 and 7 indicative for apoptosis induction. Staurosporine (25 µM) served as positive control (apoptosis increase in WM9 cells after 24 h: 1193.9 % and after 48 h: 297.4%; in WM164 cells after 24 h: 989.1% and after 48 h: 362.6%).</p>
Full article ">Figure 7
<p>Results of the LDH assay. WM9 and WM164 cells were treated with 1 µM, 5 µM, 10 µM, and 20 µM of <b>5</b> for 24 h, 48 h, or 72 h (<span class="html-italic">n</span> = 9, mean ± sem). Results are displayed as percentage of cell lysis. Control = vehicle treated cells (0.5% EtOH). Only at 20 µM, slight increases in LDH release were found.</p>
Full article ">Figure 8
<p>Effects of <b>5</b> on the cell cycle. (<b>A</b>) WM9 and (<b>B</b>) WM164 cells were treated with 10 µM and 20 µM of <b>5</b> for 24 h or 48 h (<span class="html-italic">n</span> = 6, mean). Results are displayed as percentage of cell cycle distribution. Control = vehicle treated cells (0.5% EtOH). <b>5</b> influence the distribution of the cell cycle only at high concentrations.</p>
Full article ">
20 pages, 4204 KiB  
Article
Stress and Nasal Allergy: Corticotropin-Releasing Hormone Stimulates Mast Cell Degranulation and Proliferation in Human Nasal Mucosa
by Mika Yamanaka-Takaichi, Yukari Mizukami, Koji Sugawara, Kishiko Sunami, Yuichi Teranishi, Yukimi Kira, Ralf Paus and Daisuke Tsuruta
Int. J. Mol. Sci. 2021, 22(5), 2773; https://doi.org/10.3390/ijms22052773 - 9 Mar 2021
Cited by 9 | Viewed by 11452
Abstract
Psychological stress exacerbates mast cell (MC)-dependent inflammation, including nasal allergy, but the underlying mechanisms are not thoroughly understood. Because the key stress-mediating neurohormone, corticotropin-releasing hormone (CRH), induces human skin MC degranulation, we hypothesized that CRH may be a key player in stress-aggravated nasal [...] Read more.
Psychological stress exacerbates mast cell (MC)-dependent inflammation, including nasal allergy, but the underlying mechanisms are not thoroughly understood. Because the key stress-mediating neurohormone, corticotropin-releasing hormone (CRH), induces human skin MC degranulation, we hypothesized that CRH may be a key player in stress-aggravated nasal allergy. In the current study, we probed this hypothesis in human nasal mucosa MCs (hM-MCs) in situ using nasal polyp organ culture and tested whether CRH is required for murine M-MC activation by perceived stress in vivo. CRH stimulation significantly increased the number of hM-MCs, stimulated both their degranulation and proliferation ex vivo, and increased stem cell factor (SCF) expression in human nasal mucosa epithelium. CRH also sensitized hM-MCs to further CRH stimulation and promoted a pro-inflammatory hM-MC phenotype. The CRH-induced increase in hM-MCs was mitigated by co-administration of CRH receptor type 1 (CRH-R1)-specific antagonist antalarmin, CRH-R1 small interfering RNA (siRNA), or SCF-neutralizing antibody. In vivo, restraint stress significantly increased the number and degranulation of murine M-MCs compared with sham-stressed mice. This effect was mitigated by intranasal antalarmin. Our data suggest that CRH is a major activator of hM-MC in nasal mucosa, in part via promoting SCF production, and that CRH-R1 antagonists such as antalarmin are promising candidate therapeutics for nasal mucosa neuroinflammation induced by perceived stress. Full article
(This article belongs to the Section Molecular Biology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Human nasal mucosa mast cells (hM-MCs) and the mucosal epithelium in nasal polyps (NPs) express corticotropin-releasing hormone receptor type 1 (CRH-R1). (<b>a</b>) CRH-R1 expression on c-Kit+ or (<b>b</b>) tryptase+ hM-MCs in situ. An arrow denotes double+ MCs. (<b>c</b>) Percentage of tryptase and CRH-R1 or corticotropin-releasing hormone receptor type 2 (CRH-R2) double+ MCs. <span class="html-italic">n</span> = 5. (<b>d</b>) CRH expression within the mucosal epithelium in NPs (an arrow). (<b>e</b>) CRH-R1 and CRH-R2 expression within the mucosal epithelium in NPs. Scale bar = 20 µm. Error bars indicate the standard error of the mean (SEM). ** <span class="html-italic">p</span> &lt; 0.01. MC, mast cell; hM-MCs, human nasal mucosa MCs; NPs, nasal polyps; CRH, corticotropin-releasing hormone; CRH-R1, CRH receptor type 1; CRH-R2, CRH receptor type 2; DAPI, diamidino-2-phenylindole.</p>
Full article ">Figure 2
<p>CRH increased tryptase+ hM-MC numbers and stimulated their degranulation. (<b>a</b>) Toluidine blue histochemistry with 1 day organ-cultured human NPs treated with vehicle or CRH (10<sup>−7</sup> M). An arrow denotes hM-MCs in the lamina propria. Top, non-degranulated MCs in vehicle-treated NPs. Bottom, degranulated MCs in CRH-treated NPs. <span class="html-italic">n</span> = 5; scale bar = 20 µm. (<b>b</b>) Tryptase immunohistochemistry with 1 day organ-cultured human NPs treated with vehicle or CRH (10<sup>−7</sup> M). Scale bar = 50 µm. (<b>c</b>) Quantitative immunohistomorphometry of tryptase+ cells in the lamina propria of organ-cultured NPs with CRH, anti-SCF (stem cell factor neutralizing antibody; 1 μg/mL), and/or antalarmin (10<sup>−6</sup> M); <span class="html-italic">n</span> = 6. (<b>d</b>) Tryptase immunohistochemistry. An arrow denotes MC degranulation induced by CRH. Scale bar = 10 µm. (<b>e</b>) Percentage of degranulated MCs treated with CRH, anti-SCF (1 μg/mL), and/or antalarmin (10<sup>−6</sup> M) analyzed by tryptase immunohistochemistry; <span class="html-italic">n</span> = 6. (<b>f</b>) Electron microscope images of CRH-induced MC degranulation. Scale bar = 1 µm. (<b>g</b>) Diameter of tryptase+ hM-MC granules; <span class="html-italic">n</span> = 6; scale bar = 10 µm. Error bars indicate SEM. * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001. MC, mast cell; hM-MCs, human nasal mucosa MCs; NPs, nasal polyps; CRH, corticotropin-releasing hormone; anti-SCF, stem cell factor neutralizing antibody.</p>
Full article ">Figure 3
<p>CRH increased tryptase+ hM-MC proliferation but not the apoptosis of hM-MCs. CRH-R1 gene silencing showed that CRH-R1 small interfering RNA (siRNA) inhibited the increase in tryptase+ hM-MC number and degranulation by CRH in situ. (<b>a</b>) Tryptase/Ki-67 double immunofluorescence. The arrow denotes tryptase and Ki-67 double+ cells within human nasal polyps treated with CRH; <span class="html-italic">n</span> = 6; scale bar = 10 µm. (<b>b</b>) Quantitative immunohistomorphometry of tryptase/proliferating cell nuclear antigen (PCNA) double+ cells; <span class="html-italic">n</span> = 4. (<b>c</b>) Quantitative immunohistomorphometry of tryptase/terminal deoxynucleotidyl transferase-mediated dUTP nick end-labeling (TUNEL) double+ cells; <span class="html-italic">n</span> = 4. (<b>d</b>) Quantitative immunohistomorphometry of tryptase+ cells in the lamina propria of organ-cultured NPs with CRH-R1 siRNA-treated human NPs; <span class="html-italic">n</span> = 4. (<b>e</b>) Percentage of degranulated hM-MCs in the lamina propria with CRH-R1 siRNA-treated human NPs; <span class="html-italic">n</span> = 4. Error bars indicate SEM. * <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, N.S. = not significant. MC, mast cell; hM-MCs, human nasal mucosa MCs; NPs, nasal polyps; CRH, corticotropin-releasing hormone; CRH-R1, CRH receptor type 1; CRH-R1 siRNA, CRH-R1 siRNA-treated NPs; SCR, scrambled siRNA-treated NPs; PCNA, proliferating cell nuclear antigen; TUNEL, terminal deoxynucleotidyl transferase-mediated dUTP nick end-labeling.</p>
Full article ">Figure 4
<p>CRH increased stem cell factor (SCF) expression in the mucosal epithelium in NPs. (<b>a</b>) SCF immunofluorescence of organ-cultured human NPs treated with vehicle or CRH (10<sup>−7</sup> M). CRH significantly increased SCF expression within the epithelium of 1 day organ-cultured human NPs. The arrow denotes SCF-positive immunoreactivity within the epithelium. <span class="html-italic">n</span> = 6; scale bar = 50 µm. (<b>b</b>) Double immunofluorescence for AE1/3/SCF and (<b>c</b>) tryptase/SCF of CRH-treated NPs. Most of the SCF+ cells were epithelial cells. Some tryptase+ hM-MCs within the epithelium expressed SCF. The arrow denotes SCF+ hM-MCs; <span class="html-italic">n</span> = 4 in (<b>b</b>) and 6 in (<b>c</b>). Scale bar = 10 µm in (<b>b</b>), 20 µm in (<b>c</b>). (<b>d</b>) SCF immunoreactivity in NP epithelium with CRH-R1 siRNA-treated human NPs. The arrow denotes SCF-positive immunoreactivity within the epithelium. <span class="html-italic">n</span> = 4; scale bar = 20 µm. Error bars indicate SEM. *<span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01. MC, mast cell; SCF, stem cell factor; hM-MCs, human nasal mucosa MCs; NPs, nasal polyps; CRH, corticotropin-releasing hormone; CRH-R1, CRH receptor type 1; CRH-R1 siRNA, CRH-R1 siRNA-treated NPs; SCR, scrambled siRNA-treated NPs; DAPI, diamidino-2-phenylindole.</p>
Full article ">Figure 5
<p>CRH increased the expression of CRH-R1 on tryptase+ hM-MCs and upregulated tryptase immunoreactivity within the hM-MCs. Percentage of (<b>a</b>) tryptase+ and (<b>b</b>) c-Kit+ MCs with CRH-R1 in the lamina propria of organ-cultured human NPs, analyzed by double immunofluorescence of tryptase or c-Kit/CRH-R1. <span class="html-italic">n</span> = 7 in (<b>a</b>) and 4 in (<b>b</b>). (<b>c</b>) Tryptase immunoreactivity within the hM-MCs in NPs treated with vehicle or CRH. CRH treatment significantly increased tryptase immunoreactivity within hM-MCs. <span class="html-italic">n</span> = 5; scale bar = 10 µm. Error bars indicate SEM. * <span class="html-italic">p</span> &lt; 0.05, N.S. = not significant. MC, mast cell; hM-MCs, human nasal mucosa MCs; NPs, nasal polyps; CRH, corticotropin-releasing hormone; CRH-R1, CRH receptor type 1.</p>
Full article ">Figure 6
<p>Restraint stress increased mice nasal mucosa MC (mM-MC) numbers and degranulation. (<b>a</b>) Toluidine blue histochemistry of murine nasal mucosa. The arrow denotes mM-MCs in the lamina propria, and the arrowhead denotes a degranulated mM-MC. Scale bar = 20 µm. (<b>b</b>) Quantitative immunohistomorphometry of mM-MC. (<b>c</b>) The percentage of degranulated mM-MCs in the nasal mucosa after restraint stress and nasal application of antalarmin. Both acute and chronic restraint stress significantly increased mM-MC numbers in the nasal mucosa and stimulated their degranulation. The increased number of mM-MCs and their degranulation in stressed mouse nasal mucosa was significantly inhibited by nasal application of antalarmin. Control group, <span class="html-italic">n</span> = 4 (without stress); stress group, <span class="html-italic">n</span> = 4; stress + antalarmin (5 μg/g/day) group, <span class="html-italic">n</span> = 5; and antalarmin group, <span class="html-italic">n</span> = 4 (without stress). Error bars indicate SEM. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001. MC, mast cell; mM-MCs, mice nasal mucosa MCs.</p>
Full article ">
43 pages, 8002 KiB  
Article
Design, Synthesis, and Evaluation of Novel 3-Carboranyl-1,8-Naphthalimide Derivatives as Potential Anticancer Agents
by Sebastian Rykowski, Dorota Gurda-Woźna, Marta Orlicka-Płocka, Agnieszka Fedoruk-Wyszomirska, Małgorzata Giel-Pietraszuk, Eliza Wyszko, Aleksandra Kowalczyk, Paweł Stączek, Andrzej Bak, Agnieszka Kiliszek, Wojciech Rypniewski and Agnieszka B. Olejniczak
Int. J. Mol. Sci. 2021, 22(5), 2772; https://doi.org/10.3390/ijms22052772 - 9 Mar 2021
Cited by 19 | Viewed by 4144
Abstract
We synthesized a series of novel 3-carboranyl-1,8-naphthalimide derivatives, mitonafide and pinafide analogs, using click chemistry, reductive amination and amidation reactions and investigated their in vitro effects on cytotoxicity, cell death, cell cycle, and the production of reactive oxygen species in a HepG2 cancer [...] Read more.
We synthesized a series of novel 3-carboranyl-1,8-naphthalimide derivatives, mitonafide and pinafide analogs, using click chemistry, reductive amination and amidation reactions and investigated their in vitro effects on cytotoxicity, cell death, cell cycle, and the production of reactive oxygen species in a HepG2 cancer cell line. The analyses showed that modified naphthalic anhydrides and naphthalimides bearing ortho- or meta-carboranes exhibited diversified activity. Naphthalimides were more cytotoxic than naphthalic anhydrides, with the highest IC50 value determined for compound 9 (3.10 µM). These compounds were capable of inducing cell cycle arrest at G0/G1 or G2M phase and promoting apoptosis, autophagy or ferroptosis. The most promising conjugate 35 caused strong apoptosis and induced ROS production, which was proven by the increased level of 2′-deoxy-8-oxoguanosine in DNA. The tested conjugates were found to be weak topoisomerase II inhibitors and classical DNA intercalators. Compounds 33, 34, and 36 fluorescently stained lysosomes in HepG2 cells. Additionally, we performed a similarity-based assessment of the property profile of the conjugates using the principal component analysis. The creation of an inhibitory profile and descriptor-based plane allowed forming a structure–activity landscape. Finally, a ligand-based comparative molecular field analysis was carried out to specify the (un)favorable structural modifications (pharmacophoric pattern) that are potentially important for the quantitative structure–activity relationship modeling of the carborane–naphthalimide conjugates. Full article
(This article belongs to the Section Biochemistry)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>General structure of icosahedral dicarba-<span class="html-italic">closo</span>-dodecaborane (<span class="html-italic">closo</span>-carborane, C<sub>2</sub>B<sub>10</sub>H<sub>12</sub>), and example structures of naphthalimides with boron cluster [<a href="#B20-ijms-22-02772" class="html-bibr">20</a>].</p>
Full article ">Figure 2
<p>Crystallographic structures of carborane–naphthalimide conjugates: crystal packing and molecular structure observed for <b>39</b> (<b>A</b>,<b>B</b>) and <b>41</b> (<b>C</b>,<b>D</b>).</p>
Full article ">Figure 3
<p>Intermolecular interactions seen in the crystal structure of <b>39</b>. Hydrogen bonds are indicated by dotted lines, and hydrogen donor–acceptor distances are shown in Å. The three molecules of carborane–naphthalimide conjugate are distinguished by distinct colors of carbon atoms in different shades of gray.</p>
Full article ">Figure 4
<p>Effect of compounds <b>6</b> (115 µM), <b>7</b> (104 µM), <b>8</b> (4 µM), <b>9</b> (3 µM), <b>10</b> (8 µM), <b>11</b> (5 µM), <b>15</b> (68 µM), <b>16</b> (61 µM), <b>17</b> (10 µM), <b>18</b> (15 µM), <b>19</b> (14 µM), <b>20</b> (12 µM), <b>31</b>, (53 µM), <b>32</b> (42 µM), <b>33</b> (5 µM), <b>34</b> (8 µM), <b>35</b> (9 µM), <b>36</b> (6 µM), <b>39</b> (11 µM), <b>40</b> (13 µM), <b>41</b> (10 µM), and <b>42</b> (6 µM) on cell cycle distribution in HepG2 cells. The cells were treated with these compounds at a concentration corresponding to the previously estimated IC<sub>50</sub> value. The graph presents the percentage of cells in the G0/G1, S, and G2M phases, respectively. Data are presented as mean ± SD of three independent experiments. Statistical significance is indicated by asterisks: (ns) <span class="html-italic">p</span> &gt; 0.05, (*) <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>ROS production in HepG2 cells after 24 h of incubation with compounds <b>6</b> (57.5 µM), <b>7</b> (52 µM), <b>8</b> (2 µM), <b>9</b> (1.5 µM), <b>10</b> (4 µM), <b>11</b> (2.5 µM), <b>15</b> (34 µM), <b>16</b> (30.7 µM), <b>17</b> (4.9 µM), <b>18</b> (7.5 µM), <b>19</b> (7.2 µM), <b>20</b> (6 µM), <b>32</b> (20.4 µM), <b>39</b> (5.3 µM), <b>40</b> (6.3 µM), <b>41</b> (5.2 µM), and <b>42</b> (3 µM). The concentration chosen for each compound corresponded to half of its IC<sub>50</sub> value. Intracellular ROS production was measured by dual staining with H<sub>2</sub>DCFDA/PI. The intensity of DCF fluorescence corresponded to the intracellular level of ROS in HepG2 cells. Mean fluorescence intensity was measured by flow cytometry. Data are presented as mean ± SD of three independent experiments. Statistical significance is indicated by asterisks: (ns) <span class="html-italic">p</span> &gt; 0.05, (*) <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 6
<p>Effect of compound <b>6</b> (115 µM), <b>7</b> (104 µM), <b>8</b> (4 µM), <b>9</b> (3 µM), <b>10</b> (8 µM), <b>11</b> (5 µM), <b>15</b> (68 µM), <b>16</b> (61 µM), <b>17</b> (10 µM), <b>18</b> (15 µM), <b>19</b> (14 µM), <b>20</b> (12 µM), <b>31</b> (53 µM) <b>32</b> (42 µM), <b>33</b> (5 µM), <b>34</b> (8.5 µM), <b>35</b> (9 µM), <b>36</b> (6 µM), <b>39</b> (11 µM), <b>40</b> (13 µM), <b>41</b> (10 µM), and <b>42</b> (6 µM) on cell death in HepG2 cells. The cells were treated with these compounds at a concentration corresponding to the whole IC<sub>50</sub> value. Quantitative flow cytometry analysis was performed to evaluate apoptosis (<b>A</b>,<b>B</b>), autophagy (<b>C</b>), and ferroptosis (<b>D</b>) induced by compounds after 24 h of treatment. Data are presented as mean ± SD of three independent experiments. Statistical significance is indicated by asterisks: (ns) <span class="html-italic">p</span> &gt; 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 7
<p>Confocal microscopy analysis of the co-localization of compounds <b>33</b>, <b>34</b>, and <b>36</b>. The analysis was carried out after 1 and 24 h of treatment of the cells with compounds <b>33</b> (<b>B</b>), <b>34</b> (<b>C</b>) and <b>36</b> (<b>D</b>) at the final concentration corresponding to the whole IC<sub>50</sub> values. Untreated cells were used as a control (<b>A</b>). Panels with green fluorescence show the autofluorescence of the investigated compounds (Ex/Em 488/500–600 nm), panels with red fluorescence show the autofluorescence of stained lysosomes (Ex/Em 561/585–655 nm), and panels with blue fluorescence present nuclei labeling (Ex/Em 405/430–480 nm). Merged images are shown in the right panels.</p>
Full article ">Figure 8
<p>Inhibition of the relaxation activity of human topoisomerase IIα in the presence of compounds <b>6</b> (<b>A</b>) and <b>7</b> (<b>B</b>) at concentrations of 10, 25, 50, 100, and 200 µM. SC—supercoiled DNA.</p>
Full article ">Figure 9
<p>A 3D scoreplot of carborane-based conjugates.</p>
Full article ">Figure 10
<p>Projection of carborane-containing derivatives on the plane defined by the first vs. second principal component for Dragon descriptors with IC<sub>50</sub> values in the logarithmic scale. Colors code the numerical values of pIC<sub>50</sub>.</p>
Full article ">Figure 11
<p>Projection of carborane-based derivatives on the plane defined by the first vs. second principal component for Dragon descriptors with violations of Ro5 (<b>A</b>) and calculated MW (<b>B</b>). Colors code the numerical values of Ro5 violations and MW.</p>
Full article ">Figure 12
<p>Distribution of T<sub>c</sub> coefficients (<b>A</b>) and triangular matrix of Tanimoto coefficients (<b>B</b>) for carborane conjugates.</p>
Full article ">Figure 13
<p>Grayscaled SALI plot with compounds ordered by increasing pIC<sub>50</sub> values (<b>A</b>) and neighboring plot for carborane-based molecules (<b>B</b>).</p>
Full article ">Figure 14
<p>The favorable (green) and unfavorable (yellow) steric contributions with the most potent molecule <b>9</b> as a reference (<b>A</b>) and pIC<sub>50</sub> actual vs. predicted plot for the training and test sets (<b>B</b>).</p>
Full article ">Scheme 1
<p>Synthesis of naphthalimide–<span class="html-italic">ortho</span>-/<span class="html-italic">meta</span>-carborane conjugates <b>8</b>–<b>11</b>. <span class="html-italic">Reagents and Conditions</span>: (i) HCCSi(CH<sub>3</sub>)<sub>3</sub>, Pd(PPh<sub>3</sub>)<sub>4</sub>, CuI, DMF, TEA, 2 h, 65 °C; (ii) TFA, H<sub>2</sub>O, THF, 8 h, room temperature (RT); (iii) 1-(3-azidopropyl)-1,2-dicarba-<span class="html-italic">closo</span>-dodecaborane (<b>4</b>), CuSO<sub>4</sub>⋅5H<sub>2</sub>O, sodium ascorbate, THF/H<sub>2</sub>O, 4 h, 35 °C (for <b>6</b>); 1-(3-azidopropyl)-1,7-dicarba-<span class="html-italic">closo</span>-dodecaborane (<b>5</b>), CuSO<sub>4</sub>⋅5H<sub>2</sub>O, sodium ascorbate, and THF/H<sub>2</sub>O, 2 h, 35 °C (for <b>7</b>); (iv) <span class="html-italic">N</span>,<span class="html-italic">N</span>-dimethylethylenediamine (for <b>8</b> and <b>10</b>), <span class="html-italic">N</span>-(2-aminoethyl)pyrrolidine (for <b>9</b> and <b>11</b>), EtOH, 1 h, 35 °C and then 1 h, 45 °C. The yields of the compounds are given in parentheses.</p>
Full article ">Scheme 2
<p>Synthesis of naphthalimide–<span class="html-italic">ortho-</span>/<span class="html-italic">meta</span>-carborane conjugates <b>17</b>–<b>20</b>: <span class="html-italic">Reagents and Conditions</span>: (i) NaNO<sub>2</sub>, H<sub>2</sub>SO<sub>4</sub>, H<sub>2</sub>O, 0 °C, and then CO(NH<sub>2</sub>)<sub>2</sub>, 1 h, RT, and 1 h, 100 °C; (ii) HCCCH<sub>2</sub>OH, PPh<sub>3</sub>, THF, 0 °C, and then DIAD, 72 h, RT; (iii) 1-(3-azidopropyl)-1,2-dicarba-<span class="html-italic">closo</span>-dodecaborane (<b>4</b>), CuSO<sub>4</sub>⋅5H<sub>2</sub>O, sodium ascorbate, THF/H<sub>2</sub>O, 3 h, 35 °C (for <b>15</b>); 1-(3-azidopropyl)-1,7-dicarba- <span class="html-italic">closo</span>-dodecaborane (<b>5</b>), CuSO<sub>4</sub>⋅5H<sub>2</sub>O, sodium ascorbate, THF/H<sub>2</sub>O, 4 h, 35 °C (for <b>16</b>); (iv) <span class="html-italic">N</span>,<span class="html-italic">N</span>-dimethylethylenediamine, EtOH, 1 h, 35 °C, and then 1 h, 45 °C (for <b>17</b> and <b>19</b>); <span class="html-italic">N</span>-(2-amino- ethyl)pyrrolidine, EtOH, 1 h, 35 °C, and then 1–3 h, 45 °C (for <b>18</b> and <b>20</b>). The yields of the compounds are given in parentheses.</p>
Full article ">Scheme 3
<p>Synthesis of conjugates <b>31</b>–<b>36</b> via reductive amination: <span class="html-italic">Reagents and Conditions</span>: (i) <span class="html-italic">N</span>,<span class="html-italic">N</span>-dimethylethylenediamine, EtOH, 1 h, 35 °C, and then 1 h, 45 °C (for <b>21</b>); <span class="html-italic">N</span>-(2-amino- ethyl)pyrrolidine, EtOH, 1 h, 35 °C, and then 1 h, 45 °C (for <b>22</b>); (ii) 2-(1,2-dicarba-<span class="html-italic">closo</span>-dodeca- boran-1-yl)ethanal (<b>23</b>), THF, 24 h, reflux (for <b>25</b>, <b>27</b>, and <b>28</b>); 2-(1,7-dicarba-<span class="html-italic">closo</span>-dodecaboran-1-yl)ethanal (<b>24</b>), THF, 24 h, reflux (for <b>26</b>); 2-(1,7-dicarba-<span class="html-italic">closo</span>-dodecaboran-1-yl)ethanal (<b>24</b>), MeOH, 24 h, reflux (for <b>29</b>, <b>30</b>); (iii) NaBH<sub>3</sub>CN, 24 h, RT, and then HCl, 1 h, RT. The yields of the compounds are given in parentheses.</p>
Full article ">Scheme 4
<p>Modification of naphthalimide derivatives with <span class="html-italic">ortho</span>-/<span class="html-italic">meta</span>-carborane via the formation of amide bonds: <span class="html-italic">Reagents and Conditions</span>: (i) 3-(1,2-dicarba-<span class="html-italic">closo</span>-dodecaboran-1-yl)propionic acid (<b>37</b>) or 3-(1,7-dicarba-<span class="html-italic">closo</span>-dodecaboran-1-yl)propionic acid (<b>38</b>), PyBOP, CH<sub>2</sub>Cl<sub>2</sub>, TEA, 4–6 h, RT. The yields of the compounds are given in parentheses.</p>
Full article ">
16 pages, 3351 KiB  
Article
Combined Application of Pan-AKT Inhibitor MK-2206 and BCL-2 Antagonist Venetoclax in B-Cell Precursor Acute Lymphoblastic Leukemia
by Anna Richter, Elisabeth Fischer, Clemens Holz, Julia Schulze, Sandra Lange, Anett Sekora, Gudrun Knuebel, Larissa Henze, Catrin Roolf, Hugo Murua Escobar and Christian Junghanss
Int. J. Mol. Sci. 2021, 22(5), 2771; https://doi.org/10.3390/ijms22052771 - 9 Mar 2021
Cited by 9 | Viewed by 3535
Abstract
Aberrant PI3K/AKT signaling is a hallmark of acute B-lymphoblastic leukemia (B-ALL) resulting in increased tumor cell proliferation and apoptosis deficiency. While previous AKT inhibitors struggled with selectivity, MK-2206 promises meticulous pan-AKT targeting with proven anti-tumor activity. We herein, characterize the effect of MK-2206 [...] Read more.
Aberrant PI3K/AKT signaling is a hallmark of acute B-lymphoblastic leukemia (B-ALL) resulting in increased tumor cell proliferation and apoptosis deficiency. While previous AKT inhibitors struggled with selectivity, MK-2206 promises meticulous pan-AKT targeting with proven anti-tumor activity. We herein, characterize the effect of MK-2206 on B-ALL cell lines and primary samples and investigate potential synergistic effects with BCL-2 inhibitor venetoclax to overcome limitations in apoptosis induction. MK-2206 incubation reduced AKT phosphorylation and influenced downstream signaling activity. Interestingly, after MK-2206 mono application tumor cell proliferation and metabolic activity were diminished significantly independently of basal AKT phosphorylation. Morphological changes but no induction of apoptosis was detected in the observed cell lines. In contrast, primary samples cultivated in a protective microenvironment showed a decrease in vital cells. Combined MK-2206 and venetoclax incubation resulted in partially synergistic anti-proliferative effects independently of application sequence in SEM and RS4;11 cell lines. Venetoclax-mediated apoptosis was not intensified by addition of MK-2206. Functional assessment of BCL-2 inhibition via Bax translocation assay revealed slightly increased pro-apoptotic signaling after combined MK-2206 and venetoclax incubation. In summary, we demonstrate that the pan-AKT inhibitor MK-2206 potently blocks B-ALL cell proliferation and for the first time characterize the synergistic effect of combined MK-2206 and venetoclax treatment in B-ALL. Full article
(This article belongs to the Special Issue Novel Agents and Mechanisms in Acute Leukemias)
Show Figures

Figure 1

Figure 1
<p>Expression of phosphorylated and total AKT protein determined by immunoblot and subsequent quantification. (<b>a</b>) SEM, RS4;11, REH and NALM-6 cells were incubated with the stated concentrations of MK-2206 for the indicated time periods. The displayed immunoblots are representative captions of at least two individual biological replicates. GAPDH was used as internal loading control. Blots were processed and cropped using Image Studio Lite 5.2 software and MS PowerPoint (2011) to improve clarity and conciseness. (<b>b</b>) Phospho (p)-AKT and total AKT band intensities of two independent biological replicates including the blots shown in <a href="#ijms-22-02771-f001" class="html-fig">Figure 1</a>a were determined using Image Studio Lite 5.2 software. AKT phosphorylation was calculated as the ratio of pAKT/total AKT and normalized to the DMSO control separately for every time point. One REH replicate was not quantified due to the very low basal pAKT expression and therefore low signal to noise ratio not acceptable for reliable quantification.</p>
Full article ">Figure 2
<p>Effects of MK-2206 incubation on cell proliferation and metabolic activity. Dose-response curves of SEM, RS4;11, REH and NALM-6 cells incubated with increasing concentrations of MK-2206 for 48 h (<b>a</b>,<b>c</b>) and 72 h (<b>b</b>,<b>d</b>). Proliferation (<b>a</b>,<b>b</b>) was assessed by trypan blue staining and metabolic activity (<b>c</b>,<b>d</b>) was determined by WST-1 assay. Mean ± standard deviation of at least three individual biological replicates. Significance was determined by student’s t test after testing Gaussian distribution. To increase clarity, significance levels are not indicated within the graphs and instead listed in <a href="#app1-ijms-22-02771" class="html-app">Supplementary Tables S1–S4</a>.</p>
Full article ">Figure 3
<p>Effects of MK-2206 on cell morphology and apoptosis induction. (<b>a</b>) Cells were incubated with increasing concentrations for 72 h. Cytospins were prepared and Pappenheim stained. Representative images of all cell lines treated with 0.5 µM MK-2206, 100-fold magnification. Images were acquired using the EVOS xl core microscope (Nikon). Red arrows: disintegrating cell membrane; light blue arrows: vacuolization; purple arrows: nuclear fragmentation; green arrows: cytoplasmic blebs; dark blue arrows: karyopyknosis. (<b>b</b>) SEM and RS4;11 cells were incubated with 0.25 µM MK-2206 for 72 h and subsequently stained with annexin V-FITC and PI for apoptosis analysis by flow cytometry. Cells positive for annexin V and negative for PI are indicated as early apoptotic whereas cells positive for both markers are considered late apoptotic/necrotic. (<b>c</b>) Exemplary dot plots of flow cytometric apoptosis determination in SEM cells.</p>
Full article ">Figure 4
<p>Effects of MK-2206 on primary B-ALL blasts. (<b>a,b</b>) Protein expression of phosphorylated and total AKT was measured by immunofluorescence staining (<b>a</b>) and immunoblotting (<b>b</b>). Cells of patients #0122 and #0159 were orthotopically xenografted into NSG mice as described in [<a href="#B33-ijms-22-02771" class="html-bibr">33</a>,<a href="#B34-ijms-22-02771" class="html-bibr">34</a>] and spleen cells were subsequently harvested for further analysis. (<b>a</b>) Expression of pAKT was determined by immunofluorescence using the LSM780 confocal microscope (Zeiss) at 20-fold magnification (left panel). The right hand images are enlargements of the regions indicated in the left hand red boxes. (<b>b</b>) Cells were lysed and analyzed for phosphorylated and total AKT protein expression by immunoblot. GAPDH was used as internal loading control. Blots were processed and cropped using Image Studio Lite 5.2 software and MS PowerPoint (2011) to improve clarity and conciseness. (<b>c</b>) Genetic variants in both patients were detected using Cancer hotspot panel (Ion PGM System, Thermo Fisher Scientific) and previously described in [<a href="#B34-ijms-22-02771" class="html-bibr">34</a>]. (<b>d,e</b>) Xenograft-derived primary blasts were cultured on a murine stromal cell feeder layer system and cell viability and apoptosis induction were assessed after 72 h incubation with increasing concentrations of MK-2206. (<b>d</b>) Cell viability was determined by CellTiter Blue assay in technical triplicates. Mean ± standard deviation of three individual biological replicates. Significance was determined by student’s t test after testing Gaussian distribution. * <span class="html-italic">P</span> &lt; 0.05. (<b>e</b>) Cells were stained with anti-human CD19, annexin V-FITC and PI for apoptosis analysis by flow cytometry. Cells positive for annexin V and negative for PI are indicated as early apoptotic whereas cells positive for both markers are considered late apoptotic/necrotic. Human primary blasts and murine feeder cells were discriminated by anti-human CD19 staining.</p>
Full article ">Figure 5
<p>Combined application of MK-2206 (MK) and venetoclax (ven). (<b>a</b>,<b>b</b>) Cytotoxicity of MK-2206 (0.25 µM), venetoclax (10 nM) or the combination was determined by hemolysis assay and PBMC viability analysis. Blood of five healthy voluntary donors was used. Technical triplicates were performed for both assays. Mean ± standard deviation. (<b>a</b>) Hemolytic activity was assessed by hemoglobin release after 120 min incubation with the substances, 1% SDS (positive control) or PBS (negative control). (<b>b</b>) PBMCs were isolated by density gradient centrifugation and incubated with the respective inhibitors or 10 µM DMSO (control) for 24 h. Cell viability was determined by Calcein AM assay. (<b>c,d</b>) Assessment of cell proliferation and metabolic activity was performed by trypan blue staining and WST-1 assay, respectively and compared to DMSO-treated control cells. Cells were incubated with 0.25 µM MK-2206, 10 nM (SEM (<b>c</b>)) or 2.5 nM (RS4;11 (<b>d</b>)) venetoclax or both inhibitors for 72 h. (<b>e</b>) Effects of sequential application of MK-2206 and venetoclax on SEM and RS4;11 cells were determined by WST-1 assay. Substances were applied either simultaneously (MK + ven) or the second substance was added 24 h after the first (MK &gt; ven or ven &gt; MK). Metabolic activity was assessed after 72 h. Mean ± standard deviation of at least three individual biological replicates. Significance was determined by one-way ANOVA. * <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. Asterisks demonstrate significance compared to DMSO-treated control cells or other treatments when connected by lines. (<b>f</b>) Synergy was calculated according to the Bliss independence model with values &gt; 0 indicating synergistic combinations [<a href="#B35-ijms-22-02771" class="html-bibr">35</a>]. Bliss values were calculated based on the mean biological response of at least four biological replicates. Cells were incubated with MK-2206, venetoclax or both (simultaneously or sequentially) and metabolic activity was assessed by WST-1 assay after 72 h.</p>
Full article ">Figure 6
<p>Analysis of apoptotic processes after single or combined MK-2206 (MK; 0.25 µM) and venetoclax (ven; 10 nM (SEM) or 2.5 nM (RS4;11)) incubation in SEM and RS4;11. (<b>a</b>) Cells were incubated for 72 h and stained with annexin V-FITC and PI. The amount of apoptotic cells was determined by flow cytometry with cells single positive for annexin V called early aposptotic and cells positive for both markers late apoptotic/necrotic. Mean ± standard deviation of at least three individual biological replicates. Significance was determined by one-way ANOVA. *** <span class="html-italic">P</span> &lt; 0.001. (<b>b,c</b>) Downstream apoptotic signaling was investigated by Bax translocation assay. Cells were treated with one or both inhibitors for 48 h and spun onto microscopic slides. Mitochondria were stained using MitoSpy™ red and Bax protein was stained in green. Images were taken using the Eclipse TE200 microscope (Nikon) and at 40-fold magnification. Four biological replicates were performed for each treatment and two to four individual images per biological replicate were considered for colocalization analysis. (<b>b</b>) Color-merged images were split into separate images for both channels and regions of interest were marked using Fiji program. Colocalization analysis of Bax and mitochondria was performed using the Coloc2 plugin in Fiji and Pearson’s coefficient were used to quantify the amount of Bax/mitochondria overlap. (<b>c</b>) Representative images of SEM and RS4;11 cells treated with DMSO (control), MK-2206, venetoclax or both. Mitochondria are stained in red and Bax in green. Yellow signals are a product of red and green signal overlay and indicate colocalization of both structures.</p>
Full article ">
20 pages, 2920 KiB  
Article
Preclinical Evaluation of 99mTc-ZHER2:41071, a Second-Generation Affibody-Based HER2-Visualizing Imaging Probe with a Low Renal Uptake
by Maryam Oroujeni, Sara S. Rinne, Anzhelika Vorobyeva, Annika Loftenius, Joachim Feldwisch, Per Jonasson, Vladimir Chernov, Anna Orlova, Fredrik Y. Frejd and Vladimir Tolmachev
Int. J. Mol. Sci. 2021, 22(5), 2770; https://doi.org/10.3390/ijms22052770 - 9 Mar 2021
Cited by 20 | Viewed by 4192
Abstract
Radionuclide imaging of HER2 expression in tumours may enable stratification of patients with breast, ovarian, and gastroesophageal cancers for HER2-targeting therapies. A first-generation HER2-binding affibody molecule [99mTc]Tc-ZHER2:V2 demonstrated favorable imaging properties in preclinical studies. Thereafter, the affibody scaffold has been extensively [...] Read more.
Radionuclide imaging of HER2 expression in tumours may enable stratification of patients with breast, ovarian, and gastroesophageal cancers for HER2-targeting therapies. A first-generation HER2-binding affibody molecule [99mTc]Tc-ZHER2:V2 demonstrated favorable imaging properties in preclinical studies. Thereafter, the affibody scaffold has been extensively modified, which increased its melting point, improved storage stability, and increased hydrophilicity of the surface. In this study, a second-generation affibody molecule (designated ZHER2:41071) with a new improved scaffold has been prepared and characterized. HER2-binding, biodistribution, and tumour-targeting properties of [99mTc]Tc-labelled ZHER2:41071 were investigated. These properties were compared with properties of the first-generation affibody molecules, [99mTc]Tc-ZHER2:V2 and [99mTc]Tc-ZHER2:2395. [99mTc]Tc-ZHER2:41071 bound specifically to HER2 expressing cells with an affinity of 58 ± 2 pM. The renal uptake for [99mTc]Tc-ZHER2:41071 and [99mTc]Tc-ZHER2:V2 was 25–30 fold lower when compared with [99mTc]Tc-ZHER2:2395. The uptake in tumour and kidney for [99mTc]Tc-ZHER2:41071 and [99mTc]Tc-ZHER2:V2 in SKOV-3 xenografts was similar. In conclusion, an extensive re-engineering of the scaffold did not compromise imaging properties of the affibody molecule labelled with 99mTc using a GGGC chelator. The new probe, [99mTc]Tc-ZHER2:41071 provided the best tumour-to-blood ratio compared to HER2-imaging probes for single photon emission computed tomography (SPECT) described in the literature so far. [99mTc]Tc-ZHER2:41071 is a promising candidate for further clinical translation studies. Full article
(This article belongs to the Special Issue Cancer Molecular Imaging 2.0)
Show Figures

Figure 1

Figure 1
<p>Overview of structural differences, features, and chelator positioning of ZHER2:41071, ZHER2:V2, and ZHER2:2395. All three constructs share the amino acid structure in the binding site, highlighted by the patch with a green pattern. Shared sequences are in a light blue structure, positions with scaffold variation are highlighted in black. The one close to the C-terminal chelator sequence GGGC (marked with an arrow) is consistent of an –SES- amino acids sequence, which replaced an -NDA-sequence to further improve peptide stability and increase hydrophilicity of the affibody molecule. The -SES- motif could have challenged efficient labeling by competing with the GGGC-chelator. Placement of a KVDC chelating sequence within ZHER2:2395 as compared to HER2:V2 and HER2:41071 is shown in gray.</p>
Full article ">Figure 2
<p>Circular dichroism measurements of (<b>A</b>) ZHER2:41071 and (<b>B</b>) ZHER:V2 before (black line) and after (dotted line) variable temperature measurements (VTM). The curves indicate alpha-helical conformation and reversibility after heat treatment.</p>
Full article ">Figure 3
<p>SDS–PAGE analysis of <sup>99m</sup>Tc-labelled affibody molecules. (1) [<sup>99m</sup>Tc]Tc-ZHER2:41071, (2) [<sup>99m</sup>Tc]Tc-ZHER2:2395, (3) [<sup>99m</sup>Tc]Tc-ZHER2:V2 and (4) <sup>99m</sup>TcO<sub>4</sub><sup>−</sup>. The signal was measured as digital light units (DLU) and is proportional to radioactivity at a given point of a lane in the SDS-PAGE gel. Note that the applied activity of <sup>99m</sup>TcO<sub>4</sub><sup>−</sup> was lower compared to applied activity of affibody molecules.</p>
Full article ">Figure 4
<p>In vitro binding specificity of (<b>A</b>,<b>D</b>) [<sup>99m</sup>Tc]Tc-ZHER2:41071, (<b>B</b>,<b>E</b>) [<sup>99m</sup>Tc]Tc-ZHER2:V2 and (<b>C</b>,<b>F</b>) [<sup>99m</sup>Tc]Tc-ZHER2:2395 to HER2-expressing SKOV3 (<b>A</b>–<b>C</b>) and BT-474 (<b>D</b>–<b>F</b>) cell-lines. For the pre-saturation of HER2, a 500-fold molar excess of a non-radioactive affibody molecule was added. The data are presented as an average value from three samples ± SD.</p>
Full article ">Figure 5
<p>Interaction Map of (<b>A</b>) [<sup>99m</sup>Tc]Tc-ZHER2:2395, (<b>B</b>) [<sup>99m</sup>Tc]Tc-ZHER2:41071. and (<b>C</b>) [<sup>99m</sup>Tc]Tc-ZHER2:V2 binding to HER2-expressing SKOV3 cells. Input data are obtained from LigandTracer measurement of cell-bound activity during association of labelled compounds to and dissociation from SKOV-3 cells. InteractionMap finds individual 1:1 interactions in the input data whose weighted sum explain the observed binding process. The individual interactions are displayed as coloured spots in a ka/kd plot with their colour representing their weight: warmer colours represent more abundant interactions. Binding was measured at three different concentrations: 0.33, 1, and 3 nM. The measurement was performed in duplicates.</p>
Full article ">Figure 6
<p>Normalized cellular retention of (<b>A</b>,<b>D</b>) [<sup>99m</sup>Tc]Tc-ZHER2:41071, (<b>B</b>,<b>E</b>) [<sup>99m</sup>Tc]Tc-ZHER2:V2 and (<b>C</b>,<b>F</b>) [<sup>99m</sup>Tc]Tc-ZHER2:2395 to HER2-expressing SKOV3 (<b>A</b>–<b>C</b>) and BT-474 (<b>D</b>–<b>F</b>) cell-lines. The data are presented as an average (<span class="html-italic">n</span> = 3) and SD. Error bars are not seen because they are smaller than point symbols.</p>
Full article ">Figure 7
<p>Uptake of <sup>99m</sup>Tc-labelled affibody molecules in different organs of female NMRI mice at 4 h after injection. Then, 1 µg of labelled affibody molecules (60 kBq, 100 µL in PBS) was injected into the tail vein.</p>
Full article ">Figure 8
<p>In vivo specificity of [<sup>99m</sup>Tc]Tc-ZHER2:41071 in HER2-negative Ramos xenografts and HER2-positive SKOV3 xenografts at 4 h after injection. The data are presented as the average (<span class="html-italic">n</span> = 4) and SD. * <span class="html-italic">p</span> &lt; 0.0005.</p>
Full article ">Figure 9
<p>(<b>A</b>) Biodistribution and (<b>B</b>) tumor-to-organ ratios of selected conjugates, [<sup>99m</sup>Tc]Tc-ZHER2:41071 and [<sup>99m</sup>Tc]Tc-ZHER2:V2, in BALB/C nu/nu mice bearing HER2-expressing SKOV3 xenografts 4 h after injection. The data are presented as the average value (<span class="html-italic">n</span> = 4) and SD.</p>
Full article ">Figure 10
<p>Imaging of HER2-negative Ramos xenograft (left mouse) and HER2-positive SKOV3 xenograft (right mouse) in BALB/C nu/nu mice using [<sup>99m</sup>Tc]Tc- ZHER2:41071. To confirm binding is HER2-specific, one mouse bearing Ramos (HER2-) was used for imaging.</p>
Full article ">
48 pages, 22172 KiB  
Review
A Review of the Pharmacological Activities and Recent Synthetic Advances of γ-Butyrolactones
by Joonseong Hur, Jaebong Jang and Jaehoon Sim
Int. J. Mol. Sci. 2021, 22(5), 2769; https://doi.org/10.3390/ijms22052769 - 9 Mar 2021
Cited by 50 | Viewed by 9018
Abstract
γ-Butyrolactone, a five-membered lactone moiety, is one of the privileged structures of diverse natural products and biologically active small molecules. Because of their broad spectrum of biological and pharmacological activities, synthetic methods for γ-butyrolactones have received significant attention from synthetic and [...] Read more.
γ-Butyrolactone, a five-membered lactone moiety, is one of the privileged structures of diverse natural products and biologically active small molecules. Because of their broad spectrum of biological and pharmacological activities, synthetic methods for γ-butyrolactones have received significant attention from synthetic and medicinal chemists for decades. Recently, new developments and improvements in traditional methods have been reported by considering synthetic efficiency, feasibility, and green chemistry. In this review, the pharmacological activities of natural and synthetic γ-butyrolactones are described, including their structures and bioassay methods. Mainly, we summarize recent advances, occurring during the past decade, in the construction of γ-butyrolactone classified based on the bond formation in γ-butyrolactone between (i) C5-O1 bond, (ii) C4-C5 and C2-O1 bonds, (iii) C3-C4 and C2-O1 bonds, (iv) C3-C4 and C5-O1 bonds, (v) C2-C3 and C2-O1 bonds, (vi) C3-C4 bond, and (vii) C2-O1 bond. In addition, the application to the total synthesis of natural products bearing γ-butyrolactone scaffolds is described. Full article
Show Figures

Figure 1

Figure 1
<p>Bond disconnections for the synthesis of <span class="html-italic">γ</span>-butyrolactones.</p>
Full article ">Figure 2
<p>TfOH-catalyzed oxidative lactonization with peroxyacid.</p>
Full article ">Figure 3
<p>TfOH-catalyzed oxidative lactonization with sodium periodate.</p>
Full article ">Figure 4
<p>Trifluoroacetophenone-catalyzed oxidative lactonization with hydrogen peroxide.</p>
Full article ">Figure 5
<p>Oxidative ring contraction of 3,4-dihydropyran-2-ones.</p>
Full article ">Figure 6
<p>Aminolactonization of <span class="html-italic">t</span>-butyl pentenoate with iminoiodane (<b>top</b>) and the application of the resulting <span class="html-italic">γ</span>-butyrolactone (<b>bottom</b>).</p>
Full article ">Figure 7
<p>Bromolactonization of pentenoic acid with KBr and Oxone.</p>
Full article ">Figure 8
<p>Bromolactonization of pentenoic acid with isoselenazolone.</p>
Full article ">Figure 9
<p>Acid-promoted cyclopropane opening/intramolecular ester trapping.</p>
Full article ">Figure 10
<p>Silver-mediated cyclopropane opening/intramolecular acid trapping.</p>
Full article ">Figure 11
<p>Gold-catalyzed intramolecular allylic alkylation of allylic acetate.</p>
Full article ">Figure 12
<p>Gold-NHC complex catalyzed intramolecular allylic alkylation of allylic alcohol.</p>
Full article ">Figure 13
<p>Gold-catalyzed dehydrative lactonization.</p>
Full article ">Figure 14
<p>Gold-catalyzed lactonization of allene system.</p>
Full article ">Figure 15
<p>Photoredox-catalyzed γ-butyrolactone synthesis.</p>
Full article ">Figure 16
<p>Asymmetric synthesis of <span class="html-italic">α</span>-exo-methylene-<span class="html-italic">γ</span>-butyrolactone via iridium-catalyzed 2-(alkoxycarbonyl)allylation.</p>
Full article ">Figure 17
<p>Syntheses of <span class="html-italic">γ</span>-butyrolactones via ruthenium-catalyzed hydrohydroxyalkylation. (<b>a</b>) Syntheses of spiro-<span class="html-italic">γ</span>-butyrolactones from diols and methyl acrylate; (<b>b</b>) Syntheses of polysubstituted 2,3’-spirooxindole-<span class="html-italic">γ</span>-butyrolactones from <span class="html-italic">N</span>-benzyl-3-hydroxyoxindole and acrylic esters; (<b>c</b>) Syntheses of <span class="html-italic">α</span>-exo-methylene-<span class="html-italic">γ</span>-butyrolactones from hydroxyl-substituted methacrylate and diols; (<b>d</b>) Redox level-independent formation of <b>51</b>.</p>
Full article ">Figure 18
<p>Asymmetric synthesis of <span class="html-italic">α</span>-exo-methylene <span class="html-italic">γ</span>-butyrolactone via chromium-catalyzed 2-(alkoxycarbonyl)allylation and lactonization and total synthesis of (+)-methylenolactocin.</p>
Full article ">Figure 19
<p>Asymmetric synthesis of 2,3′-spirooxindole-<span class="html-italic">α</span>-exo-methylene <span class="html-italic">γ</span>-butyrolactone via indium-catalyzed amide allylation and lactonization.</p>
Full article ">Figure 20
<p>Asymmetric syntheses of 2,3′-spirooxindole-<span class="html-italic">γ</span>-butyrolactone via NHC-catalyzed homoenolate annulation. (<b>a</b>,<b>b</b>) NHC-catalyzed 2,3′-spirooxindole-<span class="html-italic">γ</span>-butyrolactone formation from enals; (<b>c</b>,<b>d</b>) NHC-catalyzed 2,3′-spirooxindole-<span class="html-italic">γ</span>-butyrolactone formation from carboxylic acids; (<b>e</b>) NHC-catalyzed 2,3′-spirooxindole-<span class="html-italic">γ</span>-butyrolactone formation from aryl esters.</p>
Full article ">Figure 21
<p>Asymmetric synthesis of 3,4,4-trisubstituted <span class="html-italic">γ</span>-butyrolactones via NHC-catalyzed dynamic kinetic resolution.</p>
Full article ">Figure 22
<p>Synthesis of <span class="html-italic">γ</span>-butyrolactones via the alcohol-selective C-H activation mediated by photoredox catalysis.</p>
Full article ">Figure 23
<p>Synthesis of <span class="html-italic">γ</span>-butyrolactones via photoorganocatalytic C-H activation.</p>
Full article ">Figure 24
<p>Asymmetric synthesis of 4,5,5-trisubstituted-<span class="html-italic">γ</span>-butyrolactones via electroreductive C-C bond coupling.</p>
Full article ">Figure 25
<p>Synthesis of 3,3′-spirooxindole-<span class="html-italic">γ</span>-butyrolactones via peptide coupling reagent-assisted lactonization.</p>
Full article ">Figure 26
<p>Asymmetric synthesis of 4,5-disubstituted-<span class="html-italic">γ</span>-butyrolactones via organocatalyzed three-component coupling.</p>
Full article ">Figure 27
<p>Synthesis of <span class="html-italic">γ</span>-butyrolactones via ruthenium pincer-catalyzed hydrogen autotransfer.</p>
Full article ">Figure 28
<p>Synthesis of <span class="html-italic">γ</span>-butyrolactones via ionic liquid-assisted epoxide opening and lactonization.</p>
Full article ">Figure 29
<p>Polar radical crossover cycloaddition of the oxidizable alkenes and <span class="html-italic">α,β</span>-unsaturated acids.</p>
Full article ">Figure 30
<p>Polar radical crossover cycloaddition of the oxidizable alkenes and <span class="html-italic">O</span>-benzyloxime acids.</p>
Full article ">Figure 31
<p><span class="html-italic">γ</span>-Butyrolactone synthesis via the photoredox-catalyzed atom-transfer radical addition (ATRA).</p>
Full article ">Figure 32
<p>Synthesis of carbohydrate-based <span class="html-italic">γ</span>-butyrolactones through Mn(OAc)<sub>3</sub>-mediated radical lactonization.</p>
Full article ">Figure 33
<p>Synthesis of <span class="html-italic">γ</span>-butyrolactones bearing quaternary carbon centers via copper-catalyzed cyclopropanol ring-opening cross-coupling reaction.</p>
Full article ">Figure 34
<p>Synthesis of bicyclic <span class="html-italic">γ</span>-butyrolactones via palladium-catalyzed carbonylation using iron pentacarbonyl.</p>
Full article ">Figure 35
<p>Synthesis of C3-substituted <span class="html-italic">γ</span>-butyrolactones via palladium-catalyzed carbonylation cascade in the ionic liquid.</p>
Full article ">Figure 36
<p>(<b>a</b>) Synthesis of thiolated <span class="html-italic">α</span>-alkylidene-<span class="html-italic">γ</span>-butyrolactones via cobalt-catalyzed carbonylation; (<b>b</b>) Synthesis of thiolated <span class="html-italic">α</span>-alkylidene-<span class="html-italic">γ</span>-butyrolactones via palladium-catalyzed carbonylation.</p>
Full article ">Figure 37
<p>Synthesis of oxaspiro-<span class="html-italic">γ</span>-butyrolactones via palladium-catalyzed carbonylative spirolactonization and total synthesis of <span class="html-italic">α</span>-levantanolide and <span class="html-italic">α</span>-levantenolide.</p>
Full article ">Figure 38
<p>Synthesis of 3,3,5-trisubstituted-<span class="html-italic">γ</span>-butyrolactones via rhodium-catalyzed Markovnikov hydroformylation and oxidation.</p>
Full article ">Figure 39
<p>Asymmetric synthesis of 4-substituted <span class="html-italic">γ</span>-butyrolactones via rhodium-catalyzed hydroformylation and oxidation.</p>
Full article ">Figure 40
<p>Synthesis of α-alkyledene <span class="html-italic">γ</span>-butyrolactones via Ni(0)-catalyzed carboxylation and total synthesis of (±)-heteroplexisolide E.</p>
Full article ">Figure 41
<p>Synthesis of <span class="html-italic">γ</span>-butyrolactones and natural products via Rh-catalyzed C-H insertion.</p>
Full article ">
16 pages, 4271 KiB  
Article
Rational Design of Adenylate Kinase Thermostability through Coevolution and Sequence Divergence Analysis
by Jian Chang, Chengxin Zhang, Huaqiang Cheng and Yan-Wen Tan
Int. J. Mol. Sci. 2021, 22(5), 2768; https://doi.org/10.3390/ijms22052768 - 9 Mar 2021
Cited by 8 | Viewed by 3088
Abstract
Protein engineering is actively pursued in industrial and laboratory settings for high thermostability. Among the many protein engineering methods, rational design by bioinformatics provides theoretical guidance without time-consuming experimental screenings. However, most rational design methods either rely on protein tertiary structure information or [...] Read more.
Protein engineering is actively pursued in industrial and laboratory settings for high thermostability. Among the many protein engineering methods, rational design by bioinformatics provides theoretical guidance without time-consuming experimental screenings. However, most rational design methods either rely on protein tertiary structure information or have limited accuracies. We proposed a primary-sequence-based algorithm for increasing the heat resistance of a protein while maintaining its functions. Using adenylate kinase (ADK) family as a model system, this method identified a series of amino acid sites closely related to thermostability. Single- and double-point mutants constructed based on this method increase the thermal denaturation temperature of the mesophilic Escherichia coli (E. coli) ADK by 5.5 and 8.3 °C, respectively, while preserving most of the catalytic function at ambient temperatures. Additionally, the constructed mutants have improved enzymatic activity at higher temperature. Full article
(This article belongs to the Special Issue Enzymes as Biocatalysts: Current Research Trends and Applications)
Show Figures

Figure 1

Figure 1
<p>Residue Correlation Analysis (RCA) of Protein Sectors in ADK Family. (<b>A</b>) Heat map representation of RCA matrix <span class="html-italic">r<sub>ij</sub></span> for MSA of ADK. <span class="html-italic">X</span><span class="html-italic">−</span> and <span class="html-italic">y</span>−axis coordinates both correspond to residue indices of ADK from <span class="html-italic">E. coli</span>, which has 214 residues. (<b>B</b>) Three dimensional scatter plot of the 214 residues in the space formed by the three eigenvectors of the second, the third, and the fourth largest eigenvectors. Each data point represented one position. After eliminating randomized background residues (gray), the rest of the positions could be clustered into four sectors, colored green, blue, orange, and magenta. (<b>C</b>) The relative entropy angle θ of thermophilic and mesophilic sequence profile at each position. Bars were colored by sectors. Residues whose angles were larger than 0.5 in the magenta sector were selected for further mutation (marked by arrow heads). (<b>D</b>) Amino acid distribution of thermophilic sequences at the chosen mutation positions. The first series of single-point mutations were S41K, D76V, G100N, I101R, P139K, P140S, K141R, and R206F.</p>
Full article ">Figure 2
<p>Thermal stability and enzymatic activity of single-point mutants characterized by circular dichroism (CD) spectroscopy and room temperature enzymatic activity assay. (<b>A</b>) CD differential signal at 222nm of wild type and mutants S41A, D76N, G100N, I101T, P139K, P140S, K141A, K141P, and R206F from 40 °C to 70 °C. All data were normalized to a scale of zero to one. To be distinguished from each other, each curve was offset by one along the <span class="html-italic">y</span>-axis. (<b>B</b>) Forward enzymatic activity of wild type and eight mutants at 25 °C. [ATP] = 1 mM, [magnesium acetate (MgOAc<sub>2</sub>)] = 2 mM.</p>
Full article ">Figure 3
<p>Correlation between melting temperature and enzymatic activity of wt and selected mutants of ADK. (<b>A</b>) Forward reaction activities (at 0.2 mM AMP and 1 mM ATP) versus melting temperature of ADK wild type and single mutants. (<b>B</b>) Temperature dependence of enzyme activity for ADK from <span class="html-italic">E. coli</span>, single mutant R206F, double mutant R206F/S41A and ADK from <span class="html-italic">Aquifex aeolicus</span> in the direction of ADP formation. The absolute value of 100% maximal activity was calculated as <math display="inline"><semantics> <mrow> <mn>458</mn> <mo>±</mo> <mn>4</mn> <mtext> </mtext> <msup> <mi>s</mi> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </semantics></math>. (<b>C</b>) Four double mutants, R206F/S41A, R206F/G100N, R206F/P139K, R206F/P140S, all displayed a moderate <span class="html-italic">T<sub>m</sub></span> increment lower than the sum of respective <span class="html-italic">T<sub>m</sub></span> rise from two related individual mutants. The synergistic effect lowered as the Δ<span class="html-italic">T<sub>m</sub></span> sum increased. The dash line indicated the theoretical Δ<span class="html-italic">T<sub>m</sub></span> distribution of double mutants compared to R206F provided that the synergistic effect was linear.</p>
Full article ">Figure 4
<p>Mutations sites chosen by the relative entropy angle <span class="html-italic">θ</span>. (<b>A</b>) Residues colored according to our protein sectors color scheme on the “closed” conformation of ADK (PDB id: 1ANK chain A). ATP analog and AMP (both colored black) were shown in sticks. The structure on the right was obtained by rotating the diagram on the left by 90 degrees clockwise around the vertical axis. (<b>B</b>) Residues directly participating in substrate binding in <span class="html-italic">E. coli</span> ADK. All sectors except the magenta one were involved in ADK’s function execution. (<b>C</b>) Eight selected mutation sites were marked as magenta spheres in the tertiary structure of ADK.</p>
Full article ">
14 pages, 2574 KiB  
Article
Comparative Analysis of CREB3 and CREB3L2 Protein Expression in HEK293 Cells
by Kentaro Oh-hashi, Ayumi Yamamoto, Ryoichi Murase and Yoko Hirata
Int. J. Mol. Sci. 2021, 22(5), 2767; https://doi.org/10.3390/ijms22052767 - 9 Mar 2021
Cited by 6 | Viewed by 3373
Abstract
We performed a comparative analysis of two ER-resident CREB3 family proteins, CREB3 and CREB3L2, in HEK293 cells using pharmacological and genome editing approaches and identified several differences between the two. Treatment with brefeldin A (BFA) and monensin induced the cleavage of full-length CREB3 [...] Read more.
We performed a comparative analysis of two ER-resident CREB3 family proteins, CREB3 and CREB3L2, in HEK293 cells using pharmacological and genome editing approaches and identified several differences between the two. Treatment with brefeldin A (BFA) and monensin induced the cleavage of full-length CREB3 and CREB3L2; however, the level of the full-length CREB3 protein, but not CREB3L2 protein, was not noticeably reduced by the monensin treatment. On the other hand, treatment with tunicamycin (Tm) shifted the molecular weight of the full-length CREB3L2 protein downward but abolished CREB3 protein expression. Thapsigargin (Tg) significantly increased the expression of only full-length CREB3L2 protein concomitant with a slight increase in the level of its cleaved form. Treatment with cycloheximide and MG132 revealed that both endogenous CREB3 and CREB3L2 are proteasome substrates. In addition, kifunensine, an α-mannosidase inhibitor, significantly increased the levels of both full-length forms. Consistent with these findings, cells lacking SEL1L, a crucial ER-associated protein degradation (ERAD) component, showed increased expression of both full-length CREB3 and CREB3L2; however, cycloheximide treatment downregulated full-length CREB3L2 protein expression more rapidly in SEL1L-deficient cells than the full-length CREB3 protein. Finally, we investigated the induction of the expression of several CREB3 and CREB3L2 target genes by Tg and BFA treatments and SEL1L deficiency. In conclusion, this study suggests that both endogenous full-length CREB3 and CREB3L2 are substrates for ER-associated protein degradation but are partially regulated by distinct mechanisms, each of which contributes to unique cellular responses that are distinct from canonical ER signals. Full article
(This article belongs to the Special Issue Physio-Pathological Role of ERAD and Its Pharmacological Modulation)
Show Figures

Figure 1

Figure 1
<p>Expression of the CREB3 and CREB3L2 proteins in HEK293 cells. (<b>A</b>–<b>C</b>) HEK293 cells were treated with thapsigargin (Tg, 0.01 μM), tunicamycin (Tm, 2 μg/mL), brefeldin A (BFA, 0.5 μg/mL), monensin (Mone, 1 μM), nigericin (Nigr, 0.2 μM), concanamycin A (CMA, 50 nM), cycloheximide (CHX, 20 μg/mL), MG132 (MG, 10 μM), or vehicle (Control (Con)) for 8 h. The expression of the indicated protein was detected as described in the Materials and Methods. Open and filled arrowheads indicate full-length mature and cleaved CREB3 and CREB3L2 proteins, respectively. Representative results of three independent cultures are shown.</p>
Full article ">Figure 2
<p>SEL1 deficiency increased CREB3 and CREB3L2 protein expression in HEK293 cells. (<b>A</b>) Wild-type (wt) and SEL1L-deficient (SEL1L-KD) HEK293 cells were cultured (<b>A</b>) and treated with BFA (0.5 μg/mL), monensin (1 μM), Tg (0.01 μM), Tm (2 μg/mL), or vehicle for 6 h (<b>B</b>–<b>E</b>). The expression of the indicated protein was detected as described in the Materials and Methods. Open and filled arrowheads indicate full-length mature and cleaved CREB3 and CREB3L2 proteins, respectively. Representative results of three independent cultures are shown. (<b>D</b>,<b>E</b>) The relative amounts of full-length CREB3 and CREB3L2 were evaluated as described in the Materials and Methods. The level of each full-length form in untreated SEL1L-deficient cells (#2) was considered “1”. Each value represents the mean ± SEM from three independent cultures. Values marked with asterisks are significantly different from value of untreated wild-type cells (* <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2 Cont.
<p>SEL1 deficiency increased CREB3 and CREB3L2 protein expression in HEK293 cells. (<b>A</b>) Wild-type (wt) and SEL1L-deficient (SEL1L-KD) HEK293 cells were cultured (<b>A</b>) and treated with BFA (0.5 μg/mL), monensin (1 μM), Tg (0.01 μM), Tm (2 μg/mL), or vehicle for 6 h (<b>B</b>–<b>E</b>). The expression of the indicated protein was detected as described in the Materials and Methods. Open and filled arrowheads indicate full-length mature and cleaved CREB3 and CREB3L2 proteins, respectively. Representative results of three independent cultures are shown. (<b>D</b>,<b>E</b>) The relative amounts of full-length CREB3 and CREB3L2 were evaluated as described in the Materials and Methods. The level of each full-length form in untreated SEL1L-deficient cells (#2) was considered “1”. Each value represents the mean ± SEM from three independent cultures. Values marked with asterisks are significantly different from value of untreated wild-type cells (* <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>Comparison of the stability of the CREB3 and CREB3L2 proteins in HEK293 cells. (<b>A</b>) Wild-type (wt) and SEL1L-deficient HEK293 cells were treated with CHX (20 μg/mL) for the indicated times. The expression of the indicated proteins in wt and SEL1L-deficient cells was detected. Open and filled arrowheads indicate full-length mature and cleaved CREB3 and CREB3L2 proteins, respectively. Representative results of three independent cultures are shown. (<b>B</b>) The relative amounts of full-length CREB3 and CREB3L2 in wt and SEL1L-deficient cells were evaluated as described in the Materials and Methods. The amount of each protein in wt and SEL1L-deficient cells without CHX treatment was considered “1”. Each value represents the mean ± SEM from three independent cultures. Values marked with asterisks are significantly different between the indicated groups (* <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Effects of the MG132 treatment on the BFA- and monensin-induced processing of CREB3 and CREB3L2 proteins in HEK293 cells. HEK293 cells were treated with BFA (0.5 µg/mL), monensin (1 µM), or vehicle, in the presence or absence of MG132 (10 µM), for 8 h. The expression of the indicated protein was detected as described in the Materials and Methods. Open and filled arrowheads indicate full-length mature and cleaved CREB3 and CREB3L2 proteins, respectively. Representative results of three independent cultures are shown.</p>
Full article ">Figure 5
<p>Treatment with kifunensine increased the levels of the CREB3 and CREB3L2 proteins in HEK293 cells. (<b>A</b>) HEK293 cells were treated with kifunensine (Kif, 5 µg/mL), BFA (0.5 µg/mL), or vehicle for 8 h. The expression of the indicated protein was detected as described in the Materials and Methods. Open and filled arrowheads indicate full-length mature and cleaved CREB3 and CREB3L2 proteins, respectively. Representative results of four independent cultures are shown. (<b>B</b>) The relative amounts of full-length CREB3 and CREB3L2 were evaluated as described in the Materials and Methods. The amount of each full-length form in cells treated with kifunensine was considered “1”. Each value represents the mean ± SEM from four independent cultures. Values marked with asterisks are significantly different from value of untreated cells (* <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 6
<p>Effects of SEL1L deficiency on CREB3- and CREB3L2-regulated gene expression in HEK293 cells. (<b>A</b>) Wild-type (wt) and SEL1L-deficient HEK293 cells were treated with Tg (0.01 µM), BFA (0.5 µg/mL), or vehicle for 6 h. The expression of the indicated mRNAs in wt and SEL1L-deficient cells was detected as described in the Materials and Methods. Representative results of four independent cultures are shown. (<b>B</b>) The relative amount of each mRNA in wt and SEL1L-deficient cells was evaluated as described in the Materials and Methods. The amount of each mRNA in wt HEK293 cells without treatment was considered “1”. Each value represents the mean ± SEM from four independent cultures. Values marked with asterisks are significantly different from values value of untreated wild-type cells (* <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">
23 pages, 7888 KiB  
Article
Gene Expression-Related Changes in Morphologies of Organelles and Cellular Component Organization in Mucopolysaccharidoses
by Lidia Gaffke, Karolina Pierzynowska, Estera Rintz, Zuzanna Cyske, Izabela Giecewicz and Grzegorz Węgrzyn
Int. J. Mol. Sci. 2021, 22(5), 2766; https://doi.org/10.3390/ijms22052766 - 9 Mar 2021
Cited by 21 | Viewed by 3253
Abstract
Mucopolysaccharidoses (MPS) are inherited metabolic diseases characterized by accumulation of incompletely degraded glycosaminoglycans (GAGs) in lysosomes. Although primary causes of these diseases are mutations in genes coding for enzymes involved in lysosomal GAG degradation, it was demonstrated that storage of these complex carbohydrates [...] Read more.
Mucopolysaccharidoses (MPS) are inherited metabolic diseases characterized by accumulation of incompletely degraded glycosaminoglycans (GAGs) in lysosomes. Although primary causes of these diseases are mutations in genes coding for enzymes involved in lysosomal GAG degradation, it was demonstrated that storage of these complex carbohydrates provokes a cascade of secondary and tertiary changes affecting cellular functions. Potentially, this might lead to appearance of cellular disorders which could not be corrected even if the primary cause of the disease is removed. In this work, we studied changes in cellular organelles in MPS fibroblasts relative to control cells. All 11 types and subtypes of MPS were included into this study to obtain a complex picture of changes in organelles in this group of diseases. Two experimental approaches were employed, transcriptomic analyses and electron microscopic assessment of morphology of organelles. We analyzed levels of transcripts of genes grouped into two terms included into the QuickGO database, ‘Cellular component organization’ (GO:0016043) and ‘Cellular anatomical entity’ (GO:0110165), to find that number of transcripts with significantly changed levels in MPS fibroblasts vs. controls ranged from 109 to 322 (depending on MPS type) in GO:0016043, and from 70 to 208 in GO:0110165. This dysregulation of expression of genes crucial for proper structures and functions of various organelles was accompanied by severe changes in morphologies of lysosomes, nuclei, mitochondria, Golgi apparatus, and endoplasmic reticulum. Interestingly, some observed changes occurred in all/most MPS types while others were specific to particular disease types/subtypes. We suggest that severe changes in organelles in MPS cells might arise from dysregulation of expression of a battery of genes involved in organelles’ structures and functions. Intriguingly, normalization of GAG levels by using recombinant human enzymes specific to different MPS types corrected morphologies of some, but not all, organelles, while it failed to improve regulation of expression of selected genes. These results might suggest reasons for inability of enzyme replacement therapy to correct all MPS symptoms, particularly if initiated at advanced stages of the disease. Full article
(This article belongs to the Special Issue Genetic and Metabolic Molecular Research of Lysosomal Storage Disease)
Show Figures

Figure 1

Figure 1
<p>Levels of GAGs in tested lines of MPS fibroblasts relative to the control cell line (HDFa). The presented values are mean values from three independent experiments with error bars representing SD. Statistically significant differences (*; <span class="html-italic">p</span> &lt; 0.05) were found for each MPS type relative to HDFa.</p>
Full article ">Figure 2
<p>Number of up- and downregulated transcripts (at FDR &lt; 0.1; <span class="html-italic">p</span> &lt; 0.1) with division into selected processes (child terms of ‘cellular component organization’) in different MPS types relative to control cells (HDFa).</p>
Full article ">Figure 3
<p>Number of up- and downregulated transcripts (at FDR &lt; 0.1; <span class="html-italic">p</span> &lt; 0.1) with division into selected organelles (child terms of ‘cellular anatomical entity) in different MPS types relative to control cells (HDFa).</p>
Full article ">Figure 4
<p>Number of up- and down-regulated transcripts (at FDR &lt; 0.1; <span class="html-italic">p</span> &lt; 0.1) with division into selected organelles (child terms of ‘intracellular membrane-bounded organelle’ (GO:0043231) and ‘intracellular non-membrane-bounded organelle’ (GO:0043232)) in different MPS types relative to control cells (HDFa).</p>
Full article ">Figure 5
<p>Morphology of lysosomes in MPS fibroblasts in comparison to control cells (HDFa cell line). Representative electron micrographs are presented in (<b>A</b>) (size bars represent 500 nm; arrows indicate changes in these structures). Quantification of changes in lysosomes is presented in (<b>B</b>). Mean values ± SD are presented, with asterisks representing statistically significant (<span class="html-italic">p</span> &lt; 0.05) differences relative to the HDFa control.</p>
Full article ">Figure 6
<p>Morphology of nuclei in MPS fibroblasts in comparison to control cells (HDFa cell line). Representative electron micrographs are presented in (<b>A</b>) (size bars represent 2 mm). Quantification of changes in area of nucleus is presented in (<b>B</b>). Mean values ± SD are presented, with asterisks representing statistically significant (<span class="html-italic">p</span> &lt; 0.05) differences relative to the HDFa control.</p>
Full article ">Figure 7
<p>Morphology of mitochondria in MPS fibroblasts in comparison to control cells (HDFa cell line). Representative electron micrographs are presented in (<b>A</b>) (size bars represent 500 nm; arrows indicate mitochondria). Quantification of changes in mitochondria is presented in (<b>B</b>–<b>D</b>), indicating their number (<b>B</b>), length (<b>C</b>) and width (<b>D</b>). Mean values ± SD are presented, with asterisks representing statistically significant (<span class="html-italic">p</span> &lt; 0.05) differences relative to the HDFa control.</p>
Full article ">Figure 7 Cont.
<p>Morphology of mitochondria in MPS fibroblasts in comparison to control cells (HDFa cell line). Representative electron micrographs are presented in (<b>A</b>) (size bars represent 500 nm; arrows indicate mitochondria). Quantification of changes in mitochondria is presented in (<b>B</b>–<b>D</b>), indicating their number (<b>B</b>), length (<b>C</b>) and width (<b>D</b>). Mean values ± SD are presented, with asterisks representing statistically significant (<span class="html-italic">p</span> &lt; 0.05) differences relative to the HDFa control.</p>
Full article ">Figure 8
<p>Morphology of Golgi apparatus in MPS fibroblasts in comparison to control cells (HDFa cell line). Representative electron micrographs are presented in (<b>A</b>) (size bars represent 200 nm; arrows indicate Golgi structures). Quantification of changes in Golgi structures is presented in (<b>B</b>–<b>D</b>), indicating their number (<b>B</b>), length (<b>C</b>) and width (<b>D</b>). Mean values ± SD are presented, with asterisks representing statistically significant (<span class="html-italic">p</span> &lt; 0.05) differences relative to the HDFa control.</p>
Full article ">Figure 8 Cont.
<p>Morphology of Golgi apparatus in MPS fibroblasts in comparison to control cells (HDFa cell line). Representative electron micrographs are presented in (<b>A</b>) (size bars represent 200 nm; arrows indicate Golgi structures). Quantification of changes in Golgi structures is presented in (<b>B</b>–<b>D</b>), indicating their number (<b>B</b>), length (<b>C</b>) and width (<b>D</b>). Mean values ± SD are presented, with asterisks representing statistically significant (<span class="html-italic">p</span> &lt; 0.05) differences relative to the HDFa control.</p>
Full article ">Figure 9
<p>Morphology of endoplasmic reticulum in MPS fibroblasts in comparison to control cells (HDFa cell line). Representative electron micrographs are presented in (<b>A</b>) (size bars represent 500 nm). Quantification of changes in endoplasmic reticulum is presented in (<b>B</b>). Mean values ± SD are presented, with asterisks representing statistically significant (<span class="html-italic">p</span> &lt; 0.05) differences relative to the HDFa control.</p>
Full article ">Figure 10
<p>Levels of GAGs in HDFa (control) cells and MPS I and MPS II fibroblasts with and without treatment with 0.58 mg/L α-L-iduronidase (Aldurazyme, Ald) or 0.5 mg/L iduronate sulfatase (Elaprase, Ela) for 24 h. Mean values ± SD are presented, with asterisks and hashtags representing statistically significant (<span class="html-italic">p</span> &lt; 0.05) differences relative to the HDFa control and untreated MPS, respectively.</p>
Full article ">Figure 11
<p>Morphology of lysosomes (<b>A</b>), mitochondria (<b>B</b>), Golgi apparatus (<b>C</b>), and endoplasmic reticulum (<b>D</b>) in MPS fibroblasts in comparison to control cells (HDFa cell line) with and without treatment with 0.58 mg/L α-L-iduronidase (Aldurazyme, Ald) or 0.5 mg/L iduronate sulfatase (Elaprase, Ela) for 24 h. Mean values ± SD are presented, with asterisks and hashtags representing statistically significant (<span class="html-italic">p</span> &lt; 0.05) differences relative to the HDFa control and untreated MPS, respectively.</p>
Full article ">Figure 12
<p>Levels of periostin and CAPG proteins in MPS fibroblasts in comparison to control cells (HDFa cell line) with and without treatment with 0.58 mg/L α-L-iduronidase (Aldurazyme, Ald) or 0.5 mg/L iduronate sulfatase (Elaprase, Ela) for 24 h. Representative Western-blots are demonstrated in (<b>A</b>), while quantification of relative levels (assuming 1 as the value in untreated HFDa cells) of periostin and CAPG are shown in (<b>B</b>,<b>C</b>), respectively. Mean values ± SD are presented, with asterisks and hashtags representing statistically significant (<span class="html-italic">p</span> &lt; 0.05) differences relative to the HDFa control and untreated MPS, respectively.</p>
Full article ">
21 pages, 4376 KiB  
Article
Butyrate Improves Skin/Lung Fibrosis and Intestinal Dysbiosis in Bleomycin-Induced Mouse Models
by Hee Jin Park, Ok-Yi Jeong, Sung Hak Chun, Yun Hong Cheon, Mingyo Kim, Suhee Kim and Sang-Il Lee
Int. J. Mol. Sci. 2021, 22(5), 2765; https://doi.org/10.3390/ijms22052765 - 9 Mar 2021
Cited by 34 | Viewed by 5861
Abstract
Systemic sclerosis (SSc) is an autoimmune disorder characterized by fibrosis of the skin and internal organs. Despite several studies on SSc treatments, effective treatments for SSc are still lacking. Since evidence suggests an association between intestinal microbiota and SSc, we focused on butyrate, [...] Read more.
Systemic sclerosis (SSc) is an autoimmune disorder characterized by fibrosis of the skin and internal organs. Despite several studies on SSc treatments, effective treatments for SSc are still lacking. Since evidence suggests an association between intestinal microbiota and SSc, we focused on butyrate, which has beneficial effects in autoimmune diseases as a bacterial metabolite. Here, we investigated the therapeutic potential of sodium butyrate (SB) using a bleomycin-induced fibrosis mouse model of SSc and human dermal fibroblasts (HDFs). SB attenuated bleomycin-induced dermal and lung fibrosis in mice. SB influenced fecal microbiota composition (phyla Actinobacteria and Bacteroidetes, genera Bifidobacterium and Ruminococcus_g2). SB controlled macrophage differentiation in mesenteric lymph nodes, spleen, and bronchoalveolar lavage cells of mice with bleomycin-induced skin fibrosis. Profibrotic and proinflammatory gene expression was suppressed by SB administration in skin. Furthermore, SB inhibited transforming growth factor β1-responsive proinflammatory expression with increased acetylation of histone 3 in HDFs. Subcutaneous SB application had antifibrogenic effects on the skin. Butyrate ameliorated skin and lung fibrosis by improving anti-inflammatory activity in a mouse model of SSc. Butyrate may exhibit indirect and direct anti-fibrogenic action on fibroblasts by regulating macrophage differentiation and inhibition of histone deacetylase 3. These findings suggest butyrate as an SSc treatment. Full article
(This article belongs to the Section Molecular Immunology)
Show Figures

Figure 1

Figure 1
<p>Antifibrotic effects of butyrate in the BLM-induced skin fibrosis mouse model. Bleomycin (BLM) was injected subcutaneously to the back skin of mice five times a week for two weeks. Sodium butyrate (SB) was orally gavaged from two weeks before BLM injection. Skin tissues were then obtained in normal and BLM ± SB mice to evaluate fibrosis. (<b>A</b>) Representative images of the Masson’s trichrome stain and dermal thickness (arrow length) in skin tissues. Data are representative of at least four independent experiments with <span class="html-italic">n</span> = 5–7/group. (<b>B</b>–<b>D</b>) Representative immunofluorescence images of skin stained with DAPI (blue) and α-SMA (green) (<b>B</b>), representative Western blotting and quantitative analysis of α-SMA expression (<b>C</b>), and collagen content (<b>D</b>) in the skin. Data are representative of two independent experiments with <span class="html-italic">n</span> = 3–5/group; each symbol represents one mouse. Scale bars = 100 µm. DAPI: 4′,6-diamidino-2-phenylindole. ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 2
<p>Composition of fecal microbiota. Bleomycin (BLM) was injected subcutaneously for two weeks. Sodium butyrate (SB) was administered orally from two weeks before BLM injection and feces were then obtained. The 16S sequences of fecal microbiota were analyzed in normal and BLM ± SB mice. (<b>A</b>) Microbiota richness estimated by the number of operational taxonomic units (OTUs) and Shannon index. (<b>B</b>) Microbiota diversity evaluated by principal coordinate analysis plots of UniFrac distances. (<b>C</b>,<b>D</b>) The overall composition of gut microbiota (left) and statistically significant changes (right) at the phylum level (<b>C</b>) and at the family level (<b>D</b>). (<b>E</b>) Heatmap of differentially abundant microbial genera among groups. Color represents normalized (z-score) relative abundance of bacteria from green (low abundance) to red (high abundance) (<b>E</b>). <span class="html-italic">n</span> = 3/group; each symbol represents one mouse. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 in (<b>C</b>,<b>D</b>). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 vs. normal, # <span class="html-italic">p</span> &lt; 0.05 vs. BLM in (<b>E</b>).</p>
Full article ">Figure 3
<p>Flow cytometric analysis of immune cell populations in the mesenteric lymph nodes (MLN) and spleen of normal and bleomycin (BLM) ± sodium butyrate (SB) mice. BLM was injected subcutaneously for two weeks. SB was administered orally from two weeks before BLM injection. Immune cells were then obtained from the MLN and spleen. (<b>A</b>) Flow cytometric gating strategy for macrophage and DC lineages. (<b>B</b>) Flow cytometry was used to assess total leukocytes (CD45<sup>+</sup>), macrophages (CD45<sup>+</sup>CD64<sup>+</sup>CD11b<sup>+</sup>CX<sub>3</sub>CR1<sup>+</sup>), DCs (CD45<sup>+</sup>CD64<sup>−</sup>CD11c<sup>+</sup>MHCII<sup>+</sup>), B cells (B220+), and CD4+ T cells in MLN and spleen from normal and BLM ± SB mice. (<b>C</b>) Macrophage lineages were subdivided into phase 1 (P1: newly recruited monocytes), phase 2 (P2: maturing monocytes), and phase 3 (P3: monocyte/macrophage intermediates and resident macrophages) following developmental stages. Inflammatory monocytes were defined as both Ly6C<sup>+</sup>MHCII<sup>−</sup> (P1) and Ly6C<sup>+</sup>MHCII<sup>+</sup> (P2). Representative dot plots, percentages (gated on macrophage lineage) and absolute numbers of inflammatory monocytes are shown in the MLN (upper) and spleen (lower) from normal and BLM ± SB mice. Data are representative of two independent experiments with <span class="html-italic">n</span> = 4–5/group; each symbol represents one mouse. MHCII: Major histocompatibility complex class II. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 4
<p>Expression of CD11b<sup>+</sup>CX<sub>3</sub>CR1<sup>+</sup> macrophages in skin tissues from normal and bleomycin (BLM) ± sodium butyrate (SB) mice. BLM was injected subcutaneously for two weeks. SB was orally administered from two weeks before BLM injection and skin samples were then obtained. Representative immunofluorescent images of skin stained for DAPI (blue), CD11b (green), and CX<sub>3</sub>CR1 (red). Magnification, 400×. Scale bars = 100 µm. <span class="html-italic">n</span> = 6/group.</p>
Full article ">Figure 5
<p>Expression profiles of profibrotic and proinflammatory genes in skin from normal and bleomycin (BLM) ± sodium butyrate (SB) mice. BLM was injected subcutaneously for two weeks. SB was administered orally from two weeks before BLM injection and skin samples were then obtained. <span class="html-italic">n</span> = 4–6/group; each symbol represents one mouse. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 6
<p>Antifibrotic effect of butyrate on dermal fibroblasts. (<b>A</b>–<b>C</b>) Primary human dermal fibroblasts (HDFs) were stimulated with TGF-β1 (10 ng/mL) with or without sodium butyrate (SB) (0.5–1 mM) for 24–48 h. Western blotting or qPCR was performed to analyze protein and mRNA expression in the cell lysate from fibroblasts. Representative data of three independent experiments. (<b>A</b>) Expression of the α-SMA protein by Western blotting. (<b>B</b>) Expression of mRNA for <span class="html-italic">ACTA2</span>, <span class="html-italic">COL1A1</span>, <span class="html-italic">IL6</span>, and <span class="html-italic">IL1B</span> by qPCR. (<b>C</b>) Acetylated histone H3 (Ac-H3) and histone H3 (H3) protein expression. (<b>D</b>) BLM was injected subcutaneously to the back skin of mice five times a week for two weeks with simultaneous subcutaneous (s.c.) injection of SB or phosphate-buffered saline (PBS). Representative images of the Masson’s trichrome stain and dermal thickness in skin sections. <span class="html-italic">n</span> = 6/group (<span class="html-italic">n</span> = 2 for normal); each symbol represents one mouse. Scale bars = 100 µm, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 7
<p>Antifibrotic effects of sodium butyrate (SB) in bleomycin (BLM)-induced lung fibrosis mouse models. (<b>A</b>) The BLM-induced pulmonary fibrosis model was used to investigate the effect of SB on lung fibrosis in mice. BLM was administered once intratracheally (i.t.). SB was orally gavaged for five weeks, starting two weeks before BLM injection. Representative images of the Masson’s trichrome stain and histological score (Ashcroft score) in lung tissues from normal and BLM ± SB mice. Scale bars = 500 µm. (<b>B</b>–<b>E</b>) BLM-induced skin fibrosis model was used to investigate the effect of SB on lung fibrosis in mice. BLM was injected subcutaneously (s.c.) for two weeks. SB was administered orally from two weeks before BLM injection and lung tissues were then obtained. (<b>B</b>) Representative images of the Masson’s trichrome stain and Ashcroft score in lung tissues from normal and BLM ± SB mice. The results were verified through four independent experiments. <span class="html-italic">n</span> = 5–7/group in each experiment. Scale bars = 500 µm. (<b>C</b>) Representative immunohistochemical image of α-SMA (brown colors: arrows). Scale bars = 100 µm. (<b>D</b>) Representative Western blotting and quantitative analysis of α-SMA expression. (<b>E</b>) Total collagen content in lung tissues from normal and BLM ± SB mice. <span class="html-italic">n</span> = 3–6/group; each symbol represents one mouse. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 8
<p>Modulation of macrophage populations in the bronchoalveolar lavage fluid (BALF). (<b>A</b>,<b>B</b>) Flow cytometry was used to assess macrophage subsets in the BALF from normal and bleomycin (BLM) ± sodium butyrate (SB) mice in the BLM-induced skin fibrosis model. (<b>A</b>) Total leukocytes (CD45<sup>+</sup>) and macrophages (CD45<sup>+</sup>CD64<sup>+</sup>F4/80<sup>+</sup>) in the BALF from normal and BLM ± SB mice. (<b>B</b>) LMs were divided into AMs (CD11c<sup>+</sup>CD11b<sup>−</sup>) and IMs (CD11b<sup>+</sup>). Representative dot plots, percentages (gated on macrophages), and absolute numbers of AMs and IMs are shown in the BALF from normal and BLM ± SB mice. (<b>C</b>) Expression levels of mRNA of profibrotic and proinflammatory mediators in lung tissues from normal and BLM ± SB mice. <span class="html-italic">n</span> = 3–5/group; each symbol represents one mouse. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">
17 pages, 3399 KiB  
Article
Reduction of Amyloid Burden by Proliferated Homeostatic Microglia in Toxoplasma gondii-Infected Alzheimer’s Disease Model Mice
by Ji-Hun Shin, Young Sang Hwang, Bong-Kwang Jung, Seung-Hwan Seo, Do-Won Ham and Eun-Hee Shin
Int. J. Mol. Sci. 2021, 22(5), 2764; https://doi.org/10.3390/ijms22052764 - 9 Mar 2021
Cited by 8 | Viewed by 4239
Abstract
In this study, we confirmed that the number of resident homeostatic microglia increases during chronic Toxoplasma gondii infection. Given that the progression of Alzheimer’s disease (AD) worsens with the accumulation of amyloid β (Aβ) plaques, which are eliminated through microglial phagocytosis, we hypothesized [...] Read more.
In this study, we confirmed that the number of resident homeostatic microglia increases during chronic Toxoplasma gondii infection. Given that the progression of Alzheimer’s disease (AD) worsens with the accumulation of amyloid β (Aβ) plaques, which are eliminated through microglial phagocytosis, we hypothesized that T. gondii-induced microglial proliferation would reduce AD progression. Therefore, we investigated the association between microglial proliferation and Aβ plaque burden using brain tissues isolated from 5XFAD AD mice (AD group) and T. gondii-infected AD mice (AD + Toxo group). In the AD + Toxo group, amyloid plaque burden significantly decreased compared with the AD group; conversely, homeostatic microglial proliferation, and number of plaque-associated microglia significantly increased. As most plaque-associated microglia shifted to the disease-associated microglia (DAM) phenotype in both AD and AD + Toxo groups and underwent apoptosis after the lysosomal degradation of phagocytosed Aβ plaques, this indicates that a sustained supply of homeostatic microglia is required for alleviating Aβ plaque burden. Thus, chronic T. gondii infection can induce microglial proliferation in the brains of mice with progressed AD; a sustained supply of homeostatic microglia is a promising prospect for AD treatment. Full article
(This article belongs to the Section Molecular Neurobiology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Proliferation and activation of resident homeostatic microglia over 36 weeks after <span class="html-italic">Toxoplasma gondii</span> infection. (<b>A</b>) <span class="html-italic">T. gondii</span> infection was confirmed by the presence of cysts in brain tissue (H&amp;E staining; scale bar, 20 µm). (<b>B</b>) Microglia in the hippocampal formation were stained with Iba1 (red; scale bar, 100 µm). (<b>C</b>) Microglia in the hippocampal formation were stained with TMEM119 (red; scale bar, 100 µm). (<b>D</b>) Activated microglia were co-stained with CD11b and Iba1 (green and red, respectively; scale bar, 50 µm). (<b>E</b>,<b>F</b>) Mean fluorescence intensity (MFI) was calculated from fluorescence-stained images (<b>B</b>,<b>C</b>) using ImageJ. The fold changes of MFI at 3, 6, 12, and 36 weeks PI were compared with those of the control (0 weeks). H, hippocampus; C, cortex. (<b>G</b>) The number of activated microglia (Iba1<sup>+</sup>/CD11b<sup>+</sup>) was designated by cell number per mm<sup>2</sup> of the brain tissue. (<b>H</b>) Ki67-stained proliferating microglia. The yellow arrow head shows microglial cells in the mitotic phase (sky blue) co-stained with DAPI (blue) and Ki67 (green), and the number of proliferating microglia (Iba1<sup>+</sup>/Ki67<sup>+</sup>) was designated by cell number/mm<sup>2</sup> brain tissue. Scale bar; 20 µm. Quantification of MFI intensity and co-stained cell counting; two brain sections per mouse. (<b>I</b>) Microarray analysis of genes encoding trophic factors, homeostatic markers, and M1 and M2 markers in <span class="html-italic">T. gondii</span>-infected brain. Data are represented as the mean ± SEM. * Statistical significance compared with the control (* <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>Reduction of Aβ plaques and microglial proliferation in <span class="html-italic">T. gondii</span>-infected 5XFAD mouse brain. (<b>A</b>) A <span class="html-italic">T. gondii</span> cyst (black arrow) and stained amyloid plaque (blue arrow head) in a brain section (H&amp;E and Congo red staining) at 40 weeks PI. Scale bar; 20 µm. (<b>B</b>) Dense core plaques (black spots) stained using Congo red dye (scale bar; 100 µm) and acidophilic neurons (black quadrangle) in cortical layer V (scale bar; 20 µm). (<b>C</b>) The number of dense core plaques counted in Congo red-stained brain. (<b>D</b>) Concentration of Aβ in brain tissue lysate analyzed by ELISA. (<b>E</b>) DAB-color immunohistochemistry of microglia around Aβ plaques. Scale bar; 20 µm. (<b>F</b>) Microglial cells accumulated; Iba1-stained microglia in the brain. The fold change of MFI in the AD + Toxo group compared with the AD group in the Iba1-stained brain. Scale bar; 100 µm. H, hippocampus; C, cortex. (<b>G</b>) TMEM119-stained homeostatic microglia accumulation. The fold changes of MFI in the TMEM119-stained brain. Scale bar; 100 µm. H, hippocampus; C, cortex. (<b>H</b>) Protein expression of microglial trophic factors (IL-1β and TNF-α) and microglial polarization inducers (IFN-γ and IL-4) examined by ELISA. (<b>I</b>) Microarray analysis for microglial trophic factors (<span class="html-italic">Il1β</span>, <span class="html-italic">Tnfα</span>, <span class="html-italic">Mcsf</span>, and <span class="html-italic">NFKB1</span>), homeostatic microglial markers (<span class="html-italic">P2ry13</span>, <span class="html-italic">Cx3cr1</span>, and <span class="html-italic">Tmem119</span>), and inducers and markers of M1 and M2 polarized microglia (<span class="html-italic">Ifnγ</span>, <span class="html-italic">Il4</span>, <span class="html-italic">Cd86</span>, and <span class="html-italic">Cd206</span>) in AD and AD + Toxo groups compared with the gene expression in wild-type (WT) mice. Data are represented as the mean ± SEM. * Statistical significance compared with the control (* <span class="html-italic">p</span> &lt; 0.05). <sup>#</sup> Statistical significance compared with each experimental group (<sup>#</sup> <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>Plaque-associated patterns of microglia and Ly6C<sup>+</sup> monocytes. (<b>A</b>) Iba-1-stained microglia (red) around methoxy-XO4-stained amyloid β (Aβ) plaques (white). The counting result of microglia (plaque-associated or plaque-free) found around 60 amyloid plaques distributed in the brain ((10 randomly selected plaques per mouse) × 6 mice per group). Scale bar; 25 µm. (<b>B</b>) Plaque-associated homeostatic microglia (yellow, due to co-staining of Iba-1 (red) and TMEM119 (green) and indicated by an arrow). Number of plaque-associated homeostatic microglial cells. Scale bar; 20 µm. (<b>C</b>) Ly6C<sup>+</sup> monocytes (green) accumulated around Aβ plaque in brain tissues of both AD and AD + Toxo groups. (Ly6C<sup>+</sup> (green) monocytes indicated by white arrow). Number of plaque-associated Ly6C<sup>+</sup> monocytes. Scale bar; 20 µm. Data are represented as the mean ± SEM. * Statistical significance compared with the control (* <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Disease-associated microglia (DAM) phenotype of microglia and microglial phagocytosis. (<b>A</b>) DAM- or homeostatic markers of microglia investigated by gene array analysis. (<b>B</b>) qPCR results for <span class="html-italic">Cst7</span> and <span class="html-italic">Tmem119</span>. (<b>C</b>) The TREM2 signal is a predictor of the DAM phenotype. (yellow, due to co-staining of Iba-1 (red) and TREM2 (green) and indicated by an arrow); Scale bar; 20 µm. (<b>D</b>) Lipoprotein lipase (LPL) signal as a stage two marker in DAM. The number of plaque-associated and LPL-expressing microglia. The counting result of microglia found around 60 amyloid plaques distributed in the brain ((10 randomly selected plaques per mouse) × 6 mice per group). (yellow, due to co-staining of Iba-1 (red) and LPL (green) and indicated by an arrow); Scale bar; 20 µm. (<b>E</b>) LAMP1 expression, and colocalization of LAMP1 and Iba1 in microglia around the amyloid β (Aβ) plaque. Lysosomal degradation of Aβ plaque as seen by the internalized puncta per microglial cell. Both “a” and “b” indicate LAMP1 positive lysosomes (white and yellow arrow) with the colocalization of methoxy-X04 stained amyloid β (*). Scale bar; 20 µm. (<b>F</b>) Internalized puncta per microglial cell indicate lysosomal degradation of the Aβ plaque (yellow arrow). Scale bar; 20 µm. Data are represented as the mean ± SEM. * Statistical significance compared with the control (* <span class="html-italic">p</span> &lt; 0.05). <sup>#</sup> Statistical significance compared with each experimental group (<sup>#</sup> <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Apoptosis of plaque-associated microglia. TUNEL<sup>+</sup>-microglial cells surrounding the amyloid β (Aβ) plaque. Arrows represent Iba1 and TUNEL double-positive cells (light blue). Number of Iba1 and TUNEL double-positive cells in plaque-associated microglia. The counting result of plaque-associated apoptotic microglia found around 60 amyloid plaques ((randomly selected 10 plaques per mouse) × 6 mice per group). Scale bar; 20 µm. Data are represented as the mean ± SEM.</p>
Full article ">
17 pages, 972 KiB  
Review
Dysbiosis in the Development of Type I Diabetes and Associated Complications: From Mechanisms to Targeted Gut Microbes Manipulation Therapies
by Gratiela Gradisteanu Pircalabioru, Nicolae Corcionivoschi, Ozan Gundogdu, Mariana-Carmen Chifiriuc, Luminita Gabriela Marutescu, Bogdan Ispas and Octavian Savu
Int. J. Mol. Sci. 2021, 22(5), 2763; https://doi.org/10.3390/ijms22052763 - 9 Mar 2021
Cited by 18 | Viewed by 5343
Abstract
Globally, we are facing a worrying increase in type 1 diabetes mellitus (T1DM) incidence, with onset at younger age shedding light on the need to better understand the mechanisms of disease and step-up prevention. Given its implication in immune system development and regulation [...] Read more.
Globally, we are facing a worrying increase in type 1 diabetes mellitus (T1DM) incidence, with onset at younger age shedding light on the need to better understand the mechanisms of disease and step-up prevention. Given its implication in immune system development and regulation of metabolism, there is no surprise that the gut microbiota is a possible culprit behind T1DM pathogenesis. Additionally, microbiota manipulation by probiotics, prebiotics, dietary factors and microbiota transplantation can all modulate early host–microbiota interactions by enabling beneficial microbes with protective potential for individuals with T1DM or at high risk of developing T1DM. In this review, we discuss the challenges and perspectives of translating microbiome data into clinical practice. Nevertheless, this progress will only be possible if we focus our interest on developing numerous longitudinal, multicenter, interventional and double-blind randomized clinical trials to confirm their efficacy and safety of these therapeutic approaches. Full article
(This article belongs to the Special Issue Gut Microbiota and Immunity)
Show Figures

Figure 1

Figure 1
<p>Gut microbiota and T1DM. T1DM patients harbor a microbiota with reduced diversity enriched in <span class="html-italic">Bacteroides</span> species (<span class="html-italic">B. dorei</span>, <span class="html-italic">B. ovatus</span>, <span class="html-italic">B. logum</span>). The gut environment in T1DM is characterized by increased gut permeability, disrupted mucus barrier, inflammation and low production of SCFAs (particularly butyrate).</p>
Full article ">Figure 2
<p>Proposed personalized therapeutic approaches in T1DM.</p>
Full article ">
12 pages, 1136 KiB  
Review
Bone Marrow Aspirate Matrix: A Convenient Ally in Regenerative Medicine
by José Fábio Lana, Lucas Furtado da Fonseca, Gabriel Azzini, Gabriel Santos, Marcelo Braga, Alvaro Motta Cardoso Junior, William D. Murrell, Alberto Gobbi, Joseph Purita and Marco Antonio Percope de Andrade
Int. J. Mol. Sci. 2021, 22(5), 2762; https://doi.org/10.3390/ijms22052762 - 9 Mar 2021
Cited by 14 | Viewed by 4073
Abstract
The rise in musculoskeletal disorders has prompted medical experts to devise novel effective alternatives to treat complicated orthopedic conditions. The ever-expanding field of regenerative medicine has allowed researchers to appreciate the therapeutic value of bone marrow-derived biological products, such as the bone marrow [...] Read more.
The rise in musculoskeletal disorders has prompted medical experts to devise novel effective alternatives to treat complicated orthopedic conditions. The ever-expanding field of regenerative medicine has allowed researchers to appreciate the therapeutic value of bone marrow-derived biological products, such as the bone marrow aspirate (BMA) clot, a potent orthobiologic which has often been dismissed and regarded as a technical complication. Numerous in vitro and in vivo studies have contributed to the expansion of medical knowledge, revealing optimistic results concerning the application of autologous bone marrow towards various impactful disorders. The bone marrow accommodates a diverse family of cell populations and a rich secretome; therefore, autologous BMA-derived products such as the “BMA Matrix”, may represent a safe and viable approach, able to reduce the costs and some drawbacks linked to the expansion of bone marrow. BMA provides —it eliminates many hurdles associated with its preparation, especially in regards to regulatory compliance. The BMA Matrix represents a suitable alternative, indicated for the enhancement of tissue repair mechanisms by modulating inflammation and acting as a natural biological scaffold as well as a reservoir of cytokines and growth factors that support cell activity. Although promising, more clinical studies are warranted in order to further clarify the efficacy of this strategy. Full article
(This article belongs to the Section Molecular Biology)
Show Figures

Figure 1

Figure 1
<p>Bone marrow aspirate (BMA) Matrix. The combination of BMA with hyaluronic acid (HA) enables the formation of the BMA matrix, proposing a novel strategy towards the enhancement of regenerative mechanisms and possible synergistic effects. The crosslinked fibrin fibers derived from the BMA clot form a natural scaffold, supporting cellular activity and reconstruction of damaged tissue.</p>
Full article ">Figure 2
<p>Preparation of fresh autologous bone marrow aspirate matrix. (<b>A</b>) Application of local anesthesia; (<b>B</b>) target point incision; (<b>C</b>) posterior iliac crest puncture; (<b>D</b>) aspiration of fresh bone marrow; (<b>E</b>) anticoagulant-free syringes filled with bone marrow aspirate; (<b>F</b>) bone marrow aspirate mixed with hyaluronic acid; (<b>G</b>) Bone Marrow Aspirate Matrix inside the 3 mL syringes (indicated by the blue bracket).</p>
Full article ">
12 pages, 265 KiB  
Review
Machine Learning and Novel Biomarkers for the Diagnosis of Alzheimer’s Disease
by Chun-Hung Chang, Chieh-Hsin Lin and Hsien-Yuan Lane
Int. J. Mol. Sci. 2021, 22(5), 2761; https://doi.org/10.3390/ijms22052761 - 9 Mar 2021
Cited by 121 | Viewed by 13728
Abstract
Background: Alzheimer’s disease (AD) is a complex and severe neurodegenerative disease that still lacks effective methods of diagnosis. The current diagnostic methods of AD rely on cognitive tests, imaging techniques and cerebrospinal fluid (CSF) levels of amyloid-β1-42 (Aβ42), total tau protein and hyperphosphorylated [...] Read more.
Background: Alzheimer’s disease (AD) is a complex and severe neurodegenerative disease that still lacks effective methods of diagnosis. The current diagnostic methods of AD rely on cognitive tests, imaging techniques and cerebrospinal fluid (CSF) levels of amyloid-β1-42 (Aβ42), total tau protein and hyperphosphorylated tau (p-tau). However, the available methods are expensive and relatively invasive. Artificial intelligence techniques like machine learning tools have being increasingly used in precision diagnosis. Methods: We conducted a meta-analysis to investigate the machine learning and novel biomarkers for the diagnosis of AD. Methods: We searched PubMed, the Cochrane Central Register of Controlled Trials, and the Cochrane Database of Systematic Reviews for reviews and trials that investigated the machine learning and novel biomarkers in diagnosis of AD. Results: In additional to Aβ and tau-related biomarkers, biomarkers according to other mechanisms of AD pathology have been investigated. Neuronal injury biomarker includes neurofiliament light (NFL). Biomarkers about synaptic dysfunction and/or loss includes neurogranin, BACE1, synaptotagmin, SNAP-25, GAP-43, synaptophysin. Biomarkers about neuroinflammation includes sTREM2, and YKL-40. Besides, d-glutamate is one of coagonists at the NMDARs. Several machine learning algorithms including support vector machine, logistic regression, random forest, and naïve Bayes) to build an optimal predictive model to distinguish patients with AD from healthy controls. Conclusions: Our results revealed machine learning with novel biomarkers and multiple variables may increase the sensitivity and specificity in diagnosis of AD. Rapid and cost-effective HPLC for biomarkers and machine learning algorithms may assist physicians in diagnosing AD in outpatient clinics. Full article
20 pages, 2299 KiB  
Article
Anti-Apoptotic Effect of Apelin in Human Placenta: Studies on BeWo Cells and Villous Explants from Third-Trimester Human Pregnancy
by Ewa Mlyczyńska, Małgorzata Myszka, Patrycja Kurowska, Monika Dawid, Tomasz Milewicz, Marta Bałajewicz-Nowak, Paweł Kowalczyk and Agnieszka Rak
Int. J. Mol. Sci. 2021, 22(5), 2760; https://doi.org/10.3390/ijms22052760 - 9 Mar 2021
Cited by 19 | Viewed by 3123
Abstract
Previously, we demonstrated the expression of apelin and G-protein-coupled receptor APJ in human placenta cell lines as well as its direct action on placenta cell proliferation and endocrinology. The objective of this study was to examine the effect of apelin on placenta apoptosis [...] Read more.
Previously, we demonstrated the expression of apelin and G-protein-coupled receptor APJ in human placenta cell lines as well as its direct action on placenta cell proliferation and endocrinology. The objective of this study was to examine the effect of apelin on placenta apoptosis in BeWo cells and villous explants from the human third trimester of pregnancy. The BeWo cells and villous explants were incubated with apelin (2 and 20 ng/mL) alone or with staurosporine for 24 to 72 h. First, we analysed the dose- and time-dependent effect of apelin on the expression of apoptotic factors on the mRNA level by real-time PCR and on the protein level using Western blot. Next, we checked caspase 3 and 7 activity by Caspase-Glo 3/7, DNA fragmentation by the Cell Death Detection ELISA kit and oxygen consumption by the MitoXpress-Xtra Oxygen Consumption assay. We found that apelin increased the expression of pro-survival and decreased proapoptotic factors on mRNA and protein levels in both BeWo cells and villous explants. Additionally, apelin inhibited caspase 3 and 7 activity and DNA fragmentation in staurosporine-induced apoptosis as also attenuated oxidative stress by increasing extracellular oxygen consumption. The antiapoptotic effect of apelin in BeWo cells was mediated by the APJ receptor and mitogen-activated protein kinase (ERK1/2/MAP3/1) and protein kinase B (AKT). The obtained results showed the antiapoptotic effect of apelin on trophoblast cells, suggesting its participation in the development of the placenta. Full article
(This article belongs to the Special Issue Adipokines in Health and Diseases)
Show Figures

Figure 1

Figure 1
<p>Effect of apelin on protein expression of apoptotic factors in BeWo cells. The cells were incubated with apelin at doses 2 of (AP2) and 20 (AP20) ng/mL for 24, 48 and 72 h, and subsequently, Western blot analysis was performed to examine the expression of BCL2 (B-cell like lymphoma 2), BAX (Bcl-2-like protein 4), caspase 3, 8 and 9 and p53. Results are shown as stripes on gel image (<b>A</b>) and densitometry analysis relative to β-actin (<b>B</b>–<b>F</b>). Experiments were independently performed and repeated three times (<span class="html-italic">n</span> = 3). The data are arranged as means ± SEM. Different letters indicate significant differences (<span class="html-italic">p</span> &lt; 0.05) among groups; Control (<b>C</b>).</p>
Full article ">Figure 2
<p>Effect of apelin on caspase 3 and 7 activity and histone-associated DNA fragment level in BeWo cells. The cells were incubated with 2 and 20 ng/mL of apelin alone or in combination with 0.1 µL/mL staurosporine for 24, 48 and 72 h, after which caspase 3 and 7 activity (<b>A</b>,<b>B</b>) was analysed using the Caspase-Glo 3/7 assay or the level of histone-associated DNA fragments (<b>C</b>,<b>D</b>) by the Cell Death Detection ELISA kit. Experiments were independently performed and repeated three times (<span class="html-italic">n</span> = 3). The data are arranged as means ± SEM. Different letters indicate significant differences (<span class="html-italic">p</span> &lt; 0.05) among groups; Control (<b>C</b>), Staurosporine (S), Relative Luminescence Unit (RLU), Relative Absorbance Unit (RAU).</p>
Full article ">Figure 2 Cont.
<p>Effect of apelin on caspase 3 and 7 activity and histone-associated DNA fragment level in BeWo cells. The cells were incubated with 2 and 20 ng/mL of apelin alone or in combination with 0.1 µL/mL staurosporine for 24, 48 and 72 h, after which caspase 3 and 7 activity (<b>A</b>,<b>B</b>) was analysed using the Caspase-Glo 3/7 assay or the level of histone-associated DNA fragments (<b>C</b>,<b>D</b>) by the Cell Death Detection ELISA kit. Experiments were independently performed and repeated three times (<span class="html-italic">n</span> = 3). The data are arranged as means ± SEM. Different letters indicate significant differences (<span class="html-italic">p</span> &lt; 0.05) among groups; Control (<b>C</b>), Staurosporine (S), Relative Luminescence Unit (RLU), Relative Absorbance Unit (RAU).</p>
Full article ">Figure 3
<p>Effect of apelin on oxidative stress in BeWo cells. The cells were treated with apelin at 2 and 20 ng/mL for 24 h, and the oxygen consumption assay was performed. Experiments were independently performed and repeated three times (<span class="html-italic">n</span> = 3). The data are arranged as means ± SEM. Different letters indicate significant differences <span class="html-italic">(p</span> &lt; 0.05) among groups; Control (C).</p>
Full article ">Figure 4
<p>Involvement of APJ receptor, mitogen-activated protein kinase (ERK1/2/MAP3/1) and protein kinase B (AKT) in the antiapoptotic effect of apelin in BeWo cells: The cells were pre-treated for 1 h with APJ receptor, MAP3/1 and AKT kinase inhibitors ML221 (5 µM), PD098059 (1 µM) and LY290042 (1 µM). Subsequently, apelin at a dose of 2 (AP2) ng/mL was added. After 72 h of incubation, activity of caspase 3 and 7 was analysed using the Caspase-Glo 3/7 assay. Experiments were independently performed and repeated three times (<span class="html-italic">n</span> = 3). The results are presented as a percentage compared to the control (100%). The data are arranged as means ± SEM. Different letters indicate significant differences (<span class="html-italic">p</span> &lt; 0.05) among groups; Control (C).</p>
Full article ">Figure 5
<p>Effect of apelin on DNA fragmentation in villous explants from human placenta. The explants were incubated with apelin at 2 and 20 ng/mL for 24, 48 and 72 h, and the level of histone-associated DNA fragment was measured using the Cell Death Detection ELISA kit. Experiments were performed on five independent cultures of human placenta explants (<span class="html-italic">n</span> = 5). The level of DNA fragmentation was calculated to 100 µg protein of villous explants and then as percentage compared to the 24-h control sample (100%). The data are arranged as means ± SEM. Different letters indicate significant differences (<span class="html-italic">p</span> &lt; 0.05) among groups; Control (C).</p>
Full article ">Figure 6
<p>Effect of apelin on protein expression of apoptotic proteins in villous explants from human placenta. The explants were incubated with 2 (AP2) and 20 (AP20) ng/mL of apelin for 24, 48 and 72 h, and Western blot analysis was performed to determine the expression of BCL2, BAX and caspase 3. Results are shown as stripes on gel image (<b>A</b>) and densitometry analysis relative to β-actin (<b>B</b>). Experiments were performed on five independent cultures of human placenta explants (<span class="html-italic">n</span> = 5). The data are arranged as means ± SEM. Different letters indicate significant differences (<span class="html-italic">p</span> &lt; 0.05) among groups; Control (C).</p>
Full article ">Figure 7
<p>Model of apelin antiapoptotic action in BeWo cell line (<b>A</b>) and human placenta explants (<b>B</b>). Apelin inhibits apoptosis by activation of apelin receptor (APJ), mitogen-activated kinase (MAP3/1) and protein kinase B (AKT); BCL2 (B-cell lymphoma 2), BAX (Bcl-2-like protein 4); BAK1 (Bcl-2 homologous antagonist/killer); BOK (Bcl-2-related ovarian killer protein); NOD1 (Nucleotide-binding oligomerisation domain-containing protein 1); CRADD (Caspase and RIP adapter with death domain); DIABLO (Diablo IAP-Binding Mitochondrial Protein); TNFRSF25 (Tumour necrosis factor receptor superfamily member 25); BIRC6 (Baculoviral IAP repeat-containing protein 6); MCL1 (Induced myeloid leukaemia cell differentiation protein MCL1); XIAP (X-linked inhibitor of apoptosis).</p>
Full article ">
16 pages, 3740 KiB  
Article
Increased Presence of Complement Factors and Mast Cells in Alveolar Bone and Tooth Resorption
by Kathrin Luntzer, Ina Lackner, Birte Weber, Yvonne Mödinger, Anita Ignatius, Florian Gebhard, Susann-Yvonne Mihaljevic, Melanie Haffner-Luntzer and Miriam Kalbitz
Int. J. Mol. Sci. 2021, 22(5), 2759; https://doi.org/10.3390/ijms22052759 - 9 Mar 2021
Cited by 3 | Viewed by 2782
Abstract
Periodontitis is the inflammatory destruction of the tooth-surrounding and -supporting tissue, resulting at worst in tooth loss. Another locally aggressive disease of the oral cavity is tooth resorption (TR). This is associated with the destruction of the dental mineralized tissue. However, the underlying [...] Read more.
Periodontitis is the inflammatory destruction of the tooth-surrounding and -supporting tissue, resulting at worst in tooth loss. Another locally aggressive disease of the oral cavity is tooth resorption (TR). This is associated with the destruction of the dental mineralized tissue. However, the underlying pathomechanisms remain unknown. The complement system, as well as mast cells (MCs), are known to be involved in osteoclastogenesis and bone loss. The complement factors C3 and C5 were previously identified as key players in periodontal disease. Therefore, we hypothesize that complement factors and MCs might play a role in alveolar bone and tooth resorption. To investigate this, we used the cat as a model because of the naturally occurring high prevalence of both these disorders in this species. Teeth, gingiva samples and serum were collected from domestic cats, which had an appointment for dental treatment under anesthesia, as well as from healthy cats. Histological analyses, immunohistochemical staining and the CH-50 and AH-50 assays revealed increased numbers of osteoclasts and MCs, as well as complement activity in cats with TR. Calcifications score in the gingiva was highest in animals that suffer from TR. This indicates that MCs and the complement system are involved in the destruction of the mineralized tissue in this condition. Full article
(This article belongs to the Special Issue Osteoclastogenesis and Osteogenesis)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Von Kossa staining of the gingiva in control cats (C) (<span class="html-italic">n</span> = 4) and cats affected by gingivitis (G) (<span class="html-italic">n</span> = 6), periodontitis (PD) (<span class="html-italic">n</span> = 8) or tooth resorption (TR) (<span class="html-italic">n</span> = 11). Images were obtained at 100× magnification. In control and gingivitis sections, no crystals were observed (calcification score 0). In periodontitis samples, rare, tiny crystals, marked by black arrows, were present (calcification score 1). In this TR section many large crystals could be seen (calcification score 3). (<b>B</b>) There was a significant increase in the occurrence of crystals in the tooth resorption (TR) (<span class="html-italic">n</span> = 11) group compared with the control (C) (<span class="html-italic">n</span> = 4) and gingivitis (G) (<span class="html-italic">n</span> = 6) groups (both <span class="html-italic">p</span> ≤ 0.05). There was no significant difference between the periodontitis group (PD) (<span class="html-italic">n</span> = 8) and the other groups.</p>
Full article ">Figure 2
<p>(<b>A</b>) Micro-Computed Tomography (μCT) image of a canine affected by tooth resorption (TR). The affected area is marked with a white rectangle. (<b>B</b>) Tartrate-resistant acid phosphatase staining of this tooth revealed many osteoclasts in this area, directly adjacent to the dentine of the tooth. The image was obtained at 400× magnification. (<b>C</b>) The evaluation showed significantly more osteoclasts in TR (<span class="html-italic">n</span> = 11) compared to periodontitis (PD) (<span class="html-italic">n</span> = 23) in both of the mineralized tissues, bone (<span class="html-italic">p</span> ≤ 0.01) and tooth (<span class="html-italic">p</span> ≤ 0.05). There was also a significant difference in the percentage of the eroded tooth surface between these groups (<span class="html-italic">p</span> ≤ 0.01).</p>
Full article ">Figure 3
<p>(<b>A</b>) Image of the immunohistochemical histamine staining in a TR section. The image was taken at 200× magnification. MCs could be identified. (<b>B</b>) Immunohistochemical staining for histamine showed a significant increase of MCs in the TR group (<span class="html-italic">n</span> = 11) compared with the control (<span class="html-italic">n</span> = 7) and gingivitis groups (<span class="html-italic">n</span> = 10) (both <span class="html-italic">p</span> ≤ 0.01). There was no difference between the TR and periodontitis groups (<span class="html-italic">n</span> = 8).</p>
Full article ">Figure 4
<p>(<b>A</b>) Immunohistochemical staining of the complement C5a receptor (C5aR). Images were taken at 400× magnification. The endothelial layer of the vessels in the gingiva of the controls (C) (<span class="html-italic">n</span> = 4) and cats affected by gingivitis (G) (<span class="html-italic">n</span> = 6), periodontitis (PD) (<span class="html-italic">n</span> = 7) or tooth resorption (TR) (<span class="html-italic">n</span> = 12) is marked by black arrows. In the controls and gingivitis animals, only a few endothelial layers were stained. The staining in general was less marked than in in the other two groups (PD, TR). (<b>B</b>) No significant differences were observed between the groups. However, there was a trend for a difference in periodontitis and TR compared to control and gingivitis regarding the endothelial layer. In addition, there was a trend for a difference in positively stained cells of the lamina propria between the control group and the other groups.</p>
Full article ">Figure 5
<p>(<b>A</b>) Immunohistochemical staining of C5a in the gingiva of the controls (C) (<span class="html-italic">n</span> = 4) and of cats affected by gingivitis (G) (<span class="html-italic">n</span> = 6), periodontitis (PD) (<span class="html-italic">n</span> = 9) or tooth resorption (TR) (<span class="html-italic">n</span> = 12). Images were obtained at 400× magnification. No positive cells were observed in the control section. In the gingivitis section, a single positively stained cell was seen, marked by a black arrow. In the other two groups, notably more cells could be seen. (<b>B</b>) There was a trend for an increase in positively stained cells in the PD and TR groups compared with the other two groups. (<b>C</b>) Double immunofluorescence staining in the gingiva of a tooth affected by TR. Scale bar 50 μm. Mast cell granules (Avidin) were stained red, while C5a was stained green. Mast cells positive for C5a were observed, marked by white arrows.</p>
Full article ">Figure 6
<p>(<b>A</b>) Immunohistochemical staining of complement factor C3 in the gingiva of the controls (C) (<span class="html-italic">n</span> = 4) and cats affected by gingivitis (G) (<span class="html-italic">n</span> = 6), periodontitis (PD) (<span class="html-italic">n</span> = 9) or tooth resorption (TR) (<span class="html-italic">n</span> = 12). Images were obtained at 400× magnification. In the control and gingivitis groups, no positively stained cells were observed. Two C3-positive cells were seen in periodontitis, marked by black arrows. The greatest number of positively stained cells could be observed in TR. (<b>B</b>) There was a statistically significant increase in C3-positive cells in the TR group compared with the other three groups (<span class="html-italic">p</span> ≤ 0.05). (<b>C</b>) Double immunofluorescence staining in the gingiva of a tooth affected by TR. Scale bar 50 μm. Mast cell granules (Avidin) were stained red, while C5 was stained green. In this section, one mast cell was positive for C3, marked by a white arrow.</p>
Full article ">Figure 7
<p>Hemolytic assay for the detection of complement classical pathway function. The optical density of the serum samples of both the control (C) and tooth resorption (TR) groups were plotted against the dilution factor. The curves show the control (4 cats) and TR groups (12 cats) with different levels of complement function. In TR, 50% lysis occurred between serum dilutions 1:80 and 1:160, in the control group between the undiluted serum and dilution 1:10. There was a significant difference in the serum complement concentration between the two groups, being significantly higher in the TR group than in the control group (<span class="html-italic">p</span> ≤ 0.05).</p>
Full article ">Figure 8
<p>Hemolytic assay for the detection of complement alternative pathway function. The optical density of the serum samples of both the control (C) and tooth resorption (TR) groups were plotted against the dilution factor. The curves show the control (4 cats) and TR groups (10 cats) with different levels of complement function. In both groups, 50% lysis occurred between serum dilutions 1:8 and 1:16. Nevertheless, there was a significant difference in the serum complement concentration between the two groups, being significantly higher in the TR group than in the control group (<span class="html-italic">p</span> ≤ 0.05).</p>
Full article ">
17 pages, 1125 KiB  
Review
Assessing Plasmin Generation in Health and Disease
by Adam Miszta, Dana Huskens, Demy Donkervoort, Molly J. M. Roberts, Alisa S. Wolberg and Bas de Laat
Int. J. Mol. Sci. 2021, 22(5), 2758; https://doi.org/10.3390/ijms22052758 - 9 Mar 2021
Cited by 28 | Viewed by 6953
Abstract
Fibrinolysis is an important process in hemostasis responsible for dissolving the clot during wound healing. Plasmin is a central enzyme in this process via its capacity to cleave fibrin. The kinetics of plasmin generation (PG) and inhibition during fibrinolysis have been poorly understood [...] Read more.
Fibrinolysis is an important process in hemostasis responsible for dissolving the clot during wound healing. Plasmin is a central enzyme in this process via its capacity to cleave fibrin. The kinetics of plasmin generation (PG) and inhibition during fibrinolysis have been poorly understood until the recent development of assays to quantify these metrics. The assessment of plasmin kinetics allows for the identification of fibrinolytic dysfunction and better understanding of the relationships between abnormal fibrin dissolution and disease pathogenesis. Additionally, direct measurement of the inhibition of PG by antifibrinolytic medications, such as tranexamic acid, can be a useful tool to assess the risks and effectiveness of antifibrinolytic therapy in hemorrhagic diseases. This review provides an overview of available PG assays to directly measure the kinetics of plasmin formation and inhibition in human and mouse plasmas and focuses on their applications in defining the role of plasmin in diseases, including angioedema, hemophilia, rare bleeding disorders, COVID-19, or diet-induced obesity. Moreover, this review introduces the PG assay as a promising clinical and research method to monitor antifibrinolytic medications and screen for genetic or acquired fibrinolytic disorders. Full article
(This article belongs to the Special Issue Fibrinogen/Fibrin, Factor XIII and Fibrinolysis in Diseases)
Show Figures

Figure 1

Figure 1
<p>Schematic representation of the primary structure of human glutamic (Glu)-plasminogen. The catalytic triad (His603, Asp646, and Ser741) within the protease domain, the activation site (Arg561–Val562), and the 24 disulfide bridges as well as the signal peptide are indicated. NTP, N-terminal peptide; K1–K5, kringles 1–5. Previously published [<a href="#B3-ijms-22-02758" class="html-bibr">3</a>] and adapted from Schaller and Gerber [<a href="#B4-ijms-22-02758" class="html-bibr">4</a>].</p>
Full article ">Figure 2
<p>A schematic representation of the fibrinolytic system. Fibrinolysis is initiated when the product of coagulation, thrombin, cleaves fibrinopeptides A and B from fibrinogen, leading to fibrin network formation. Thrombin activates factor XIII that crosslinks fibrin. Tissue plasminogen activator (tPA) or urokinase-type plasminogen activator (uPA) converts plasminogen into plasmin. Plasmin cleaves fibrin into fibrin degradation products (FDPs), including D-dimer. Direct plasmin inhibitors are α<sub>2</sub>-antiplasmin, α<sub>2</sub>-macroglobulin, and C-1 esterase inhibitor (C1-INH). Plasmin formation is also downregulated by several inhibitors, including plasminogen activator inhibitor 1 (PAI-1) which inhibits tPA and uPA. Interaction of thrombin with thrombomodulin (TM) activates thrombin-activated fibrinolysis inhibitor (TAFI). Activated TAFI (TAFIa) reduces plasminogen activation by cleaving C-lysines from fibrin and FDPs. Activation of plasmin can also be inhibited by lysine analogues tranexamic acid (TXA) and ε-aminocaproic acid (EACA).</p>
Full article ">
15 pages, 23641 KiB  
Review
Diabetes, Obesity, and Inflammation: Impact on Clinical and Radiographic Features of Breast Cancer
by Braden Miller, Hunter Chalfant, Alexandra Thomas, Elizabeth Wellberg, Christina Henson, Molly W. McNally, William E. Grizzle, Ajay Jain and Lacey R. McNally
Int. J. Mol. Sci. 2021, 22(5), 2757; https://doi.org/10.3390/ijms22052757 - 9 Mar 2021
Cited by 19 | Viewed by 5219
Abstract
Obesity, diabetes, and inflammation increase the risk of breast cancer, the most common malignancy in women. One of the mainstays of breast cancer treatment and improving outcomes is early detection through imaging-based screening. There may be a role for individualized imaging strategies for [...] Read more.
Obesity, diabetes, and inflammation increase the risk of breast cancer, the most common malignancy in women. One of the mainstays of breast cancer treatment and improving outcomes is early detection through imaging-based screening. There may be a role for individualized imaging strategies for patients with certain co-morbidities. Herein, we review the literature regarding the accuracy of conventional imaging modalities in obese and diabetic women, the potential role of anti-inflammatory agents to improve detection, and the novel molecular imaging techniques that may have a role for breast cancer screening in these patients. We demonstrate that with conventional imaging modalities, increased sensitivity often comes with a loss of specificity, resulting in unnecessary biopsies and overtreatment. Obese women have body size limitations that impair image quality, and diabetes increases the risk for dense breast tis-sue. Increased density is known to obscure the diagnosis of cancer on routine screening mammography. Novel molecu-lar imaging agents with targets such as estrogen receptor, human epidermal growth factor receptor 2 (HER2), pyrimi-dine analogues, and ligand-targeted receptor probes, among others, have potential to reduce false positive results. They can also improve detection rates with increased resolution and inform therapeutic decision making. These emerg-ing imaging techniques promise to improve breast cancer diagnosis in obese patients with diabetes who have dense breasts, but more work is needed to validate their clinical application. Full article
(This article belongs to the Special Issue Molecular Imaging in Diabetes, Obesity, and Infections 2.0)
Show Figures

Figure 1

Figure 1
<p>Imaging appearance of different breasts of increasing density using three types of imaging modalities. Digital mammography depicts how breasts appear more opaque as breast density increases. T1-weighted non-fat-suppressed MR images show increasing amount of fibroglandular tissue (FGT). T1-weighted fat-suppressed contrast-enhanced subtraction MR images show increasing amounts of background parenchymal enhancement (BPE), which reflects the vascularity of the fibroglandular tissue. (Adapted from [<a href="#B23-ijms-22-02757" class="html-bibr">23</a>]).</p>
Full article ">Figure 2
<p>Image of ultrasound placement for ultrasound-guided biopsy. (<b>A</b>) Image of ultrasound-guided biopsy for medial breast lesions. (<b>B</b>) In the prefire position, needle tip (white arrowhead) should be placed at the margin of most lesions (black arrow), with needle parallel or near parallel to chest wall. (<b>C</b>) Postfire images should be documented and should definitively outline the needle traversing the lesion (black arrow). The tip of the needle may lie beyond the lesion margin in some cases, depending on the size of the target (white arrowhead).</p>
Full article ">Figure 3
<p>Mammography images depicting different appearances of calcifications. (<b>A</b>) Classic appearance of breast arterial calcifications (BAC) (white arrow) that are easy to discriminate and associated with diabetes. (<b>B</b>) Groups of microcalcifications (yellow circles) in a woman with ductal carcinoma in situ (DCIS) is shown in its most typical appearance on digital breast tomosynthesis (DBT). (<b>C</b>) A patient with DCIS with mammographic findings showing segmental fine pleomorphic and fine-linear branching calcifications. (<b>D</b>) A skin calcification resembles microcalcifications (white arrow) and often requires alternative views to localize to the skin. Any or all of these findings can be seen in a single mammogram and can create a convoluted picture for radiologists. (Adapted from [<a href="#B65-ijms-22-02757" class="html-bibr">65</a>]).</p>
Full article ">Figure 4
<p>Patterns of HER2–PET/CT compared with FDG–PET/CT, Maximum intensity projection. Lesion uptake was considered concerning for malignant disease when visually higher than blood pool. (<b>A</b>) Entire tumor burden showed concerning tracer uptake; (<b>B</b>) Major portion of the tumor burden showed concerning tracer uptake; (<b>C</b>) A very small subset of the tumor burden showed concerning tracer uptake; (<b>D</b>) The entire visualized tumor burden lacked concerning tracer uptake. Adapted from [<a href="#B98-ijms-22-02757" class="html-bibr">98</a>].</p>
Full article ">Figure 5
<p>Osteopontin-probe identifies triple-negative breast cancer using multispectral optoacoustic tomography in a murine model. Serial slices are shown with * indicating positive osteopontin-probe uptake within the orthotopic triple negative breast tumor. Adapted from [<a href="#B103-ijms-22-02757" class="html-bibr">103</a>].</p>
Full article ">
33 pages, 727 KiB  
Review
Omics Approaches in Adipose Tissue and Skeletal Muscle Addressing the Role of Extracellular Matrix in Obesity and Metabolic Dysfunction
by Augusto Anguita-Ruiz, Mireia Bustos-Aibar, Julio Plaza-Díaz, Andrea Mendez-Gutierrez, Jesús Alcalá-Fdez, Concepción María Aguilera and Francisco Javier Ruiz-Ojeda
Int. J. Mol. Sci. 2021, 22(5), 2756; https://doi.org/10.3390/ijms22052756 - 9 Mar 2021
Cited by 17 | Viewed by 6133
Abstract
Extracellular matrix (ECM) remodeling plays important roles in both white adipose tissue (WAT) and the skeletal muscle (SM) metabolism. Excessive adipocyte hypertrophy causes fibrosis, inflammation, and metabolic dysfunction in adipose tissue, as well as impaired adipogenesis. Similarly, disturbed ECM remodeling in SM has [...] Read more.
Extracellular matrix (ECM) remodeling plays important roles in both white adipose tissue (WAT) and the skeletal muscle (SM) metabolism. Excessive adipocyte hypertrophy causes fibrosis, inflammation, and metabolic dysfunction in adipose tissue, as well as impaired adipogenesis. Similarly, disturbed ECM remodeling in SM has metabolic consequences such as decreased insulin sensitivity. Most of described ECM molecular alterations have been associated with DNA sequence variation, alterations in gene expression patterns, and epigenetic modifications. Among others, the most important epigenetic mechanism by which cells are able to modulate their gene expression is DNA methylation. Epigenome-Wide Association Studies (EWAS) have become a powerful approach to identify DNA methylation variation associated with biological traits in humans. Likewise, Genome-Wide Association Studies (GWAS) and gene expression microarrays have allowed the study of whole-genome genetics and transcriptomics patterns in obesity and metabolic diseases. The aim of this review is to explore the molecular basis of ECM in WAT and SM remodeling in obesity and the consequences of metabolic complications. For that purpose, we reviewed scientific literature including all omics approaches reporting genetic, epigenetic, and transcriptomic (GWAS, EWAS, and RNA-seq or cDNA arrays) ECM-related alterations in WAT and SM as associated with metabolic dysfunction and obesity. Full article
(This article belongs to the Special Issue Proteolysis of Extracellular Matrix in Human Disease)
Show Figures

Figure 1

Figure 1
<p>Venn diagram collecting the genes reported in omics approaches in adipose tissue. This figure shows the overlapping between genetics, epigenomics, and transcriptomic unveiling the role of ECM <span class="html-italic">loci</span> in obesity and metabolic comorbidities. <span class="html-italic">ADIPOQ</span>, adiponectin, C1Q, and collagen domain-containing; <span class="html-italic">CD36</span>, cluster of differentiation 36; <span class="html-italic">COL5A1</span>, collagen type V alpha 1 chain; <span class="html-italic">IL6</span>, interleukin-6; <span class="html-italic">PPARG</span>, peroxisome proliferator-activated receptor gamma; <span class="html-italic">SIRT1</span>, sirtuin-1; <span class="html-italic">TCF7L2</span>, transcription factor 7 like 2.</p>
Full article ">
16 pages, 2647 KiB  
Article
Epigenetic Changes Governing Scn5a Expression in Denervated Skeletal Muscle
by David Carreras, Rebecca Martinez-Moreno, Mel·lina Pinsach-Abuin, Manel M. Santafe, Pol Gomà, Ramon Brugada, Fabiana S. Scornik, Guillermo J. Pérez and Sara Pagans
Int. J. Mol. Sci. 2021, 22(5), 2755; https://doi.org/10.3390/ijms22052755 - 9 Mar 2021
Cited by 6 | Viewed by 3191
Abstract
The SCN5A gene encodes the α-subunit of the voltage-gated cardiac sodium channel (NaV1.5), a key player in cardiac action potential depolarization. Genetic variants in protein-coding regions of the human SCN5A have been largely associated with inherited cardiac arrhythmias. Increasing evidence also [...] Read more.
The SCN5A gene encodes the α-subunit of the voltage-gated cardiac sodium channel (NaV1.5), a key player in cardiac action potential depolarization. Genetic variants in protein-coding regions of the human SCN5A have been largely associated with inherited cardiac arrhythmias. Increasing evidence also suggests that aberrant expression of the SCN5A gene could increase susceptibility to arrhythmogenic diseases, but the mechanisms governing SCN5A expression are not yet well understood. To gain insights into the molecular basis of SCN5A gene regulation, we used rat gastrocnemius muscle four days following denervation, a process well known to stimulate Scn5a expression. Our results show that denervation of rat skeletal muscle induces the expression of the adult cardiac Scn5a isoform. RNA-seq experiments reveal that denervation leads to significant changes in the transcriptome, with Scn5a amongst the fifty top upregulated genes. Consistent with this increase in expression, ChIP-qPCR assays show enrichment of H3K27ac and H3K4me3 and binding of the transcription factor Gata4 near the Scn5a promoter region. Also, Gata4 mRNA levels are significantly induced upon denervation. Genome-wide analysis of H3K27ac by ChIP-seq suggest that a super enhancer recently described to regulate Scn5a in cardiac tissue is activated in response to denervation. Altogether, our experiments reveal that similar mechanisms regulate the expression of Scn5a in denervated muscle and cardiac tissue, suggesting a conserved pathway for SCN5A expression among striated muscles. Full article
(This article belongs to the Special Issue Epigenetic Mechanisms of Cardiac Disease)
Show Figures

Figure 1

Figure 1
<p>Comparison of exon 6 sequence between reference cardiac adult isoform (top sequence) and the sequence obtained from the denervated skeletal muscle (bottom sequence). The electropherogram corresponds to the product obtained for the denervated muscle.</p>
Full article ">Figure 2
<p>Differential RNA-seq analysis of denervated skeletal muscle compared to control, innervated muscle. (<b>A</b>) Pie Chart shows the percentage of genes with no significant expression changes (grey), significantly upregulated (orange) and significantly downregulated (blue). (<b>B</b>) Heatmap showing the 50 top upregulated and downregulated genes upon denervation. Each row represents the expression level of a particular gene in control (<span class="html-italic">n</span> = 4) and denervated samples (<span class="html-italic">n</span> = 4).</p>
Full article ">Figure 3
<p>Gene Ontology (GO) enrichment annotation analysis of differentially expressed (DE) genes. GO categorization of DE genes in molecular functions (MF), cellular components (CC), biological processes (BP), and Kyoto Encyclopedia of Genes and Genomes (KEGG) Pathway (shown as -log<sub>10</sub> <span class="html-italic">p</span>-value). Only the top ten significant upregulated and downregulated annotations are shown (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p><span class="html-italic">Scn5a</span> is the only gene upregulated within the topological associated domain (TAD)<b>.</b> Limits of the TAD in which <span class="html-italic">Scn5a</span> is included, and changes in the expression of genes located within the TAD are shown as log2FC.</p>
Full article ">Figure 5
<p>Enrichment of H3K27ac and H3K4me3 and binding of Gata4 near the <span class="html-italic">Scn5a</span> promoter in the denervated muscle. (<b>A</b>) Schematic view of the localization of the four primer pairs used by qPCR. Numbers indicate outermost left base pairs of amplicons. (<b>B</b>) Chromatin immunoprecipitation (ChIP) assays from control (white) and denervated (blue) samples using antibodies against H3K27ac, H3K4me3 and Gata4 and followed by qPCR. Results are shown as percentage of input (mean ± SE). H3K27ac (<span class="html-italic">n</span> = 6), H3K4me3 (<span class="html-italic">n</span> = 6) and Gata4 (<span class="html-italic">n</span> = 3). Significance was analyzed with <span class="html-italic">t</span>-test (<b>*</b> <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 6
<p>Denervation induces the activation of a super enhancer downstream of <span class="html-italic">Scn5a</span>. Integrative Genome Viewer (IGV) view of the H3K27ac distribution pattern in the <span class="html-italic">Scn5a</span> TAD and the reads of RNA-seq data in control and denervated gastrocnemius muscle samples. Expanded view of the super enhancer region located between <span class="html-italic">Exog</span> and <span class="html-italic">Scn5a</span> gene (bottom).</p>
Full article ">
Previous Issue
Next Issue
Back to TopTop