[go: up one dir, main page]

Next Issue
Volume 22, January-2
Previous Issue
Volume 21, December-2
 
 
ijms-logo

Journal Browser

Journal Browser

Int. J. Mol. Sci., Volume 22, Issue 1 (January-1 2021) – 472 articles

Cover Story (view full-size image): We successfully developed new a supramolecular polymer containing difunctional adenine-containing end groups that spontaneously self-assembles into nanosized spherical micelles in aqueous solution due to the presence of self-complementary hydrogen-bonded adenine interactions. Subsequently, a pro-photosensitizer and antitumor drug could be effectively encapsulated within supramolecular micelles, in order to achieve effective drug delivery combined with photo-chemotherapy. More importantly, pro-photosensitizer-loaded micelles exhibited significantly enhanced cellular uptake after visible light irradiation, and the light-triggered disassembly of micellar structures rapidly increased the production of reactive oxygen species within the cells, which thus accelerated apoptotic cell death. Thus, this newly developed photoresponsive nanocarrier could serve as a novel anticancer delivery system to effectively [...] Read more.
  • Issues are regarded as officially published after their release is announced to the table of contents alert mailing list.
  • You may sign up for e-mail alerts to receive table of contents of newly released issues.
  • PDF is the official format for papers published in both, html and pdf forms. To view the papers in pdf format, click on the "PDF Full-text" link, and use the free Adobe Reader to open them.
Order results
Result details
Section
Select all
Export citation of selected articles as:
19 pages, 3350 KiB  
Article
Tryptophan Derivatives by Saccharomyces cerevisiae EC1118: Evaluation, Optimization, and Production in a Soybean-Based Medium
by Michele Dei Cas, Ileana Vigentini, Sara Vitalini, Antonella Laganaro, Marcello Iriti, Rita Paroni and Roberto Foschino
Int. J. Mol. Sci. 2021, 22(1), 472; https://doi.org/10.3390/ijms22010472 - 5 Jan 2021
Cited by 10 | Viewed by 7920
Abstract
Given the pharmacological properti es and the potential role of kynurenic acid (KYNA) in human physiology and the pleiotropic activity of the neurohormone melatonin (MEL) involved in physiological and immunological functions and as regulator of antioxidant enzymes, this study aimed at evaluating the [...] Read more.
Given the pharmacological properti es and the potential role of kynurenic acid (KYNA) in human physiology and the pleiotropic activity of the neurohormone melatonin (MEL) involved in physiological and immunological functions and as regulator of antioxidant enzymes, this study aimed at evaluating the capability of Saccharomyces cerevisiae EC1118 to release tryptophan derivatives (dTRPs) from the kynurenine (KYN) and melatonin pathways. The setting up of the spectroscopic and chromatographic conditions for the quantification of the dTRPs in LC-MS/MS system, the optimization of dTRPs’ production in fermentative and whole-cell biotransformation approaches and the production of dTRPs in a soybean-based cultural medium naturally enriched in tryptophan, as a case of study, were included in the experimental plan. Variable amounts of dTRPs, with a prevalence of metabolites of the KYN pathway, were detected. The LC-MS/MS analysis showed that the compound synthesized at highest concentration is KYNA that reached 9.146 ± 0.585 mg/L in fermentation trials in a chemically defined medium at 400 mg/L TRP. Further experiments in a soybean-based medium confirm KYNA as the main dTRPs, whereas the other dTRPs reached very lower concentrations. While detectable quantities of melatonin were never observed, two MEL isomers were successfully measured in laboratory media. Full article
(This article belongs to the Special Issue Tryptophan in Nutrition and Health)
Show Figures

Figure 1

Figure 1
<p><span class="html-italic">S. cerevisiae</span> EC1118 growth curve in YNBT100 medium and TRP consumption.</p>
Full article ">Figure 2
<p>(<b>a</b>) Schematic view of the metabolites from the KYN pathway accumulated in YNBT100 medium and (<b>b</b>,<b>c</b>) kinetics production of the extracellular intermediates of the KYN pathway in <span class="html-italic">S. cerevisiae</span> EC1118. (<b>d</b>) Kinetics production of the extracellular intermediates of the MEL pathway in <span class="html-italic">S. cerevisiae</span> EC1118 in YNBT100.</p>
Full article ">Figure 3
<p>Melatonin isomers (MI) production by <span class="html-italic">S.cerevisiae</span> EC1118. Chromatogram in (<b>A</b>) shows the melatonin transition (<span class="html-italic">m/z</span> 233.2 &gt; 174.2) of a pure standard containing melatonin (MEL, rt 6.80) tryptophan ethyl ester (TEE, rt 6.11) both at a concentration of 2 ng/mL and internal standard (MEL OCD3, rt 6.84, <span class="html-italic">m/z</span> 236.2 &gt; 177.2, black). Chromatograms in (<b>B</b>) show the superimposition of melatonin trace (<span class="html-italic">m/z</span> 233.2 &gt; 174.2) from three different <span class="html-italic">S. cerevisiae</span> EC1118 fermentation supernatants at 24 h (blue), 48 h (red), and 72 h (green). In (<b>B</b>) is evident the absence of the melatonin peak that should be expected at the same retention time of its labeled analogue (rt 6.80, black). Among melatonin isomers in fermentation medium it was evidenced the presence of tryptophan ethyl ester (TEE, rt 6.06) and two isomers, named MI1 (rt 4.70) and MI2 (rt 5.6), with a not-elucidated structure.</p>
Full article ">Figure 4
<p><span class="html-italic">S. cerevisiae</span> EC1118 growth curve in YNBT400 medium and TRP consumption.</p>
Full article ">Figure 5
<p>TRP derivatives produced by <span class="html-italic">S. cerevisiae</span> EC1118 in YNBT400 medium. Error bars indicate the standard deviation.</p>
Full article ">Figure 6
<p>TRP derivatives produced by <span class="html-italic">S. cerevisiae</span> EC1118 in WCB experiments. Error bars indicate the standard deviation among the mean values of biological replicates.</p>
Full article ">Figure 7
<p>TRP derivatives produced by <span class="html-italic">S. cerevisiae</span> EC1118 in fermentation experiments in YNBSOY medium. Error bars indicate the standard deviation among the means of biological replicates. Letters indicate the grouping information using the Fisher LSD Method and 95% confidence.</p>
Full article ">
29 pages, 1129 KiB  
Review
More Than Just Simple Interaction between STIM and Orai Proteins: CRAC Channel Function Enabled by a Network of Interactions with Regulatory Proteins
by Sascha Berlansky, Christina Humer, Matthias Sallinger and Irene Frischauf
Int. J. Mol. Sci. 2021, 22(1), 471; https://doi.org/10.3390/ijms22010471 - 5 Jan 2021
Cited by 21 | Viewed by 5883
Abstract
The calcium-release-activated calcium (CRAC) channel, activated by the release of Ca2+ from the endoplasmic reticulum (ER), is critical for Ca2+ homeostasis and active signal transduction in a plethora of cell types. Spurred by the long-sought decryption of the molecular nature of [...] Read more.
The calcium-release-activated calcium (CRAC) channel, activated by the release of Ca2+ from the endoplasmic reticulum (ER), is critical for Ca2+ homeostasis and active signal transduction in a plethora of cell types. Spurred by the long-sought decryption of the molecular nature of the CRAC channel, considerable scientific effort has been devoted to gaining insights into functional and structural mechanisms underlying this signalling cascade. Key players in CRAC channel function are the Stromal interaction molecule 1 (STIM1) and Orai1. STIM1 proteins span through the membrane of the ER, are competent in sensing luminal Ca2+ concentration, and in turn, are responsible for relaying the signal of Ca2+ store-depletion to pore-forming Orai1 proteins in the plasma membrane. A direct interaction of STIM1 and Orai1 allows for the re-entry of Ca2+ from the extracellular space. Although much is already known about the structure, function, and interaction of STIM1 and Orai1, there is growing evidence that CRAC under physiological conditions is dependent on additional proteins to function properly. Several auxiliary proteins have been shown to regulate CRAC channel activity by means of direct interactions with STIM1 and/or Orai1, promoting or hindering Ca2+ influx in a mechanistically diverse manner. Various proteins have also been identified to exert a modulatory role on the CRAC signalling cascade although inherently lacking an affinity for both STIM1 and Orai1. Apart from ubiquitously expressed representatives, a subset of such regulatory mechanisms seems to allow for a cell-type-specific control of CRAC channel function, considering the rather restricted expression patterns of the specific proteins. Given the high functional and clinical relevance of both generic and cell-type-specific interacting networks, the following review shall provide a comprehensive summary of regulators of the multilayered CRAC channel signalling cascade. It also includes proteins expressed in a narrow spectrum of cells and tissues that are often disregarded in other reviews of similar topics. Full article
(This article belongs to the Special Issue STIMulating Ca2+ Homeostasis)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Identified positive regulators of SOCE. Known protein-localisation together with potential identified interaction sites between STIM1/Orai1 and regulatory-interacting proteins are represented in green boxes.</p>
Full article ">Figure 2
<p>Identified negative regulators of store-operated Ca<sup>2+</sup> entry (SOCE). Known protein-localisation together with potential identified interaction sites between STIM1/Orai1 and regulatory-interacting proteins are represented in red boxes.</p>
Full article ">
17 pages, 4463 KiB  
Article
Apigenin, a Partial Antagonist of the Estrogen Receptor (ER), Inhibits ER-Positive Breast Cancer Cell Proliferation through Akt/FOXM1 Signaling
by Thu Ha Pham, Yann Le Page, Frédéric Percevault, François Ferrière, Gilles Flouriot and Farzad Pakdel
Int. J. Mol. Sci. 2021, 22(1), 470; https://doi.org/10.3390/ijms22010470 - 5 Jan 2021
Cited by 27 | Viewed by 6157
Abstract
Approximately 80% of breast cancer (BC) cases express the estrogen receptor (ER), and 30–40% of these cases acquire resistance to endocrine therapies over time. Hyperactivation of Akt is one of the mechanisms by which endocrine resistance is acquired. Apigenin (Api), a flavone found [...] Read more.
Approximately 80% of breast cancer (BC) cases express the estrogen receptor (ER), and 30–40% of these cases acquire resistance to endocrine therapies over time. Hyperactivation of Akt is one of the mechanisms by which endocrine resistance is acquired. Apigenin (Api), a flavone found in several plant foods, has shown beneficial effects in cancer and chronic diseases. Here, we studied the therapeutic potential of Api in the treatment of ER-positive, endocrine therapy-resistant BC. To achieve this objective, we stably overexpressed the constitutively active form of the Akt protein in MCF-7 cells (named the MCF-7/Akt clone). The proliferation of MCF-7/Akt cells is partially independent of estradiol (E2) and exhibits an incomplete response to the anti-estrogen agent 4-hydroxytamoxifen, demonstrating the resistance of these cells to hormone therapy. Api exerts an antiproliferative effect on the MCF-7/Akt clone. Api inhibits the proliferative effect of E2 by inducing G2/M phase cell cycle arrest and apoptosis. Importantly, Api inhibits the Akt/FOXM1 signaling pathway by decreasing the expression of FOXM1, a key transcription factor involved in the cell cycle. Api also alters the expression of genes regulated by FOXM1, including cell cycle-related genes, particularly in the MCF-7/Akt clone. Together, our results strengthen the therapeutic potential of Api for the treatment of endocrine-resistant BC. Full article
(This article belongs to the Special Issue Antitumor Activities of Natural Compounds From Plants)
Show Figures

Figure 1

Figure 1
<p>Characterization of MCF-7 cell clones expressing wild-type Akt (MCF-7/Ctrl) and its constitutively active form (MCF-7/Akt). (<b>A</b>) Equal amounts of protein extracts from MCF-7/Ctrl cells and MCF-7/Akt cells treated with or without tetracycline (Tet) were analyzed by Western blotting with an antibody specific for Akt. Notably, myr-Akt1 has a larger molecular mass than wild-type Akt1. (<b>B</b>,<b>C</b>) Protein expression and subcellular localization of phospho-Bad (<b>B</b>) and phospho-mTOR (<b>C</b>) in MCF-7/Ctrl cells and MCF-7/Akt cells, as examined by immunofluorescence. (<b>D</b>) Equal amounts of protein extracts from MCF-7/Ctrl cells and MCF-7/Akt cells were analyzed by Western blotting with an antibody specific for FOXO3a (left, above). Protein expression and subcellular localization of FOXO3a in MCF-7/Ctrl cells and MCF-7/Akt cells examined by immunofluorescence (right). Quantification of the nuclear/cytoplasmic localization ratio of FOXO3a (left, below). Fluorescence intensity in the nucleus and cytoplasm of each cell was measured using ImageJ software. The median nuclear/cytoplasmic localization ratio of the control condition was used as a threshold to determine the percentage of cells below (black) and above (white) this threshold in each condition. (<b>E</b>) Gene expression of GATA3 (above) FOXA1 and (below) in MCF-7/Ctrl cells and MCF-7/Akt cells, as revealed by RT-PCR. Statistical analyses were performed with a Mann-Whitney test. * <span class="html-italic">p</span>-value &lt; 0.05, ** <span class="html-italic">p</span>-value &lt; 0.01 is considered to be significantly different from the MCF-7/Ctrl cell group.</p>
Full article ">Figure 2
<p>Effects of apigenin on the proliferation and cell cycle of MCF-7/Ctrl cells and MCF-7/Akt cells. (<b>A</b>) MCF-7/Ctrl cells and MCF-7/Akt cells were treated with solvent (0.1% (<span class="html-italic">v</span>/<span class="html-italic">v</span>) DMSO and 0.1% (<span class="html-italic">v</span>/<span class="html-italic">v</span>) ethanol) as the control (Cont), 1 nM estradiol (E2), or 10 µM apigenin (Api) alone or in combination with E2. (<b>B</b>) MCF-7/Ctrl cells and MCF-7/Akt cells were treated with solvent (Cont) or 1 µM 4-hydroxytamoxifen (4-OHT) alone or in combination with Api in the presence or absence of E2. Cell numbers were determined using the sulforhodamine B assay after 6 days of treatment. The experiment was conducted 3 times in quintuplicate. The results are expressed as the fold change in cell number compared with that in the control MCF-7/Ctrl cells and are presented as the mean ± SEM. For the cell cycle assays, MCF-7/Ctrl cells (<b>C</b>,<b>D</b>) and MCF-7/Akt cells (<b>E</b>,<b>F</b>) were treated with solvent (Cont), 4-OHT and Api alone or in combination in the presence or absence of E2 for approximately 24 h. The cell cycle was analyzed by flow cytometry after propidium iodide staining. The experiment was conducted 3 times. The results are expressed as the percentage of cells in each phase of the cell cycle and are presented as the mean ± SEM. (<b>G</b>) Equal amounts of protein extracts from MCF-7/Ctrl cells and MCF-7/Akt cells were analyzed by Western blotting with antibodies specific for cyclin B1 and β-actin. MCF-7/Ctrl cells and MCF-7/Akt cells were treated with solvent (Cont), E2 and Api alone or in combination with E2 for 24 h. For apoptosis analysis, MCF-7/Ctrl cells (<b>H</b>,<b>I</b>) and MCF-7/Akt cells (<b>J</b>,<b>K</b>) were treated with solvent (Cont), 4-OHT and Api alone, or in combination, in the presence or absence of E2 for approximately 24 h. Apoptotic cells were detected using terminal deoxynucleotidyl dUTP nick end labeling (TUNEL) assays. The experiment was conducted 3 times. The percentage of apoptotic cells was quantified and expressed as the fold change of the Cont-treated of the MCF-7/Ctrl clone. The results are presented as the mean ± SEM. Statistical analyses were performed with one-way ANOVA followed by Tukey’s post hoc test. * <span class="html-italic">p</span>-value &lt; 0.05, ** <span class="html-italic">p</span>-value &lt; 0.01, *** <span class="html-italic">p</span>-value &lt; 0.001 indicate significant differences from the control group in the same clone. # <span class="html-italic">p</span>-value &lt; 0.05, ## <span class="html-italic">p</span>-value &lt; 0.01, ### <span class="html-italic">p</span>-value &lt; 0.001 indicate significant differences from the E2 group in the same clone.</p>
Full article ">Figure 3
<p>Effect of apigenin on ER target genes. MCF-7/Ctrl cells and MCF-7/Akt cells were treated with 0.1% (<span class="html-italic">v</span>/<span class="html-italic">v</span>) DMSO and 0.1% (<span class="html-italic">v</span>/<span class="html-italic">v</span>) ethanol as the control (Cont), 1 nM estradiol (E2), or 10 µM apigenin (Api) alone or in combination with E2 for 24 h. The gene expression of estrogen receptor 1 (ERα) (<b>A</b>), progesterone receptor (PgR) (<b>B</b>), amphiregulin (AREG) (<b>C</b>) and chemokine (C-X-C motif) ligand 12 (CXCL12) (<b>D</b>) was assessed by real-time PCR and normalized to the expression of the housekeeping genes GAPDH and TBP. The experiment was conducted three times in triplicate. The results are expressed as the fold change in gene expression compared with that in the control of the MCF-7/Ctrl cells and are presented as the mean ± SEM. Statistical analyses were performed with one-way ANOVA followed by Tukey’s post hoc test. * <span class="html-italic">p</span>-value &lt; 0.05, ** <span class="html-italic">p</span>-value &lt; 0.01, *** <span class="html-italic">p</span>-value &lt; 0.001 indicate significant differences from the control group in the same clone. # <span class="html-italic">p</span>-value &lt; 0.05, ## <span class="html-italic">p</span>-value &lt; 0.01 indicate significant differences from the E2 group in the same clone.</p>
Full article ">Figure 4
<p>Effects of apigenin on the ER genomic activity. MCF-7/Ctrl cells and MCF-7/Akt cells were transiently transfected with an estrogen-responsive element luciferase reporter plasmid (ERE-Luc) (<b>A</b>), activator protein 1 luciferase reporter plasmid (AP1-Luc) (<b>B</b>) or specificity protein 1 luciferase reporter plasmid (SP1-Luc) (<b>C</b>) as well as a CMV-β-galactosidase plasmid as a control to assess transfection efficiency. Then, the cells were treated with 0.1% (<span class="html-italic">v</span>/<span class="html-italic">v</span>) DMSO and 0.1% (<span class="html-italic">v</span>/<span class="html-italic">v</span>) ethanol as the control (Cont), 1 nM estradiol (E2), or 10 µM apigenin (Api) alone or in combination with E2 for 24 h. The experiment was conducted 3 times in triplicate. The results are expressed as the fold change in luciferase activity compared with that in the control of the MCF-7/Ctrl cells and are presented as the mean ± SEM. Statistical analyses were performed with one-way ANOVA followed by Tukey’s post hoc test. * <span class="html-italic">p-</span>value &lt; 0.05, ** <span class="html-italic">p</span>-value &lt; 0.01, *** <span class="html-italic">p</span>-value &lt; 0.001 indicate significant differences from the control group in the same clone. # <span class="html-italic">p</span>-value &lt; 0.05, ## <span class="html-italic">p</span>-value &lt; 0.01, ### <span class="html-italic">p</span>-value &lt; 0.001 indicate significant differences from the E2 group in the same clone.</p>
Full article ">Figure 5
<p>Apigenin inhibits the expression of FOXM1 and modulates the expression of FOXM1-related cell cycle genes. MCF-7/Ctrl cells and MCF-7/Akt cells were treated with 0.1% (<span class="html-italic">v</span>/<span class="html-italic">v</span>) DMSO and 0.1% (<span class="html-italic">v</span>/<span class="html-italic">v</span>) ethanol as the control (Cont), 1 nM estradiol (E2), or 10 µM apigenin (Api) alone or in combination with E2 for 24 h. (<b>A</b>) Equal amounts of protein extracts from MCF-7/Ctrl cells and MCF-7/Akt cells were analyzed by Western blotting with antibodies specific to FOXM1 and β-actin. The gene expression of forkhead box M1 (FOXM1) (<b>B</b>), polo-like kinase 1 (PLK1) (<b>C</b>), cyclin B1 (CCNB1) (<b>D</b>), cell division cycle 25A (CDC25A) (<b>E</b>), centromere protein A (CENPA) (<b>F</b>) and cyclin-dependent kinase inhibitor 1A (CDKN1A/p21) (<b>G</b>) was assessed by real-time PCR and normalized to the expression of the housekeeping genes GAPDH and TBP. The experiment was conducted three times in triplicate. The results are expressed as the fold change in gene expression compared with that in the control of the MCF-7/Ctrl cells and are presented as the mean ± SEM. Statistical analyses were performed with one-way ANOVA followed by Tukey’s post hoc test. * <span class="html-italic">p</span>-value &lt; 0.05, ** <span class="html-italic">p</span>-value &lt; 0.01, *** <span class="html-italic">p</span>-value &lt; 0.001 indicate significant differences from the control group in the same clone. # <span class="html-italic">p</span>-value &lt; 0.05, ## <span class="html-italic">p</span>-value &lt; 0.01, ### <span class="html-italic">p</span>-value &lt; 0.001 indicate significant differences from the E2 group in the same clone.</p>
Full article ">
12 pages, 1189 KiB  
Review
Intracellular Ca2+ Signaling in Protozoan Parasites: An Overview with a Focus on Mitochondria
by Pedro H. Scarpelli, Mateus F. Pecenin and Celia R. S. Garcia
Int. J. Mol. Sci. 2021, 22(1), 469; https://doi.org/10.3390/ijms22010469 - 5 Jan 2021
Cited by 16 | Viewed by 4392
Abstract
Ca2+ signaling has been involved in controling critical cellular functions such as activation of proteases, cell death, and cell cycle control. The endoplasmatic reticulum plays a significant role in Ca2+ storage inside the cell, but mitochondria have long been recognized as [...] Read more.
Ca2+ signaling has been involved in controling critical cellular functions such as activation of proteases, cell death, and cell cycle control. The endoplasmatic reticulum plays a significant role in Ca2+ storage inside the cell, but mitochondria have long been recognized as a fundamental Ca2+ pool. Protozoan parasites such as Plasmodium falciparum, Toxoplasma gondii, and Trypanosoma cruzi display a Ca2+ signaling toolkit with similarities to higher eukaryotes, including the participation of mitochondria in Ca2+-dependent signaling events. This review summarizes the most recent knowledge in mitochondrial Ca2+ signaling in protozoan parasites, focusing on the mechanism involved in mitochondrial Ca2+ uptake by pathogenic protists. Full article
(This article belongs to the Special Issue Mitochondrial Calcium Signaling)
Show Figures

Figure 1

Figure 1
<p>Fluorescent microscopy <span class="html-italic">Plasmodium falciparum</span> ring, trophozoites, and schizonts stages using a mitochondrial green fluorescent protein (GFP) construction: the parasite nucleus was stained by HOECHST33342 (blue) and mitochondria with MitoTracker Red CMX Ros (red) to demonstrate co-localization with Mito-Emerald-GFP (green) (<b>A</b>). <span class="html-italic">P. chabaudi</span> parasites were stained with Fluo4 (cytoplasmatic calcium indicator) and Rhod2 (mitochondrial calcium indicator): effect of melatonin (MLT) and thapsigargin (Thg) on Ca<sup>2+</sup> (<b>B</b>); addition of thapsigargin (<b>C</b>); effect of melatonin and thapsigargin on Ca<sup>2+</sup> fluorescence in the presence of FCCP (<b>D</b>); addition of thapsigargin in the presence of FCCP (<b>E</b>). Traces represent fluorescence intensity ratio of Ca<sup>2+</sup> probes Rhod-2 AM-mitochondria (open circles) and Fluo-3 AM-cytosol (fill squares). The images were retrieved with the author’s consent from the following references: [<a href="#B84-ijms-22-00469" class="html-bibr">84</a>] (<b>A</b>) and [<a href="#B83-ijms-22-00469" class="html-bibr">83</a>] (<b>B</b>–<b>E</b>).</p>
Full article ">
14 pages, 14154 KiB  
Article
Distinct Responses of Arabidopsis Telomeres and Transposable Elements to Zebularine Exposure
by Klára Konečná, Pavla Polanská Sováková, Karin Anteková, Jiří Fajkus and Miloslava Fojtová
Int. J. Mol. Sci. 2021, 22(1), 468; https://doi.org/10.3390/ijms22010468 - 5 Jan 2021
Cited by 7 | Viewed by 3901
Abstract
Involvement of epigenetic mechanisms in the regulation of telomeres and transposable elements (TEs), genomic regions with the protective and potentially detrimental function, respectively, has been frequently studied. Here, we analyzed telomere lengths in Arabidopsis thaliana plants of Columbia, Landsberg erecta and Wassilevskija ecotypes [...] Read more.
Involvement of epigenetic mechanisms in the regulation of telomeres and transposable elements (TEs), genomic regions with the protective and potentially detrimental function, respectively, has been frequently studied. Here, we analyzed telomere lengths in Arabidopsis thaliana plants of Columbia, Landsberg erecta and Wassilevskija ecotypes exposed repeatedly to the hypomethylation drug zebularine during germination. Shorter telomeres were detected in plants growing from seedlings germinated in the presence of zebularine with a progression in telomeric phenotype across generations, relatively high inter-individual variability, and diverse responses among ecotypes. Interestingly, the extent of telomere shortening in zebularine Columbia and Wassilevskija plants corresponded to the transcriptional activation of TEs, suggesting a correlated response of these genomic elements to the zebularine treatment. Changes in lengths of telomeres and levels of TE transcripts in leaves were not always correlated with a hypomethylation of cytosines located in these regions, indicating a cytosine methylation-independent level of their regulation. These observations, including differences among ecotypes together with distinct dynamics of the reversal of the disruption of telomere homeostasis and TEs transcriptional activation, reflect a complex involvement of epigenetic processes in the regulation of crucial genomic regions. Our results further demonstrate the ability of plant cells to cope with these changes without a critical loss of the genome stability. Full article
Show Figures

Figure 1

Figure 1
<p>Experimental design. Seeds of <span class="html-italic">A. thaliana</span> of Columbia-0 (Col), Landsberg erecta (Ler), and Wassilevskija (Ws) ecotypes were germinated for 7 days on a control medium (O) or a medium supplemented with 250 µM zebularine (Z), and plants were cultivated in soil without the drug (first generation, 1G). Seeds from one parental plant (P) were germinated to limit the contribution of inter-individual variability in telomere lengths. Seeds from selected 1G plants were germinated either on a medium supplemented with zebularine (Z/Z) or on a control medium (Z/O and O/O), and plants were grown in the soil (second generation, 2G). The same protocol was followed once more to get the third generation (Z/Z/Z, Z/Z/O, and O/O/O, 3G) plants. This nomenclature is used in the following text and figures.</p>
Full article ">Figure 2
<p>Lengths of telomeres in <span class="html-italic">A. thaliana</span> plants grown from seeds exposed to the zebularine during germination. (<b>a</b>) Seeds of Columbia-0 (Col); (<b>b</b>) Landsberg erecta (Ler); (<b>c</b>) Wassilewskija (Ws) ecotypes were exposed to the 250 µM zebularine during germination for three generations; in parallel, progenies of zebularine plants were germinated on a control medium (for experimental design and nomenclature of samples, see <a href="#ijms-22-00468-f001" class="html-fig">Figure 1</a>). Leaves for analysis were collected from 8-week-old plants. DNA was extracted from 2 g of leaves, 5 µg of DNA was digested by MseI restriction endonuclease and analyzed by Southern hybridization using radioactively labeled telomeric probe. Each sample (line) corresponded to the leaf taken from one independently cultivated plant. P, parental plant; Z, first generation (1G) of plants grown from seedlings germinated in the presence of 250 µM zebularine; Z/Z, plants grown from seedlings germinated for two generations (2G) in the presence of 250 µM zebularine; Z/Z/Z, plants grown from seedlings germinated for three generations (3G) in the presence of 250 µM zebularine; Z/O, 2G plants grown from seedlings of 1G zebularine plants (Z) germinated on control medium; Z/Z/O, 3G plants grown from seedlings of 2G zebularine plants (Z/Z) germinated on a control medium. Black arrows indicate plants that were propagated to the subsequent generation. In 3G plants with distinct telomere phenotypes (S, short telomeres and L, long telomeres), the levels of transposable elements (TEs) transcripts, TE methylation and methylation of telomeric cytosines were analyzed (Figures 3–5). M, DNA fragment size marker; GeneRuler 1 kb DNA Ladder (Thermo Fisher Scientific). In the right panels, lengths of telomeres were presented using a box and whisker plot with the bottom part (black) and the top part (white) representing the lower and upper quartiles, respectively, separated by the median. The ends of whiskers represent the minimal and maximal telomere lengths reflecting the range of the hybridization signal.</p>
Full article ">Figure 3
<p>Relative levels of transcripts of selected TEs in seedlings of <span class="html-italic">A. thaliana</span> plants exposed repeatedly to 250 µM zebularine during germinations and in leaves of plants grown from these seedlings. RNA was extracted using the NucleoSpin RNA kit (Macherey-Nagel) from 100 mg of 7-day-old seedlings and leaves of 8-week-old plants. Relative levels of transcripts in Z/Z/Z 7-day-old seedlings (7ds) and leaves were compared to levels detected in control organs (7ds O/O/O, leaf O/O/O/, for experimental design and nomenclature of samples, see <a href="#ijms-22-00468-f001" class="html-fig">Figure 1</a>). S, plants with short telomeres and L, plants with long telomeres (see <a href="#ijms-22-00468-f002" class="html-fig">Figure 2</a>a). Signals of TE transcripts were related to that of the <span class="html-italic">ubiquitin10</span> reference gene, levels of TE transcripts in 7ds O/O/O samples were arbitrarily chosen as 1. Data were evaluated by two sample F-test and t-test. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001. Note the logarithmic scale at the y axis.</p>
Full article ">Figure 4
<p>SPM11 transposon methylation analyzed by bisulfite sequencing. Methylation of 44 cytosines located in an amplified 336 bp region (6 cytosines in the CG sequence context, 14 cytosines in the CHG, 24 cytosines in the CHH, H = A, T, C) was evaluated by the CyMATE software [<a href="#B27-ijms-22-00468" class="html-bibr">27</a>]. Five to ten clones were analyzed per sample; mean values of methylated cytosines and standard deviations are presented. O/O/O, leaf of the control Col plant; Z/Z/Z L, leaf of 3G plant with long telomeres grown from seedlings exposed to zebularine during germinations; Z/Z/Z S, leaf of 3G plant with short telomeres grown from seedlings exposed to zebularine during germinations (see <a href="#ijms-22-00468-f002" class="html-fig">Figure 2</a>a). For experimental design and nomenclature of samples, see <a href="#ijms-22-00468-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure 5
<p>Relative levels of methylated cytosines in telomeric repeats in the first and third generations of 7-day-old seedlings (7ds) of <span class="html-italic">A. thaliana</span> plants germinated on control medium or exposed to 250 µM zebularine, and in leaves of plants grown from these seedlings in soil. Mixed samples were analyzed; DNAs were isolated from 1 g of seedlings cultivated on three Petri dishes and 2 g of leaves of three plants grown from these seedlings, except of individual 3G zebularine plants with short and long telomeres (leaf Z/Z/Z S, leaf Z/Z/Z L). For bisulfite conversion, DNAs were mixed at an equimolar ratio. Bisulfite-modified DNAs were hybridized with the oligonucleotide probe reflecting fraction of telomeres with methylated cytosines (<b>a</b>), and the probe complementary to the G-strand of telomeres to determine loading (<b>b</b>). The same membrane was sequentially hybridized with both probes. PC, positive control, genomic DNA from Col leaves non-converted by sodium bisulfite; NC, negative control, pUC19 plasmid DNA. (<b>c</b>) Relative methylation of cytosines in telomeric repeats. Intensities of hybridization signals in (<b>a</b>,<b>b</b>) were evaluated by the MultiGauge software (FujiFilm) and expressed as the methylation/loading ratio. Signal ratios in respective control samples of the first generation (7ds O, leaf O) were arbitrarily taken as 1. For experimental design and nomenclature of samples, see <a href="#ijms-22-00468-f001" class="html-fig">Figure 1</a>.</p>
Full article ">
13 pages, 5397 KiB  
Article
AtPiezo Plays an Important Role in Root Cap Mechanotransduction
by Xianming Fang, Beibei Liu, Qianshuo Shao, Xuemei Huang, Jia Li, Sheng Luan and Kai He
Int. J. Mol. Sci. 2021, 22(1), 467; https://doi.org/10.3390/ijms22010467 - 5 Jan 2021
Cited by 23 | Viewed by 5375
Abstract
Plants encounter a variety of mechanical stimuli during their growth and development. It is currently believed that mechanosensitive ion channels play an essential role in the initial perception of mechanical force in plants. Over the past decade, the study of Piezo, a mechanosensitive [...] Read more.
Plants encounter a variety of mechanical stimuli during their growth and development. It is currently believed that mechanosensitive ion channels play an essential role in the initial perception of mechanical force in plants. Over the past decade, the study of Piezo, a mechanosensitive ion channel in animals, has made significant progress. It has been proved that the perception of mechanical force in various physiological processes of animals is indispensable. However, little is still known about the function of its homologs in plants. In this study, by investigating the function of the AtPiezo gene in the model plant Arabidopsis thaliana, we found that AtPiezo plays a role in the perception of mechanical force in plant root cap and the flow of Ca2+ is involved in this process. These findings allow us to understand the function of AtPiezo from the perspective of plants and provide new insights into the mechanism of plant root cap in response to mechanical stimuli. Full article
(This article belongs to the Section Molecular Plant Sciences)
Show Figures

Figure 1

Figure 1
<p>Evolutionary Analysis of Piezo in Plants. (<b>A</b>) Maximum likelihood unrooted phylogenetic tree of Piezo homologs. Green branches belong to the plant kingdom. (<b>B</b>) Predicted topology of the AtPiezo monomer. Beam and CTD domains are indicated in black. (<b>C</b>) The alignment of beam, inner helix and CTD domains of AtPiezo and its orthologs. Protein multiple sequence alignment of AtPiezo (<span class="html-italic">Arabidopsis thaliana</span>), GmPiezo1a (<span class="html-italic">Glycine max</span>), OsPiezo (<span class="html-italic">Oryza sativa Japonica Group</span>), ZmPiezo (<span class="html-italic">Zea mays</span>), PpPiezo1a (<span class="html-italic">Physcomitrella patens</span>), OlPiezo (<span class="html-italic">Ostreococcus lucimarinus</span>), MmPiezo1 (<span class="html-italic">Mus musculus</span>), and HmPiezo1 (<span class="html-italic">Homo sapiens</span>). The highly conserved amino acids among Piezo orthologues are highlighted in yellow. The consensus is with a threshold of &gt;50%.</p>
Full article ">Figure 2
<p>Expression patterns of <span class="html-italic">AtPiezo</span> in <span class="html-italic">Arabidopsis</span>. (<b>A</b>) Histochemical GUS staining is shown in 0.5-day-old, 2-day-old, and 10-day-old seedlings. Bars = 0.5 mm. (<b>B</b>) Longitudinal and cross sections of the root cap in different positions of <span class="html-italic">pAtPiezo::GUS</span> plant. Bars = 50 μm and 20 μm. (<b>C</b>) The expression of <span class="html-italic">pAtPiezo::NLS-YFP</span> in the root cap. Bars = 50 μm. (<b>D</b>) RT-PCR analysis of <span class="html-italic">AtPiezo</span> transcripts in different tissues.</p>
Full article ">Figure 3
<p><span class="html-italic">atpiezo</span> mutants exhibit reduced rooting ability. (<b>A</b>) Schematic diagram of <span class="html-italic">AtPiezo</span> gene structure. Exons are indicated by black boxes. The black lines represent introns. Untranslated regions are indicated by gray boxes. The insertion sites of two T-DNA insertion alleles are indicated by the black solid triangles. The allele of delete fragment is indicated by the black hollow triangle. Bar = 1000 bp. (<b>B</b>) Representative images of WT and <span class="html-italic">atpiezo</span> mutants grown on 1/2MS medium five days after germination, Bar = 1 cm. (<b>C</b>) Representative images of WT and <span class="html-italic">atpiezo</span> mutants grown in the soil 55 days after germination. Bar = 5 cm. (<b>D</b>) The rooting phenotype of Col-0, <span class="html-italic">piezo-1</span>, <span class="html-italic">piezo-2 and piezo-c1</span> horizontally grown on the agar medium. Five-day-old seedlings of different plant lines were grown on the medium containing 0.7%, 0.8%, or 0.9% agar. White solid arrows indicate the seedlings rooting in the medium. Bar = 1 cm. (<b>E</b>) The rate of WT and <span class="html-italic">atpiezo</span> mutants foraging into the medium five days after germination. Data are presented as mean ± SD (<span class="html-italic">n</span> = 6 plates, more than 30 seedlings in each plate) and analyzed with two-way ANOVA and Tukey’s multiple comparison test (Different lowercase letters indicate significant differences at <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>The <span class="html-italic">atpiezo</span> mutant plants show altered growth status inside the medium. (<b>A</b>) Representative images of growth status of primary roots of 7-day-old plants in the agar medium. White solid arrows indicate the helixes at primary roots. Black solid arrows indicate the positions where the root tips reach. Bar = 1 cm. (<b>B</b>) The rate of helical roots of WT and <span class="html-italic">atpiezo</span> mutants in the media with different agar concentrations. Data are presented as mean ± SD (<span class="html-italic">n</span> = 6 plates, more than 10 seedlings in each plate) and analyzed with two-way ANOVA and Tukey’s multiple comparison test (Different lowercase letters indicate significant differences at <span class="html-italic">p</span> &lt; 0.05). (<b>C</b>) The distance from the root tip to the medium surface of WT and <span class="html-italic">atpiezo</span> mutants in the media with different agar concentrations. Boxplots span the first to the third quartiles of the data. A line in the box represents the median (<span class="html-italic">n</span> &gt; 20). The results were analyzed with two-way ANOVA and Tukey’s multiple comparison test (Different lowercase letters indicate significant differences at <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>AtPiezo affects the shape of root caps in response to mechanical forces. (<b>A</b>) Schematic diagram of root cap index. The index is the ratio of the width across the QC and the length from QC to the root tip. (<b>B</b>) Root cap shape of Col-0, <span class="html-italic">piezo1</span>, and <span class="html-italic">piezo-2</span> grown on the medium with 0.8% agar. The root cap of 4-day-old plants that vertically grew on the agar medium (Vertical). The root cap of the plants that horizontally grew for 24 h followed by vertical growth for three days on the medium (Horizontal). Bar = 50 μm. (<b>C</b>) Statistic analysis of root cap index for different plants under vertical and horizontal growth conditions. Boxplots span the first to the third quartiles of the data. A line in the box represents the median (<span class="html-italic">n</span> = 30). The results were analyzed with two-way ANOVA and Tukey’s multiple comparison test (Different lowercase letters indicate significant differences at <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 6
<p>AtPiezo affects the Ca<sup>2+</sup> flux of plant root cap. (<b>A</b>) Schematic diagram of the position for Ca<sup>2+</sup> flux measurement using NMT. (<b>B</b>) The net Ca<sup>2+</sup> fluxes in Col-0, <span class="html-italic">piezo-1</span>, and <span class="html-italic">piezo-c1</span> seedlings that were grown on medium for five days. Boxplots span the first to the third quartiles of the data. A line in the box represents the median (<span class="html-italic">n</span> = 9). The results were analyzed with one-way ANOVA and Tukey’s multiple comparison test (Different lowercase letters indicate significant differences at <span class="html-italic">p</span> &lt; 0.05) (<b>C</b>) Real-time kinetics of Ca<sup>2+</sup> flow in the tip of root cap. Data are presented as mean ± SE (<span class="html-italic">n</span> = 5).</p>
Full article ">
20 pages, 343 KiB  
Review
Determination of Intraprostatic and Intratesticular Androgens
by Markéta Šimková, Jiří Heráček, Pavel Drašar and Richard Hampl
Int. J. Mol. Sci. 2021, 22(1), 466; https://doi.org/10.3390/ijms22010466 - 5 Jan 2021
Cited by 5 | Viewed by 3799
Abstract
Androgens represent the main hormones responsible for maintaining hormonal balance and function in the prostate and testis. As they are involved in prostate and testicular carcinogenesis, more detailed information of their active concentration at the site of action is required. Since the introduction [...] Read more.
Androgens represent the main hormones responsible for maintaining hormonal balance and function in the prostate and testis. As they are involved in prostate and testicular carcinogenesis, more detailed information of their active concentration at the site of action is required. Since the introduction of the term intracrinology as the local formation of active steroid hormones from inactive precursors of the adrenal gland, mainly dehydroepiandrosterone (DHEA) and DHEA-S, it is evident that blood circulating levels of sex steroid hormones need not reflect their actual concentrations in the tissue. Here, we review and critically evaluate available methods for the analysis of human intraprostatic and intratesticular steroid concentrations. Since analytical approaches have much in common in both tissues, we discuss them together. Preanalytical steps, including various techniques for separation of the analytes, are compared, followed by the end-point measurement. Advantages and disadvantages of chromatography-mass spectrometry (LC-MS, GC-MS), immunoanalytical methods (IA), and hybrid (LC-IA) are discussed. Finally, the clinical information value of the determined steroid hormones is evaluated concerning differentiating between patients with cancer or benign hyperplasia and between patients with different degrees of infertility. Adrenal-derived 11-oxygenated androgens are mentioned as perspective prognostic markers for these purposes. Full article
21 pages, 4503 KiB  
Article
Prion-Associated Neurodegeneration Causes Both Endoplasmic Reticulum Stress and Proteasome Impairment in a Murine Model of Spontaneous Disease
by Alicia Otero, Marina Betancor, Hasier Eraña, Natalia Fernández Borges, José J. Lucas, Juan José Badiola, Joaquín Castilla and Rosa Bolea
Int. J. Mol. Sci. 2021, 22(1), 465; https://doi.org/10.3390/ijms22010465 - 5 Jan 2021
Cited by 13 | Viewed by 4021
Abstract
Prion diseases are a group of neurodegenerative disorders that can be spontaneous, familial or acquired by infection. The conversion of the prion protein PrPC to its abnormal and misfolded isoform PrPSc is the main event in the pathogenesis of prion diseases [...] Read more.
Prion diseases are a group of neurodegenerative disorders that can be spontaneous, familial or acquired by infection. The conversion of the prion protein PrPC to its abnormal and misfolded isoform PrPSc is the main event in the pathogenesis of prion diseases of all origins. In spontaneous prion diseases, the mechanisms that trigger the formation of PrPSc in the central nervous system remain unknown. Several reports have demonstrated that the accumulation of PrPSc can induce endoplasmic reticulum (ER) stress and proteasome impairment from the early stages of the prion disease. Both mechanisms lead to an increment of PrP aggregates in the secretory pathway, which could explain the pathogenesis of spontaneous prion diseases. Here, we investigate the role of ER stress and proteasome impairment during prion disorders in a murine model of spontaneous prion disease (TgVole) co-expressing the UbG76V-GFP reporter, which allows measuring the proteasome activity in vivo. Spontaneously prion-affected mice showed a significantly higher accumulation of the PKR-like ER kinase (PERK), the ER chaperone binding immunoglobulin protein (BiP/Grp78), the ER protein disulfide isomerase (PDI) and the UbG76V-GFP reporter than age-matched controls in certain brain areas. The upregulation of PERK, BiP, PDI and ubiquitin was detected from the preclinical stage of the disease, indicating that ER stress and proteasome impairment begin at early stages of the spontaneous disease. Strong correlations were found between the deposition of these markers and neuropathological markers of prion disease in both preclinical and clinical mice. Our results suggest that both ER stress and proteasome impairment occur during the pathogenesis of spontaneous prion diseases. Full article
(This article belongs to the Section Molecular Neurobiology)
Show Figures

Figure 1

Figure 1
<p>PRKR-like endoplasmic reticulum kinase (PERK) upregulation in the thalamic area of clinical TgU1<sup>+</sup>/TgVole<sup>+</sup> mice. (<b>A</b>) Nuclear staining was observed for PERK in numerous cells of all groups of mice. The number of immunopositive cells was higher and immunostaining was stronger in TgU1<sup>+</sup>/TgVole<sup>+</sup> mice, especially in the thalamus and hypothalamus. Images correspond to the hypothalamic area in all mice. (<b>B</b>) PERK distribution in the brains of clinical and preclinical TgU1<sup>+</sup>/TgVole<sup>+</sup> mice and their age-matched controls. PERK immunostaining was analyzed using a semiquantitative scale from 0 (lack of immunostaining) to 5 (very intense immunostaining) in nine different brain areas: frontal cortex (Fc), septal area (Sa), cortex at the level of the thalamus (Tc), hippocampus (Hc), thalamus (T), hypothalamus (Ht), mesencephalon (Mes), cerebellum (Cbl) and medulla oblongata (Mo). The number of animals studied was the following: clinical TgU1<sup>+</sup>/TgVole<sup>+</sup> <span class="html-italic">n</span> = 6 (4 female, 2 male), clinical TgU1<sup>+</sup>/TgVole<sup>−</sup> controls <span class="html-italic">n</span> = 8 (6 female, 2 male), preclinical TgU1<sup>+</sup>/TgVole<sup>+</sup> <span class="html-italic">n</span> = 5 (2 female, 3 male) and preclinical TgU1<sup>+</sup>/TgVole<sup>−</sup> controls <span class="html-italic">n</span> = 5 (2 female, 3 male). Comparison of the PERK immunolabeling revealed significant differences between clinical TgU1<sup>+</sup>/TgVole<sup>+</sup> mice and clinical TgU1<sup>+</sup>/TgVole<sup>−</sup> controls and between preclinical TgU1<sup>+</sup>/TgVole<sup>+</sup> mice and preclinical TgU1<sup>+</sup>/TgVole<sup>−</sup> controls in different brain areas. (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, Mann-Whitney U test).</p>
Full article ">Figure 2
<p>Binding immunoglobulin protein (BiP) expression levels are higher in clinical and preclinical TgU1<sup>+</sup>/TgVole<sup>+</sup> mice than in healthy TgU1<sup>+</sup>/TgVole<sup>-</sup> controls in certain brain areas. (<b>A</b>) All groups of mice presented uniform cytoplasmic labeling of BiP in numerous neurons and a diffuse staining in the neuropil. Immunostaining intensity, rather than the number of positive cells, was higher in both clinical and preclinical TgU1<sup>+</sup>/TgVole<sup>+</sup> compared to their age-matched controls. (<b>B</b>) BiP distribution in the brains of clinical and preclinical TgU1<sup>+</sup>/TgVole<sup>+</sup> mice and their age-matched controls. BiP immunostaining was analyzed using a semiquantitative scale from 0 (lack of immunostaining) to 5 (very intense immunostaining) in nine different brain areas: frontal cortex (Fc), septal area (Sa), cortex at the level of the thalamus (Tc), hippocampus (Hc), thalamus (T), hypothalamus (Ht), mesencephalon (Mes), cerebellum (Cbl) and medulla oblongata (Mo). The number of animals included in each group was: clinical TgU1<sup>+</sup>/TgVole<sup>+</sup> <span class="html-italic">n</span> = 6 (4 female, 2 male), clinical TgU1<sup>+</sup>/TgVole<sup>−</sup> controls <span class="html-italic">n</span> = 8 (6 female, 2 male), preclinical TgU1<sup>+</sup>/TgVole<sup>+</sup> <span class="html-italic">n</span> = 5 (2 female, 3 male) and preclinical TgU1<sup>+</sup>/TgVole<sup>−</sup> controls <span class="html-italic">n</span> = 5 (2 female, 3 male). Comparison of the BiP immunolabeling revealed significant differences between clinical TgU1<sup>+</sup>/TgVole<sup>+</sup> mice and clinical TgU1<sup>+</sup>/TgVole<sup>−</sup> controls and between preclinical TgU1<sup>+</sup>/TgVole<sup>+</sup> mice and preclinical TgU1<sup>+</sup>/TgVole<sup>−</sup> controls in different brain areas. No differences were observed between clinical and preclinical TgU1<sup>+</sup>/TgVole<sup>+</sup> mice. (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, Mann-Whitney U test).</p>
Full article ">Figure 3
<p>Protein disulfide isomerase (PDI) accumulation is more intense in clinical TgU1<sup>+</sup>/TgVole<sup>+</sup> mice than in preclinical TgU1<sup>+</sup>/TgVole<sup>+</sup> mice and healthy TgU1<sup>+</sup>/TgVole<sup>−</sup> controls in all brain areas. (<b>A</b>) A strong intraneuronal PDI labeling was observed in the gigantocellular reticular nucleus of the medulla oblongata of clinical TgU1<sup>+</sup>/TgVole<sup>+</sup> mice. Intraneuronal PDI immunolabeling was also observed in the other groups of mice, but the intensity of the immunostaining and the number of immunopositive cells were reduced compared with clinical TgU1<sup>+</sup>/TgVole<sup>+</sup> mice (arrows). Insert picture contains two neurons showing a strong accumulation of PDI. (<b>B</b>) PDI distribution in the brains of clinical and preclinical TgU1<sup>+</sup>/TgVole<sup>+</sup> mice and their age-matched controls. PDI immunolabeling was semiquantitatively analyzed and scored on a scale of 0 (absence of immunolabeling) to 5 (very intense immunolabeling) in nine brain areas. Each group included the following number of animals: clinical TgU1<sup>+</sup>/TgVole<sup>+</sup> <span class="html-italic">n</span> = 6 (4 female, 2 male), clinical TgU1<sup>+</sup>/TgVole<sup>−</sup> controls <span class="html-italic">n</span> = 8 (6 female, 2 male), preclinical TgU1<sup>+</sup>/TgVole<sup>+</sup> <span class="html-italic">n</span> = 5 (2 female, 3 male) and preclinical TgU1<sup>+</sup>/TgVole<sup>−</sup> controls <span class="html-italic">n</span> = 5 (2 female, 3 male). Comparison of the PDI immunolabeling profiles revealed significant differences between the group of clinical TgU1<sup>+</sup>/TgVole<sup>+</sup> mice and clinical TgU1<sup>+</sup>/TgVole<sup>−</sup> controls in certain brain areas and between preclinical TgU1<sup>+</sup>/TgVole<sup>+</sup> mice and preclinical TgU1<sup>+</sup>/TgVole<sup>−</sup> controls. Significant differences in PDI immunostaining were also noticed between clinical and preclinical TgU1<sup>+</sup>/TgVole<sup>+</sup> mice and between clinical and preclinical TgU1<sup>+</sup>/TgVole<sup>−</sup> controls. (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p &lt;</span> 0.01, <span class="html-italic">*** p &lt;</span> 0.001, Mann-Whitney U test).</p>
Full article ">Figure 4
<p>Ub<sup>G76V</sup>-GFP accumulation is more intense in clinical and preclinical TgU1<sup>+</sup>/TgVole<sup>+</sup> mice than in healthy TgU1<sup>+</sup>/TgVole<sup>−</sup> controls in certain brain areas. (<b>A</b>) Strong Ub<sup>G76V</sup>-GFP immunostaining was observed in the thalamus of clinical and preclinical TgU1<sup>+</sup>/TgVole<sup>+</sup> mice, affecting cells whose morphology is compatible with reactive astrocytes. Healthy aged TgU1<sup>+</sup>/TgVole<sup>−</sup> mice (clinical TgU1<sup>+</sup>/TgVole<sup>−</sup> controls) showed granular immunostaining and filamentous ubiquitin aggregates in the neuropil (arrow). Healthy young TgU1<sup>+</sup>/TgVole<sup>−</sup> mice (preclinical TgU1<sup>+</sup>/TgVole<sup>−</sup> controls) showed slight intraneuronal Ub<sup>G76V</sup>-GFP immunolabeling (arrows). (<b>B</b>) Ub<sup>G76V</sup>-GFP distribution in the brains of clinical and preclinical TgU1<sup>+</sup>/TgVole<sup>+</sup> mice and their age-matched controls. Ub<sup>G76V</sup>-GFP immunolabeling was semiquantitatively analyzed and scored on a scale of 0 (absence of immunolabeling) to 5 (very intense immunolabeling) in nine different brain areas. The number of animals within each group was: clinical TgU1<sup>+</sup>/TgVole<sup>+</sup> <span class="html-italic">n</span> = 6 (4 female, 2 male), clinical TgU1<sup>+</sup>/TgVole<sup>−</sup> controls <span class="html-italic">n</span> = 8 (6 female, 2 male), preclinical TgU1<sup>+</sup>/TgVole<sup>+</sup> <span class="html-italic">n</span> = 5 (2 female, 3 male) and preclinical TgU1<sup>+</sup>/TgVole<sup>−</sup> controls <span class="html-italic">n</span> = 5 (2 female, 3 male). Comparison of the Ub<sup>G76V</sup>-GFP immunolabeling profiles revealed significant differences between the group of clinical TgU1<sup>+</sup>/TgVole<sup>+</sup> mice and clinical TgU1<sup>+</sup>/TgVole<sup>−</sup> controls and between preclinical TgU1<sup>+</sup>/TgVole<sup>+</sup> mice and preclinical TgU1<sup>+</sup>/TgVole<sup>−</sup> controls in numerous brain areas. No differences were observed between clinical and preclinical TgU1<sup>+</sup>/TgVole<sup>+</sup> mice. (* <span class="html-italic">p</span> &lt; 0.05, <span class="html-italic">*** p &lt;</span> 0.001, Mann-Whitney U test).</p>
Full article ">Figure 5
<p>Ub<sup>G76V</sup>-GFP intracellular accumulation is observed in brain areas showing prion-associated neuropathology. (<b>A</b>) Hippocampus from a clinical TgU1<sup>+</sup>/TgVole<sup>+</sup> mouse stained with hematoxylin and eosin and immunostained for Ub<sup>G76V</sup>-GFP and glial fibrillary acidic protein (GFAP). This animal shows severe spongiosis in the hippocampus and intense immunostaining for Ub<sup>G76V</sup>-GFP and GFAP. Strong Ub<sup>G76V</sup>-GFP immunolabeling is observed, affecting numerous cells that appear to be reactive astrocytes. (<b>B)</b> Dual immunofluorescence staining with anti-GFAP and anti-GFP antibodies revealed that numerous reactive astrocytes accumulated the Ub<sup>G76V</sup>-GFP reporter, suggesting that these cells can compensate for proteasome impairment and accumulate high amounts of ubiquitin conjugates before succumbing to the cytotoxic effect.</p>
Full article ">
14 pages, 979 KiB  
Review
Metabolome-Driven Regulation of Adenovirus-Induced Cell Death
by Anastasia Laevskaya, Anton Borovjagin, Peter S. Timashev, Maciej S. Lesniak and Ilya Ulasov
Int. J. Mol. Sci. 2021, 22(1), 464; https://doi.org/10.3390/ijms22010464 - 5 Jan 2021
Cited by 4 | Viewed by 4383
Abstract
A viral infection that involves virus invasion, protein synthesis, and virion assembly is typically accompanied by sharp fluctuations in the intracellular levels of metabolites. Under certain conditions, dramatic metabolic shifts can result in various types of cell death. Here, we review different types [...] Read more.
A viral infection that involves virus invasion, protein synthesis, and virion assembly is typically accompanied by sharp fluctuations in the intracellular levels of metabolites. Under certain conditions, dramatic metabolic shifts can result in various types of cell death. Here, we review different types of adenovirus-induced cell death associated with changes in metabolic profiles of the infected cells. As evidenced by experimental data, in most cases changes in the metabolome precede cell death rather than represent its consequence. In our previous study, the induction of autophagic cell death was observed following adenovirus-mediated lactate production, acetyl-CoA accumulation, and ATP release, while apoptosis was demonstrated to be modulated by alterations in acetate and asparagine metabolism. On the other hand, adenovirus-induced ROS production and ATP depletion were demonstrated to play a significant role in the process of necrotic cell death. Interestingly, the accumulation of ceramide compounds was found to contribute to the induction of all the three types of cell death mentioned above. Eventually, the characterization of metabolite analysis could help in uncovering the molecular mechanism of adenovirus-mediated cell death induction and contribute to the development of efficacious oncolytic adenoviral vectors. Full article
(This article belongs to the Special Issue Autophagy in Cancer Progression and Therapeutics)
Show Figures

Figure 1

Figure 1
<p>Schematic representation of the tentative cell death mechanism operating through a ceramide-mediated mitophagy. Adenoviral infection induces ceramide accumulation through de novo synthesis from palmitoyl CoA and serine [<a href="#B19-ijms-22-00464" class="html-bibr">19</a>]. Subsequently, ceramide mediates the LC3 binding to the mitochondrial membrane and promotes lethal mitophagy leading to autophagic cell death [<a href="#B25-ijms-22-00464" class="html-bibr">25</a>].</p>
Full article ">Figure 2
<p>Calcium influx contributes to the induction of necrotic cell death. Calcium influx is promoted by many factors, including endoplasmic reticulum stress, acidosis (due to lactate accumulation) [<a href="#B70-ijms-22-00464" class="html-bibr">70</a>,<a href="#B71-ijms-22-00464" class="html-bibr">71</a>], and ATP depletion (particularly due to assembly of virions and RIPK1-dependent inhibition of ANT in the mitochondrial membrane mediated by ceramide accumulation). Subsequently, an elevated concentration of calcium induces ROS production through increasing mitochondrial permeability [<a href="#B56-ijms-22-00464" class="html-bibr">56</a>,<a href="#B72-ijms-22-00464" class="html-bibr">72</a>,<a href="#B73-ijms-22-00464" class="html-bibr">73</a>]. A simultaneous decrease in ATP levels and the induced ROS production result in necrotic cell death. Abbreviations: ER—endoplasmic reticulum, ANT—adenosine nucleotide translocase, MCI—mitochondrial complex I, LDH—lactate dehydrogenase, “+”—increase effect.</p>
Full article ">Figure 3
<p>Metabolome fluctuations that trigger different types of cell death. Elevated acetate levels stimulating uptake of glucose and lactate along with asparagine deficiency can be linked to the induction of apoptosis. Calcium influx, production of reactive oxygen species (ROS), increased glucose uptake and lactate secretion as well as an elevation in acetyl-CoA concentration are all capable of triggering autophagic cell death. In addition, ROS production, calcium elevation and ATP depletion are known to induce necroptosis. Furthermore, accumulation of ceramide species is capable of triggering all three types of cell death described in this review. Up arrow means it increases and down arrow means it decreases.</p>
Full article ">
16 pages, 1694 KiB  
Review
The Association between Hepatic Encephalopathy and Diabetic Encephalopathy: The Brain-Liver Axis
by So Yeong Cheon and Juhyun Song
Int. J. Mol. Sci. 2021, 22(1), 463; https://doi.org/10.3390/ijms22010463 - 5 Jan 2021
Cited by 22 | Viewed by 7545
Abstract
Hepatic encephalopathy (HE) is one of the main consequences of liver disease and is observed in severe liver failure and cirrhosis. Recent studies have provided significant evidence that HE shows several neurological symptoms including depressive mood, cognitive dysfunction, impaired circadian rhythm, and attention [...] Read more.
Hepatic encephalopathy (HE) is one of the main consequences of liver disease and is observed in severe liver failure and cirrhosis. Recent studies have provided significant evidence that HE shows several neurological symptoms including depressive mood, cognitive dysfunction, impaired circadian rhythm, and attention deficits as well as motor disturbance. Liver disease is also a risk factor for the development of diabetes mellitus. Diabetic encephalopathy (DE) is characterized by cognitive dysfunction and motor impairment. Recent research investigated the relationship between metabolic changes and the pathogenesis of neurological disease, indicating the importance between metabolic organs and the brain. Given that a diverse number of metabolites and changes in the brain contribute to neurologic dysfunction, HE and DE are emerging types of neurologic disease. Here, we review significant evidence of the association between HE and DE, and summarise the common risk factors. This review may provide promising therapeutic information and help to design a future metabolic organ-related study in relation to HE and DE. Full article
(This article belongs to the Special Issue Oxidative Stress and Inflammation in Chronic Diseases)
Show Figures

Figure 1

Figure 1
<p>BBB breakdown and brain edema may aggravate brain dysfunction in HE and DE. Circulating factors can cause BBB breakdown, accompanied by astrocyte swelling and dysfunction, and glial activation. These can induce neuroinflammation, which is responsible for neuronal cell damage in HE and DE.</p>
Full article ">Figure 2
<p>The neurotransmitter imbalance in HE and DE. The changes in neurotransmitters, such as glutamate, GABA, serotonin, dopamine, and noradrenaline, may lead to motor dysfunction, learning and memory dysfunction, and emotional disturbance. The levels of dopamine and choline are decreased and the levels of glutamate, GABA, and norepinephrine are increased in HE and DE. However, Dopamine and serotonin levels are elevated in HE. Conversely, serotonin level is reduced in HE.</p>
Full article ">Figure 3
<p>Insulin resistance and impaired glucose metabolism may aggravate brain dysfunction in HE and DE. Peripheral insulin and glucose metabolism can influence brain homeostasis, and impaired glucose tolerance and insulin resistance can cause brain dysfunction in HE and DE.</p>
Full article ">
34 pages, 8154 KiB  
Article
The Thyroid Hormone Transporter Mct8 Restricts Cathepsin-Mediated Thyroglobulin Processing in Male Mice through Thyroid Auto-Regulatory Mechanisms That Encompass Autophagy
by Vaishnavi Venugopalan, Alaa Al-Hashimi, Maren Rehders, Janine Golchert, Vivien Reinecke, Georg Homuth, Uwe Völker, Mythili Manirajah, Adam Touzani, Jonas Weber, Matthew S. Bogyo, Francois Verrey, Eva K. Wirth, Ulrich Schweizer, Heike Heuer, Janine Kirstein and Klaudia Brix
Int. J. Mol. Sci. 2021, 22(1), 462; https://doi.org/10.3390/ijms22010462 - 5 Jan 2021
Cited by 8 | Viewed by 4594
Abstract
The thyroid gland is both a thyroid hormone (TH) generating as well as a TH responsive organ. It is hence crucial that cathepsin-mediated proteolytic cleavage of the precursor thyroglobulin is regulated and integrated with the subsequent export of TH into the blood circulation, [...] Read more.
The thyroid gland is both a thyroid hormone (TH) generating as well as a TH responsive organ. It is hence crucial that cathepsin-mediated proteolytic cleavage of the precursor thyroglobulin is regulated and integrated with the subsequent export of TH into the blood circulation, which is enabled by TH transporters such as monocarboxylate transporters Mct8 and Mct10. Previously, we showed that cathepsin K-deficient mice exhibit the phenomenon of functional compensation through cathepsin L upregulation, which is independent of the canonical hypothalamus-pituitary-thyroid axis, thus, due to auto-regulation. Since these animals also feature enhanced Mct8 expression, we aimed to understand if TH transporters are part of the thyroid auto-regulatory mechanisms. Therefore, we analyzed phenotypic differences in thyroid function arising from combined cathepsin K and TH transporter deficiencies, i.e., in Ctsk-/-/Mct10-/-, Ctsk-/-/Mct8-/y, and Ctsk-/-/Mct8-/y/Mct10-/-. Despite the impaired TH export, thyroglobulin degradation was enhanced in the mice lacking Mct8, particularly in the triple-deficient genotype, due to increased cathepsin amounts and enhanced cysteine peptidase activities, leading to ongoing thyroglobulin proteolysis for TH liberation, eventually causing self-thyrotoxic thyroid states. The increased cathepsin amounts were a consequence of autophagy-mediated lysosomal biogenesis that is possibly triggered due to the stress accompanying intrathyroidal TH accumulation, in particular in the Ctsk-/-/Mct8-/y/Mct10-/- animals. Collectively, our data points to the notion that the absence of cathepsin K and Mct8 leads to excessive thyroglobulin degradation and TH liberation in a non-classical pathway of thyroid auto-regulation. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Thyroglobulin storage status in combined cathepsin K and TH transporter deficiency. Thyroid cryo-sections were immunolabelled with antibodies against Tg (green), and the intraluminal Tg status was assessed in (<b>A</b>) WT, (<b>B</b>) <span class="html-italic">Ctsk</span><sup>-/-</sup>, (<b>C</b>) <span class="html-italic">Ctsk</span><sup>-/-</sup>/<span class="html-italic">Mct10</span><sup>-/-</sup>, (<b>D</b>) <span class="html-italic">Ctsk</span><sup>-/-</sup>/<span class="html-italic">Mct8</span><sup>-/y</sup>, and (<b>E</b>) <span class="html-italic">Ctsk</span><sup>-/-</sup>/<span class="html-italic">Mct8</span><sup>-/y</sup>/<span class="html-italic">Mct10</span><sup>-/-</sup> mice by confocal laser scanning microscopy. Single-channel fluorescence and corresponding phase contrast micrographs are depicted in the right panels as indicated. Tg staining in the lumen was either homogenous and rather faint, corresponding to tightly compacted Tg (asterisks), or appeared multilayered, referring to solubilized Tg (arrows), depending on the accessibility of intraluminal Tg for binding of the Tg-specific antibodies. Bar graphs indicate the proportion of follicles displaying compacted (<b>F</b>) and solubilized Tg (<b>G</b>) relative to the total number of investigated follicles, respectively, in the genotypes. Mice lacking both, cathepsin K and TH transporters, showed a decrease in the number of follicles with compacted Tg, and accordingly an increase in the number of follicles exhibiting Tg in multilayers. Animals analyzed: <span class="html-italic">n</span> = 3 per genotype, numbers of follicles analyzed: <span class="html-italic">n</span> = 1151, 688, 633, 743, and 852 for WT, <span class="html-italic">Ctsk</span><sup>-/-</sup>, <span class="html-italic">Ctsk</span><sup>-/-</sup>/<span class="html-italic">Mct10</span><sup>-/-</sup>, <span class="html-italic">Ctsk</span><sup>-/-</sup>/<span class="html-italic">Mct8</span><sup>-/y</sup>, and <span class="html-italic">Ctsk</span><sup>-/-</sup>/<span class="html-italic">Mct8</span><sup>-/y</sup>/<span class="html-italic">Mct10</span><sup>-/-</sup> mice, respectively. Nuclei were counter-stained with Draq5<sup>TM</sup> (red). Scale bars represent 50 µm. Data is depicted as means ± SD. Levels of significance are indicated as * for <span class="html-italic">p</span> &lt; 0.05 and *** for <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 2
<p>Thyroglobulin cross-linkage in combined cathepsin K and TH transporter deficiencies. Proteins isolated from whole thyroid tissue lysates of the indicated genotypes and WT controls were separated on 8–18% horizontal SDS-gels under non-reducing conditions, transferred onto nitrocellulose membrane, and immunoblotted using Tg-specific antibodies. A representative immunoblot is shown (<b>A</b>). The molecular mass markers are given in the left margin. Bands representing multimers, dimers, monomers, and fragments of Tg are indicated in the right margin. Bar charts (<b>B</b>–<b>E</b>) represent densitometry analyses of total Tg (<b>B</b>), Tg multimers (<b>C</b>), Tg dimers (<b>D</b>), Tg monomers/Tg dimers (<b>E</b>), and Tg fragments/Tg dimers (<b>F</b>) in the investigated genotypes as fold changes over WT controls. No genotypic differences in band intensities of Tg multimers or Tg dimers were observed, indicating that Tg cross-linkage most likely remained unaltered upon cathepsin K and/or TH transporter deficiencies. The ratio of band intensities of Tg monomers over Tg dimers (representing Tg solubilization), as well as Tg fragments over Tg dimers (representing Tg degradation) showed an increase in <span class="html-italic">Ctsk</span><sup>-/-</sup>/<span class="html-italic">Mct8</span><sup>-/y</sup> and <span class="html-italic">Ctsk</span><sup>-/-</sup>/<span class="html-italic">Mct8</span><sup>-/y</sup>/<span class="html-italic">Mct10</span><sup>-/-</sup> thyroid tissue, suggesting enhanced solubilization and altered proteolytic processing of Tg. Animals analyzed: <span class="html-italic">n</span> = 3–4 per genotype. Densitometry data was normalized to total Ponceau-stained protein per lane and is depicted as means ± SD. Levels of significance are indicated as ** for <span class="html-italic">p</span> &lt; 0.01 and *** for <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 3
<p>Protein glycosylation in combined cathepsin K and TH transporter deficiencies. Thyroid tissue sections from (<b>A</b>) WT, (<b>B</b>) <span class="html-italic">Ctsk</span><sup>-/-</sup>, (<b>C</b>) <span class="html-italic">Ctsk</span><sup>-/-</sup>/<span class="html-italic">Mct10</span><sup>-/-</sup>, (<b>D</b>) <span class="html-italic">Ctsk</span><sup>-/-</sup>/<span class="html-italic">Mct8</span><sup>-/y</sup>, and (<b>E</b>) <span class="html-italic">Ctsk</span><sup>-/-</sup>/<span class="html-italic">Mct8</span><sup>-/y</sup>/<span class="html-italic">Mct10</span><sup>-/-</sup> mice were stained with biotinylated lectin ConA and Alexa 546-conjugated streptavidin (green) to assess any difference in glycosylation states by confocal laser scanning microscopy. Merged, single-channel fluorescence, and corresponding phase contrast micrographs are displayed as indicated. The intensity of ConA staining was measured using a Cell Profiler pipeline and normalized to the numbers of cells (<b>F</b>). Mice lacking both cathepsin K and either or both TH transporters showed a significant decrease in the ConA signal. Animals analyzed: <span class="html-italic">n</span> = 3 per genotype with 8–10 micrographs quantified per animal, respectively. Nuclei were counter-stained with Draq5<sup>TM</sup> (red). Scale bars represent 50 µm. Data is depicted as fold changes over WT controls as means ± SD. Levels of significance are indicated as * for <span class="html-italic">p</span> &lt; 0.05 and *** for <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 4
<p>Gross thyroglobulin degradation states in thyroid glands of mice lacking cathepsin K and TH transporters. Whole thyroid tissue lysates of indicated genotypes and WT controls were separated on 8–18% horizontal SDS-gels under reducing conditions, transferred onto nitrocellulose membrane, and subsequently probed with anti-Tg antibodies. Shown is a representative immunoblot (<b>A</b>). The molecular mass markers are given in the left margin. Bands representing dimers, monomers, and fragments of Tg are indicated in the right margin of the immunoblot. Bar charts (<b>B</b>–<b>E</b>) represent densitometry analyses of total Tg (<b>B</b>), Tg dimers (<b>C</b>), Tg monomers (<b>D</b>), and Tg fragments (<b>E</b>) in the investigated genotypes as fold changes over WT controls. <span class="html-italic">Ctsk<sup>-</sup></span><sup>/-</sup>/<span class="html-italic">Mct8</span><sup>-/y</sup> and <span class="html-italic">Ctsk</span><sup>-/-</sup>/<span class="html-italic">Mct8</span><sup>-/y</sup>/<span class="html-italic">Mct10</span><sup>-/-</sup> mice showed a decrease in mono- and dimeric Tg amounts, while exhibiting an increase in the amounts of Tg fragments, thereby indicating enhanced Tg degradation. Animals analyzed: <span class="html-italic">n</span> = 3–4 per genotype. Densitometry data was normalized to total Ponceau-stained protein per lane and is depicted as means ± SD. Levels of significance are indicated as * for <span class="html-italic">p</span> &lt; 0.05, ** for <span class="html-italic">p</span> &lt; 0.01, and *** for <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 5
<p>Active cysteine peptidases in thyroid glands of mice lacking cathepsin K alone or in combination with TH transporter deficiencies. Whole thyroid tissue from WT, <span class="html-italic">Ctsk</span><sup>-/-</sup>, <span class="html-italic">Ctsk</span><sup>-/-</sup>/<span class="html-italic">Mct10</span><sup>-/-</sup>, <span class="html-italic">Ctsk</span><sup>-/-</sup>/<span class="html-italic">Mct8</span><sup>-/y</sup> and <span class="html-italic">Ctsk</span><sup>-/-</sup>/<span class="html-italic">Mct8</span><sup>-/y</sup>/<span class="html-italic">Mct10</span><sup>-/-</sup> mice was homogenized in lysis buffer containing 5 µM biotinylated activity-based probe DCG-04 before protein separation and staining of the blots with HRP-conjugated streptavidin to exclusively detect proteolytically active cysteine peptidases (<b>A</b>). Molecular mass markers are indicated in the left margin. Controls were conducted with WT tissue that was lysed without the addition of DCG-04, demonstrating streptavidin detection of some high molecular mass bands representing endogenous biotinylated proteins (<b>B</b>). However, these are most likely not cysteine peptidases ranging from 20–30 kDa in molecular mass. Therefore, the densities of the resulting bands below the 32-kDa molecular mass marker were normalized to total Ponceau-stained protein per lane, and are presented as fold changes over WT (<b>C</b>). It is important to note that the DCG-04 activity-based probe binds in equimolar ratio to active cysteine peptidases, only. The quantitation of signal intensities is therefore representative of proteolytic activity. The relative signal intensity of active cysteine peptidases was higher in thyroid lysates of <span class="html-italic">Ctsk</span><sup>-/-</sup>/<span class="html-italic">Mct8</span><sup>-/y</sup> and further enhanced in <span class="html-italic">Ctsk</span><sup>-/-</sup>/<span class="html-italic">Mct8</span><sup>-/y</sup>/<span class="html-italic">Mct10</span><sup>-/-</sup> mice in comparison to WT controls (<b>C</b>). Animals analyzed: <span class="html-italic">n</span> = 3–4 per genotype. Data is depicted as means ± SD. Levels of significance are indicated as * for <span class="html-italic">p</span> &lt; 0.05 and *** for <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 6
<p>Cystatin C and D levels in combined cathepsin K and TH transporter deficiencies. Thyroid cryo-sections from mice of the indicated genotypes and WT controls were stained with antibodies specific for cystatin C (<b>A</b>–<b>E</b>) or cystatin D (<b>G</b>–<b>K</b>) (green). Merged, single-channel fluorescence, and corresponding phase contrast micrographs are displayed as indicated. Micrographs show that cystatin C predominantly localized to the thyroid follicle lumen (<b>A</b>–<b>E</b>, asterisks) while cystatin D mainly localized to the apical pericellular space of thyrocytes (<b>G</b>–<b>K</b>, arrows) in all genotypes investigated. Intraluminal cystatin C (<b>D</b> and <b>E</b>, asterisks) and D (<b>J</b> and <b>K</b>, asterisks) signals appeared enhanced in mice lacking Mct8. The intensities of cystatin C and D staining were measured using a Cell Profiler pipeline and normalized to the numbers of cells (<b>F</b> and <b>L</b>, respectively). <span class="html-italic">Ctsk</span><sup>-/-</sup>/<span class="html-italic">Mct8</span><sup>-/y</sup>/<span class="html-italic">Mct10</span><sup>-/-</sup> mice showed a two-fold increase in thyroidal cystatin C and D signals when compared to WT controls. Animals analyzed: <span class="html-italic">n</span> = 3 per genotype with 7–12 micrographs quantified per animal, respectively. Nuclei were counter-stained with Draq5<sup>TM</sup> (red). Scale bars represent 50 µm. Data is depicted as fold changes over WT controls as means ± SD. Levels of significance are indicated as * for <span class="html-italic">p</span> &lt; 0.05 and ** for <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 7
<p>Maturation states of cathepsin B, D, and L remain unaltered in combined cathepsin K and TH transporter deficiency. Whole thyroid tissue lysates from WT, <span class="html-italic">Ctsk</span><sup>-/-</sup>, <span class="html-italic">Ctsk</span><sup>-/-</sup>/<span class="html-italic">Mct10</span><sup>-/-</sup>, <span class="html-italic">Ctsk</span><sup>-/-</sup>/<span class="html-italic">Mct8</span><sup>-/y</sup>, and <span class="html-italic">Ctsk</span><sup>-/-</sup>/<span class="html-italic">Mct8</span><sup>-/y</sup>/<span class="html-italic">Mct10</span><sup>-/-</sup> mice were immunoblotted using antibodies specific for cathepsins B, D, or L, as indicated (<b>A</b>–<b>C</b>, respectively, left panels). Molecular mass markers are displayed in the left margins. Bands representing proform (pro), single-chain (SC), and heavy-chain (HC) of two-chain forms are indicated in the right margin of the immunoblots. The density of the resulting bands was normalized to total Ponceau-stained protein per lane. Bar graphs represent amounts of total cathepsins (<b>A</b>–<b>C</b>, right panels) and corresponding processed forms (<b>A</b>–<b>C</b>, bottom panels), respectively, as fold changes over WT. The total amounts of all investigated cathepsins were predominantly increased in <span class="html-italic">Ctsk</span><sup>-/-</sup>/<span class="html-italic">Mct8</span><sup>-/y</sup>/<span class="html-italic">Mct10</span><sup>-/-</sup> mice when compared to WT controls. The maturation states of cathepsins B and L did not differ in any genotype. Animals analyzed: <span class="html-italic">n</span> = 3–4 per genotype. Data is depicted as means ± SD. Levels of significance are indicated as * for <span class="html-italic">p</span> &lt; 0.05, ** for <span class="html-italic">p</span> &lt; 0.01, and *** for <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 8
<p>Assessing the induction of lysosomal biogenesis in mice lacking cathepsin K and TH transporters. Thyroid tissue sections from (<b>A</b>) WT, (<b>B</b>) <span class="html-italic">Ctsk</span><sup>-/-</sup>, (<b>C</b>) <span class="html-italic">Ctsk</span><sup>-/-</sup>/<span class="html-italic">Mct10</span><sup>-/-</sup>, (<b>D</b>) <span class="html-italic">Ctsk</span><sup>-/-</sup>/<span class="html-italic">Mct8</span><sup>-/y</sup>, and (<b>E</b>) <span class="html-italic">Ctsk</span><sup>-/-</sup>/<span class="html-italic">Mct8</span><sup>-/y</sup>/<span class="html-italic">Mct10</span><sup>-/-</sup> mice were stained with antibodies against Lamp1 (green) and imaged by confocal laser scanning microscopy (<b>A</b>–<b>E</b>). As expected, Lamp1 signals were observed on vesicular membranes in all investigated genotypes. The intensity of Lamp1 staining (<b>F</b>) and the numbers of vesicles containing Lamp1 or cathepsins B, D or L (<b>G</b>–<b>J</b>, respectively) were determined using a Cell Profiler pipeline and normalized to the numbers of cells. Note that Lamp1 signals and the numbers of vesicles per cell were significantly increased in thyroid tissue of the triple-deficient genotype. Animals analyzed: <span class="html-italic">n</span> = 3 per genotype with 8–10 micrographs quantified per animal, respectively. Nuclei were counter-stained with Draq5<sup>TM</sup> (red). Scale bars represent 20 µm. Data is depicted as fold changes over WT controls as means ± SD. Levels of significance are indicated as * for <span class="html-italic">p</span> &lt; 0.05, ** for <span class="html-italic">p</span> &lt; 0.01, and *** for <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 9
<p>Proteome analysis of lysosomal constituents in thyroid tissue of <span class="html-italic">Ctsk</span><sup>-/-</sup>/<span class="html-italic">Mct8</span><sup>-/y</sup>/<span class="html-italic">Mct10</span><sup>-/-</sup> mice. Snap-frozen thyroid tissue from the triple-deficient murine model and WT controls were analyzed using LC-MS/MS. The bars depict fold changes of derived protein levels (means) of commonly known targets of lysosomal biogenesis and those that play a role in lysosomal function in comparison to WT controls (WT = 1, left panel). The <span class="html-italic">q</span>-values derived by Welch’s t-test indicated that the protein levels of 33 out of 39 lysosomal proteins were significantly increased in <span class="html-italic">Ctsk</span><sup>-/-</sup>/<span class="html-italic">Mct8</span><sup>-/y</sup>/<span class="html-italic">Mct10</span><sup>-/-</sup> vs. WT thyroid tissue (center panel). Bars and <span class="html-italic">q</span>-values for corresponding proteins that show significant differences are indicated in bold. Volcano plot of proteome data displaying the pattern of differential thyroid protein abundance for <span class="html-italic">Ctsk</span><sup>-/-</sup>/<span class="html-italic">Mct8</span><sup>-/y</sup>/<span class="html-italic">Mct10</span><sup>-/-</sup> mice relative to WT animals is shown (right panel). The <span class="html-italic">x</span>-axis indicates the log<sub>2</sub> of the protein ratios in the comparison and the <span class="html-italic">y</span>-axis indicates the negative decadic logarithm of the <span class="html-italic">p</span>-values. Significantly differentially abundant proteins (<span class="html-italic">p</span> ≤ 0.05, fold change ≥ |1.5|) are highlighted as red or blue points, indicating proteins present in increased and decreased amounts in <span class="html-italic">Ctsk</span><sup>-/-</sup>/<span class="html-italic">Mct8</span><sup>-/y</sup>/<span class="html-italic">Mct10</span><sup>-/-</sup> mice relative to WT animals, respectively. Lysosomal proteins are explicitly labelled.</p>
Full article ">Figure 10
<p>Autophagy in the thyroid glands of mice lacking cathepsin K and TH transporters. Mouse thyroid glands from the indicated genotypes were sectioned and stained with LC3-specific antibodies (<b>A</b>, green). Immunofluorescence analyses revealed diffuse staining patterns for LC3 in WT, <span class="html-italic">Ctsk</span><sup>-/-</sup>, and <span class="html-italic">Ctsk</span><sup>-/-</sup>/<span class="html-italic">Mct10</span><sup>-/-</sup>, while LC3 signals were punctate and rather vesicular (arrows) in <span class="html-italic">Ctsk</span><sup>-/-</sup>/<span class="html-italic">Mct8</span><sup>-/y</sup> and <span class="html-italic">Ctsk</span><sup>-/-</sup>/<span class="html-italic">Mct8</span><sup>-/y</sup>/<span class="html-italic">Mct10</span><sup>-/-</sup> thyroids (<b>A</b>). Nuclei were counter-stained with Draq5<sup>TM</sup> (<b>A</b>, red). Scale bars represent 50 µm. Thyroid tissue lysates were separated on horizontal SDS-gels and immunoblotted for LC3 (<b>B</b>, left panel) or p62 (<b>C</b>, left panel). LC3-II and p62 band intensities were determined by densitometry and normalized to total Ponceau-stained protein per lane (<b>B</b> and <b>C</b>, right panels, respectively). Densitometry analyses confirmed that autophagy was induced in <span class="html-italic">Ctsk</span><sup>-/-</sup>/<span class="html-italic">Mct8</span><sup>-/y</sup> and <span class="html-italic">Ctsk</span><sup>-/-</sup>/<span class="html-italic">Mct8</span><sup>-/y</sup>/<span class="html-italic">Mct10</span><sup>-/-</sup> mice. In both genotypes, the LC3-II signal was enhanced while the p62 signal was diminished in comparison to WT controls, corresponding to increased autophagosomal numbers and autophagic flux, respectively. Animals analyzed: <span class="html-italic">n</span> = 5–6 per genotype with 8–10 micrographs quantified per animal in A, respectively, and <span class="html-italic">n</span> = 3–4 per genotype in (<b>B</b>) and (<b>C</b>). Densitometry data is depicted as fold changes over WT as means ± SD. Levels of significance are indicated as * for <span class="html-italic">p</span> &lt; 0.05, ** for <span class="html-italic">p</span> &lt; 0.01, and *** for <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 11
<p>Intrathyroidal iodine levels in combined cathepsin K and TH transporter deficiency. Whole thyroid tissue lysates from WT, <span class="html-italic">Ctsk</span><sup>-/-</sup>, <span class="html-italic">Ctsk</span><sup>-/-</sup>/<span class="html-italic">Mct10</span><sup>-/-</sup>, <span class="html-italic">Ctsk</span><sup>-/-</sup>/<span class="html-italic">Mct8</span><sup>-/y</sup>, and <span class="html-italic">Ctsk</span><sup>-/-</sup>/<span class="html-italic">Mct8</span><sup>-/y</sup>/<span class="html-italic">Mct10</span><sup>-/-</sup> mice were normalized to 1 µg of total protein and the SK reaction was performed to determine intrathyroidal iodine concentrations. Displayed is a bar graph representing iodine concentrations in all genotypes as fold changes over WT controls (<b>A</b>). Intrathyroidal iodine concentrations were significantly enhanced in <span class="html-italic">Ctsk</span><sup>-/-</sup>/<span class="html-italic">Mct8</span><sup>-/y</sup>, and more prominently in <span class="html-italic">Ctsk</span><sup>-/-</sup>/<span class="html-italic">Mct8</span><sup>-/y</sup>/<span class="html-italic">Mct10</span><sup>-/-</sup> mice. Snap-frozen thyroid tissue from <span class="html-italic">Ctsk</span><sup>-/-</sup>/<span class="html-italic">Mct8</span><sup>-/y</sup>/<span class="html-italic">Mct10</span><sup>-/-</sup> mice was further subjected to Omics analyses to study differential expression of genes and respective levels of proteins involved in iodine supply through the blood circulation (CD31) or by iodide uptake (NIS) (<b>B</b>), neither of which showed significant differences when compared to WT controls, while NIS mRNA and protein showed a trend toward increased levels. (<b>C</b>) Thyroid tissue lysates were separated on horizontal SDS-gels and immunoblotted for NIS (<b>C</b>), and the band intensities were determined by densitometry and normalized to total Ponceau-stained protein per lane (<b>D</b>). Molecular mass markers are displayed in the left margins (<b>C</b>). Bands representing NIS are indicated in the right margin of the immunoblots (<b>C</b>). Immunoblot and densitometry analyses confirmed that NIS amounts were not altered in <span class="html-italic">Ctsk</span><sup>-/-</sup>/<span class="html-italic">Mct8</span><sup>-/y</sup>/<span class="html-italic">Mct10</span><sup>-/-</sup> mice in comparison to WT controls (<b>C</b> and <b>D</b>). Densitometry data is depicted as fold changes over WT (<b>D</b>). Animals analyzed: <span class="html-italic">n</span> = 3–5 per genotype. Data is depicted as means ± SD. Levels of significance are indicated as ** for <span class="html-italic">p</span> &lt; 0.01 and *** for <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 12
<p>Nuclear T3 amounts in thyroids lacking cathepsin K and TH transporters. Cryo-sections of thyroid glands from WT, <span class="html-italic">Ctsk</span><sup>-/-</sup>, <span class="html-italic">Ctsk</span><sup>-/-</sup>/<span class="html-italic">Mct10</span><sup>-/-</sup>, <span class="html-italic">Ctsk</span><sup>-/-</sup>/<span class="html-italic">Mct8</span><sup>-/y</sup>, and <span class="html-italic">Ctsk</span><sup>-/-</sup>/<span class="html-italic">Mct8</span><sup>-/y</sup>/<span class="html-italic">Mct10</span><sup>-/-</sup> mice were incubated with T3-specific antibodies (green) and analyzed by confocal laser scanning microscopy (A′–E′, respectively). T3 signals in the nuclei of thyrocytes (yellow) for all investigated genotypes are depicted upon desaturation of the T3-channel (<b>A</b>–<b>E</b>, respectively). Areas occupied by nuclear T3 over the total nuclear area was determined by a Cell Profiler-based pipeline representing an observer-unbiased approach, and the proportion of nuclear T3 signals are displayed as fold changes over WT controls (<b>F</b>). <span class="html-italic">Ctsk</span><sup>-/-</sup>/<span class="html-italic">Mct8</span><sup>-/y</sup> and <span class="html-italic">Ctsk</span><sup>-/-</sup>/<span class="html-italic">Mct8</span><sup>-/y</sup>/<span class="html-italic">Mct10</span><sup>-/-</sup> mice show enhanced accumulation of T3 within the nuclei of thyroid epithelial cells, indicating thyrotoxicity. Note that a decrease in nuclear T3 was observed in <span class="html-italic">Ctsk</span><sup>-/-</sup> thyroids which feature enhanced Mct8-mediated TH export. Nuclei were counter-stained with Draq5<sup>TM</sup> (red). The Mct8-deficient genotypes showed dead cells in follicle lumina with nuclear T3 (<b>D</b> and <b>E</b>, arrowheads). Scale bars represent 50 µm. Animals analyzed: <span class="html-italic">n</span> = 3 per genotype with 6–10 micrographs quantified per animal in A-E, respectively. Data is depicted as means ± SD. Levels of significance are indicated as * for <span class="html-italic">p</span> &lt; 0.05, ** for <span class="html-italic">p</span> &lt; 0.01, and *** for <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 13
<p>Automated image analysis using a Cell Profiler pipeline. Schematic diagram outlining the modules used for quantifying cathepsin amounts and numbers of cathepsin-positive vesicles within thyrocytes. As an example, a thyroid tissue section stained with cathepsin B antibody is shown. The input image contains three channels, namely, cathepsin (green), CMO stain for cell cytoplasm (red), and Draq5<sup>TM</sup> counter-stain for nuclei (blue). (1) The input image was split into individual channels and converted to gray scale output images OrigRed, OrigGreen, and OrigBlue, respectively. (2) The OrigRed image representative of the CMO cytoplasmic staining was converted to a binary image after applying a threshold to eliminate any unspecific signal in the lumen. (3) The OrigBlue image was used to identify the nuclei. (4) To obtain total cathepsin intensity, the signal from the OrigGreen image was measured. (5) OrigRed was used to mask OrigGreen to exclusively detect immuno-positive signals in the epithelium (‘IntraCellular_Cath’). (6) IntraCellular_Cath was used to identify vesicles which were counted. Nuclear counts (step 3) were used to normalize as indicated in illustrations of the respective analyses.</p>
Full article ">
24 pages, 2925 KiB  
Article
Intermittent Hypoxic Conditioning Rescues Cognition and Mitochondrial Bioenergetic Profile in the Triple Transgenic Mouse Model of Alzheimer’s Disease
by Sónia C. Correia, Nuno J. Machado, Marco G. Alves, Pedro F. Oliveira and Paula I. Moreira
Int. J. Mol. Sci. 2021, 22(1), 461; https://doi.org/10.3390/ijms22010461 - 5 Jan 2021
Cited by 14 | Viewed by 4999
Abstract
The lack of effective disease-modifying therapeutics to tackle Alzheimer’s disease (AD) is unsettling considering the actual prevalence of this devastating neurodegenerative disorder worldwide. Intermittent hypoxic conditioning (IHC) is a powerful non-pharmacological procedure known to enhance brain resilience. In this context, the aim of [...] Read more.
The lack of effective disease-modifying therapeutics to tackle Alzheimer’s disease (AD) is unsettling considering the actual prevalence of this devastating neurodegenerative disorder worldwide. Intermittent hypoxic conditioning (IHC) is a powerful non-pharmacological procedure known to enhance brain resilience. In this context, the aim of the present study was to investigate the potential long-term protective impact of IHC against AD-related phenotype, putting a special focus on cognition and mitochondrial bioenergetics and dynamics. For this purpose, six-month-old male triple transgenic AD mice (3×Tg-AD) were submitted to an IHC protocol for two weeks and the behavioral assessment was performed at 8.5 months of age, while the sacrifice of mice occurred at nine months of age and their brains were removed for the remaining analyses. Interestingly, IHC was able to prevent anxiety-like behavior and memory and learning deficits and significantly reduced brain cortical levels of amyloid-β (Aβ) in 3×Tg-AD mice. Concerning brain energy metabolism, IHC caused a significant increase in brain cortical levels of glucose and a robust improvement of the mitochondrial bioenergetic profile in 3×Tg-AD mice, as mirrored by the significant increase in mitochondrial membrane potential (ΔΨm) and respiratory control ratio (RCR). Notably, the improvement of mitochondrial bioenergetics seems to result from an adaptative coordination of the distinct but intertwined aspects of the mitochondrial quality control axis. Particularly, our results indicate that IHC favors mitochondrial fusion and promotes mitochondrial biogenesis and transport and mitophagy in the brain cortex of 3×Tg-AD mice. Lastly, IHC also induced a marked reduction in synaptosomal-associated protein 25 kDa (SNAP-25) levels and a significant increase in both glutamate and GABA levels in the brain cortex of 3×Tg-AD mice, suggesting a remodeling of the synaptic microenvironment. Overall, these results demonstrate the effectiveness of the IHC paradigm in forestalling the AD-related phenotype in the 3×Tg-AD mouse model, offering new insights to AD therapy and forcing a rethink concerning the potential value of non-pharmacological interventions in clinical practice. Full article
(This article belongs to the Section Biochemistry)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Effect of IHC on locomotor activity and anxiety-like behavior in the 3×Tg-AD mouse model during the open-field test. (<b>A</b>) Total distance travelled and (<b>B</b>) time spent in the center of the open-field arena during the open field test. Data are presented as box-and-whisker plots representing median and interquartile range (IQR), with minimum and maximum values of 5–9 animals from each experimental condition. Statistical significance: ** <span class="html-italic">p</span> &lt; 0.01 when compared with control WT mice.</p>
Full article ">Figure 2
<p>Effect of IHC on learning and spatial memory in the 3×Tg-AD mouse model during the MWM test. (<b>A</b>) Mean escape latencies to reach the hidden platform during the training trials, which were conducted for 4 consecutive days. (<b>B</b>) The percent time in the target quadrant during the probe trial performed 24 h after the last training trial. (<b>C</b>) The number of platform crossings during the probe trial. Data are presented as box-and-whisker plots representing median and IQR, with minimum and maximum values of 5–9 animals from each experimental condition. Statistical significance: * <span class="html-italic">p</span> &lt; 0.05 when compared with control WT mice.</p>
Full article ">Figure 3
<p>Effect of IHC on AD hallmarks in the brain cortex of 3×Tg-AD mouse model. (<b>A</b>) Using the antibody 6E10, the Aβ levels were detected by immuno-dot blot. (<b>B</b>) Representative Western blot image and densiometric analysis of tau levels detect with AT8 antibody which recognizes Ser202/Thr205 residues. β-actin was used as an internal loading control. Data are presented as box-and-whisker plots representing median and IQR, with minimum and maximum values of 5 animals from each experimental condition. Statistical significance: ** <span class="html-italic">p</span> &lt; 0.01 when compared with control WT mice; <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 when compared with 3×Tg-AD mice.</p>
Full article ">Figure 4
<p>Effect of IHC on energy metabolism in the brain cortex of 3×Tg-AD mouse model. The levels of the metabolites (<b>A</b>) glucose, (<b>B</b>) lactate, (<b>C</b>) alanine, and (<b>D</b>) succinate were detected in the brain cortex by NMR spectroscopy. (<b>E</b>) Citrate synthase activity was measured spectrophotometrically at 412 nm. Data are presented as box-and-whisker plots representing median and IQR, with minimum and maximum values of 5–7 animals from each experimental condition. Statistical significance: * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001 when compared with control WT mice; <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 when compared with 3×Tg-AD mice.</p>
Full article ">Figure 5
<p>Effect of IHC on mitochondrial fusion-fission machinery in the brain cortex of 3×Tg-AD mouse model. (<b>A</b>) Representative Western blot image and densiometric analysis of OPA-1, (<b>B</b>) Mnf-1, (<b>C</b>) Mnf-2, (<b>D</b>) DRP1 (<b>E</b>) Fis1 protein levels. β-actin was used as an internal loading control. Active form of DRP1 (p<sup>Ser616</sup>-DRP1) was normalized to total DRP1 levels. Data are presented as box-and-whisker plots representing median and IQR, with minimum and maximum values of 5 animals from each experimental condition. Statistical significance: *** <span class="html-italic">p</span> &lt; 0.001 when compared with control WT mice; <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05 when compared with 3×Tg-AD mice.</p>
Full article ">Figure 6
<p>Effect of IHC on mitochondrial biogenesis in the brain cortex of 3×Tg-AD mouse model. (<b>A</b>) Representative Western blot image and densiometric analysis of TFAM and (<b>B</b>) TOM20 protein levels. β-actin was used as an internal loading control. (<b>C</b>) The relative mtDNA copy number was determined by RT-PCR. Data are presented as box-and-whisker plots representing median and IQR, with minimum and maximum values of 5 animals from each experimental condition. Statistical significance: <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05, <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 when compared with 3×Tg-AD mice.</p>
Full article ">Figure 7
<p>Effect of IHC on mitophagy-related markers in the brain cortex of 3×Tg-AD mouse model. (<b>A</b>) Representative Western blot image and densiometric analysis of PINK-1, (<b>B</b>) Parkin, (<b>C</b>) LAMP-1 and (<b>D</b>) LC3-II protein levels. β-actin was used as an internal loading control. Data are presented as box-and-whisker plots representing median and IQR, with minimum and maximum values of 5 animals from each experimental condition. Statistical significance: ** <span class="html-italic">p</span> &lt; 0.01 when compared with control WT mice; <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05, <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 when compared with 3×Tg-AD mice.</p>
Full article ">Figure 8
<p>Effect of IHC on mitochondrial transport-related proteins in the brain cortex of 3×Tg-AD mouse model. (<b>A</b>) Representative Western blot image and densiometric analysis of dynein, (<b>B</b>) KIF-5B and (<b>C</b>) syntaphilin protein levels. β-actin was used as an internal loading control. Data are presented as box-and-whisker plots representing median and IQR, with minimum and maximum values of 5 animals from each experimental condition. Statistical significance: * <span class="html-italic">p</span> &lt; 0.05 when compared with control WT mice; <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05, <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 when compared with 3×Tg-AD mice.</p>
Full article ">Figure 9
<p>Effect of IHC on synaptic integrity markers and neurotransmitters levels in the brain cortex of 3×Tg-AD mouse model. (<b>A</b>) Representative Western blot image and densiometric analysis of SNAP-25, (<b>B</b>) synaptotagamin-I and (<b>C</b>) PSD-95 protein levels. β-actin was used as an internal loading control. The levels of the neurotransmitters (<b>D</b>) GABA and (<b>E</b>) glutamate were detected in the hippocampus by NMR spectroscopy. Data are presented as box-and-whisker plots representing median and IQR, with minimum and maximum values of 5 animals from each experimental condition. Statistical significance: * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 when compared with control WT mice; <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05, <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 when compared with 3×Tg-AD mice.</p>
Full article ">Figure 10
<p>Experimental design. 6-month-old male WT and 3×Tg-AD mice were randomly divided into two groups: (i) a conditioned and (ii) a non-conditioned control group. In the IHC paradigm, mice were exposed to 9 hypoxic episodes over 2 weeks for either 2 or 4 h at either 8% or 11% O<sub>2</sub> during a 2-week period. Open-field and MWM tests were performed around 8.5-month-old. Mice were euthanized at 9 months of age.</p>
Full article ">
19 pages, 4475 KiB  
Article
STAT3 Is an Upstream Regulator of Granzyme G in the Maternal-To-Zygotic Transition of Mouse Embryos
by Huan Ou-Yang, Shinn-Chih Wu, Li-Ying Sung, Shiao-Hsuan Yang, Shang-Hsun Yang, Kowit-Yu Chong and Chuan-Mu Chen
Int. J. Mol. Sci. 2021, 22(1), 460; https://doi.org/10.3390/ijms22010460 - 5 Jan 2021
Cited by 7 | Viewed by 3722
Abstract
The maternal-to-zygotic transition (MZT), which controls maternal signaling to synthesize zygotic gene products, promotes the preimplantation development of mouse zygotes to the two-cell stage. Our previous study reported that mouse granzyme g (Gzmg), a serine-type protease, is required for the MZT. In this [...] Read more.
The maternal-to-zygotic transition (MZT), which controls maternal signaling to synthesize zygotic gene products, promotes the preimplantation development of mouse zygotes to the two-cell stage. Our previous study reported that mouse granzyme g (Gzmg), a serine-type protease, is required for the MZT. In this study, we further identified the maternal factors that regulate the Gzmg promoter activity in the zygote to the two-cell stage of mouse embryos. A full-length Gzmg promoter from mouse genomic DNA, FL-pGzmg (−1696~+28 nt), was cloned, and four deletion constructs of this Gzmg promoter, Δ1-pGzmg (−1369~+28 nt), Δ2-pGzmg (−939~+28 nt), Δ3-pGzmg (−711~+28 nt) and Δ4-pGzmg (−417~+28 nt), were subsequently generated. Different-sized Gzmg promoters were used to perform promoter assays of mouse zygotes and two-cell stage embryos. The results showed that Δ4-pGzmg promoted the highest expression level of the enhanced green fluorescent protein (EGFP) reporter in the zygotes and two-cell embryos. The data suggested that time-specific transcription factors upregulated Gzmg by binding cis-elements in the −417~+28-nt Gzmg promoter region. According to the results of the promoter assay, the transcription factor binding sites were predicted and analyzed with the JASPAR database, and two transcription factors, signal transducer and activator of transcription 3 (STAT3) and GA-binding protein alpha (GABPα), were identified. Furthermore, STAT3 and GABPα are expressed and located in zygote pronuclei and two-cell nuclei were confirmed by immunofluorescence staining; however, only STAT3 was recruited to the mouse zygote pronuclei and two-cell nuclei injected with the Δ4-pGzmg reporter construct. These data indicated that STAT3 is a maternal transcription factor and may upregulate Gzmg to promote the MZT. Furthermore, treatment with a STAT3 inhibitor, S3I-201, caused mouse embryonic arrest at the zygote and two-cell stages. These results suggest that STAT3, a maternal protein, is a critical transcription factor and regulates Gzmg transcription activity in preimplantation mouse embryos. It plays an important role in the maternal-to-zygotic transition during early embryonic development. Full article
(This article belongs to the Special Issue Regulation of Gene Expression During Embryonic Development)
Show Figures

Figure 1

Figure 1
<p><span class="html-italic">Granzyme g</span> (<span class="html-italic">Gzmg</span>) promoter deletion constructs and the strategy of the promoter assay performed with a dual fluorescence co-injection system in mouse zygotes and 2-cell stage embryos. (<b>A</b>) The full-length <span class="html-italic">Gzmg</span> promoter (FL-<span class="html-italic">Gzmg</span>; −1724~+28 nt) was cloned from the mouse genome, and four deletion constructs of the <span class="html-italic">Gzmg</span> promoter were generated: Δ1-<span class="html-italic">Gzmg</span> (−1397~+28 ng), Δ2-<span class="html-italic">Gzmg</span> (−967~+28 nt), Δ3-<span class="html-italic">Gzmg</span> (−739~+28 nt) and Δ4-<span class="html-italic">Gzmg</span> (−445~+28 nt). The deletion constructs of the <span class="html-italic">Gzmg</span> promoter were inserted into a plasmid of enhanced green fluorescent protein-N1 (pEGFP-N1) reporter expression vector to replace the original cytomegalovirus (CMV) promoter in the plasmid. (<b>B</b>) Time-course experiments of dual fluorescence plasmids co-injected with p<span class="html-italic">Gzmg</span>-EGFP-N1 and mCherry control plasmid DNA in the zygote stage (20 h post-human chorionic gonadotropin (hCG)) and 2-cell stage (40 h post-hCG) embryos for the <span class="html-italic">Gzmg</span> promoter activity assay.</p>
Full article ">Figure 2
<p>Representative images and quantitative data from the Gzmg promoter assay obtained by a dual fluorescence co-injection system in mouse zygotes and 2-cell stage embryos. (<b>A</b>) Fluorescence images from the Gzmg promoter assay of the mouse zygote stage. (<b>B</b>) Fluorescence images from the Gzmg promoter assay of the mouse 2-cell stage. Light: Bright field used to show the normal morphology of the control mouse zygotes and 2-cell stage embryos. EGFP: green fluorescence field representing the expression levels of different promoter lengths of the pGzmg-EGFP-N1 deletion constructs. mCherry: red fluorescence field representing the expression levels of co-injected mCherry control plasmid DNA. (<b>C</b>) Quantitative data of the Gzmg promoter assay obtained by the dual fluorescence system. The graph shows the means ± SD of at least four replicates for each group of embryos. The asterisk (*) indicates zygote and 2-cell stage embryo groups with the same types of Gzmg promoter constructs that show significant differences according to a Student’s <span class="html-italic">t</span>-test (<span class="html-italic">p</span> &lt; 0.05). The number signs (<sup>#</sup> and <sup>##</sup>) indicate that the groups are significantly different from the Δ4-pGzmg construct groups of the same injected zygote or 2-cell stage embryos, as determined by one-way ANOVA (<sup>#</sup>, <span class="html-italic">p</span> &lt; 0.05 and <sup>##</sup>, <span class="html-italic">p</span> &lt; 0.01). The dagger (<sup>†</sup>) indicates a nonembryonic promoter construct from the mouse lung Clara cell-specific protein (ccsp) gene promoter, used as a negative control, that shows significant differences with all types of the Gzmg promoter constructs as determined by one-way ANOVA (<span class="html-italic">p</span> &lt; 0.05). RFUs: relative fluorescence intensity units.</p>
Full article ">Figure 3
<p>Immunofluorescence staining showing signal transducer and activator of transcription 3 (STAT3) and GA-binding protein alpha (GABPα) expression and localization in mouse zygotes and 2-cell stage embryos. Localization and expression of two transcription factors, STAT3 (<b>A</b>) and GABPα (<b>B</b>), in zygote stage embryos 20 h post-hCG and 2-cell stage embryos 40 h post-hCG. DAPI: 4′,6-diamidino-2-phenylindole, a blue-fluorescent dye that binds to AT-rich regions of double-stranded DNA and is used for nuclear localization staining. The abundance of green fluorescence representing the expression and localization of transcription factors STAT3 (zygote, <span class="html-italic">n</span> = 5 and 2-cell embryo, <span class="html-italic">n</span> = 6) and GABPα (zygote, <span class="html-italic">n</span> = 5 and 2-cell embryo, <span class="html-italic">n</span> = 6) in early-stage mouse embryos. Scale bar = 20 µm. Quantitative data of the fluorescence intensities of STAT3 (<b>C</b>) and GABPα (<b>D</b>) located in the zygotic pronuclei or 2-cell stage embryonic nuclei. The asterisk (*) indicates a significant difference between two pronuclei (*, <span class="html-italic">p</span> &lt; 0.05 and **, <span class="html-italic">p</span> &lt; 0.01) in each embryo as analyzed by Student’s <span class="html-italic">t</span>-test.</p>
Full article ">Figure 4
<p>STAT3 accumulates in the pronuclei and nuclei after microinjection of the Δ4-Gzmg promoter construct plasmid in mouse zygotes and 2-cell stage embryos. (<b>A</b>) Plot profile yellow lines represent the abundance and localization of transcription factors STAT3 (left panel) and GABPα (right panel), with or without Δ4-Gzmg promoter construct plasmid injection into one side of the zygotic pronuclei and 2-cell embryonic nuclei. Yellow arrowheads indicate nuclei after Δ4-Gzmg promoter construct plasmid injection. White arrowheads indicate nuclei without Δ4-Gzmg promoter construct plasmid injection. Scale bar = 10 µm. (<b>B</b>) Quantitative data were calculated by absolute difference values of fluorescence intensities of STAT3 (zygote, <span class="html-italic">n</span> = 5 and 2-cell embryo, <span class="html-italic">n</span> = 6) and GABPα (zygote, <span class="html-italic">n</span> = 6 and 2-cell embryo, <span class="html-italic">n</span> = 5) between the Δ4-Gzmg promoter plasmid-injected nucleus and the noninjected nucleus from the same zygote or 2-cell stage embryo. The asterisks (** and ***) indicate a significant difference as determined by Student’s <span class="html-italic">t</span>-test. (**, <span class="html-italic">p</span> &lt; 0.01 and ***, <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 5
<p>Abnormal chromosome plate in arrested zygote stage embryos treated with STAT3 inhibitor S3I-201. (<b>A</b>) Representative images of abnormal metaphase chromosome plates in arrested zygotes after treatment with 100-μM S3I-201 inhibitor in KSOM-AA embryo culture medium. Chromosomal DNA was stained with DAPI, a blue fluorescent dye. All samples were analyzed under a laser scanning confocal microscope. Scale bar = 20 µm. (<b>B</b>) Representative images of the different morphologies of the abnormal metaphase chromosome plates.</p>
Full article ">Figure 6
<p>A novel N-terminal domain (NTD)-deleted STAT3 in the pronucleus of mouse zygote stage embryos. (<b>A</b>) A graphic hypothesis showing that the N-terminal domain (NTD) deletion of STAT3 may be detected by two different recognition domains of STAT3 antibodies. One is labeled with an anti-F-2 antibody that specifically recognizes the N-terminal domain of STAT3 and is used for full-length STAT3α detection. The other is an anti-C-20 antibody that specifically recognizes the C-terminal domain of STAT3 and is used for the detection of both NTD-deleted STAT3 and STAT3α. (<b>B</b>) Immunolocalization of NTD-deleted STAT3 and full-length STAT3α in mouse zygote stage embryos by anti-F-2 (red fluorescent color) and anti-C-20 (green fluorescent color) antibodies. DNA was stained with DAPI (blue fluorescent color) to define pronuclei localization. Scale bar = 10 µm.</p>
Full article ">
21 pages, 3704 KiB  
Article
Characterization of Urine Stem Cell-Derived Extracellular Vesicles Reveals B Cell Stimulating Cargo
by Asmaa A. Zidan, Mohammed Al-Hawwas, Griffith B. Perkins, Ghada M. Mourad, Catherine J. M. Stapledon, Larisa Bobrovskaya, Xin-Fu Zhou and Plinio R. Hurtado
Int. J. Mol. Sci. 2021, 22(1), 459; https://doi.org/10.3390/ijms22010459 - 5 Jan 2021
Cited by 18 | Viewed by 4906
Abstract
Elucidation of the biological functions of extracellular vesicles (EVs) and their potential roles in physiological and pathological processes is an expanding field of research. In this study, we characterized USC–derived EVs and studied their capacity to modulate the human immune response in vitro. [...] Read more.
Elucidation of the biological functions of extracellular vesicles (EVs) and their potential roles in physiological and pathological processes is an expanding field of research. In this study, we characterized USC–derived EVs and studied their capacity to modulate the human immune response in vitro. We found that the USC–derived EVs are a heterogeneous population, ranging in size from that of micro–vesicles (150 nm–1 μm) down to that of exosomes (60–150 nm). Regarding their immunomodulatory functions, we found that upon isolation, the EVs (60–150 nm) induced B cell proliferation and IgM antibody secretion. Analysis of the EV contents unexpectedly revealed the presence of BAFF, APRIL, IL–6, and CD40L, all known to play a central role in B cell stimulation, differentiation, and humoral immunity. In regard to their effect on T cell functions, they resembled the function of mesenchymal stem cell (MSC)–derived EVs previously described, suppressing T cell response to activation. The finding that USC–derived EVs transport a potent bioactive cargo opens the door to a novel therapeutic avenue for boosting B cell responses in immunodeficiency or cancer. Full article
(This article belongs to the Special Issue Interaction of Nanomaterials with the Immune System)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Urine stem cells (USCs) characterization; (<b>A</b>) Phase–contrast light micrograph (LM) at passage zero showing stellate shaped cells with membrane projection reaching nearby cells (original magnification ×100, scale bar = 100 µm). (<b>B</b>) hematoxylin and eosin staining of USCs at passage three, showing basophilic stellate cells with perinuclear eosinophilia, large central nuclei, and cytoplasmic projections Actively dividing cells were occasionally seen represented by kissing cells (arrow) (original magnification ×400, scale bar = 20 µm). (<b>C</b>) Periodic acid–Schiff (PAS) staining of USCs at passage three, showing perinuclear magenta color (original magnification ×400, scale bar = 20 µm). (<b>D</b>) Toluidine Blue staining of USC pellet following chondrogenic differentiation showing blue staining proteoglycan–rich extracellular matrix (original magnification ×40, scale bar = 200 µm). (<b>E</b>) Oil Red O staining of the induced adipogenic cells showing red staining of intracellular lipid (original magnification ×400, scale bar = 20 µm). (<b>F</b>) Alizarin Red staining of induced osteogenic cells showing a dark orange discoloration of the calcification nodules (original magnification ×100, scale bar = 100 µm). (<b>G</b>) A representative sample of immune phenotypical characterization of 10<sup>5</sup> USCs (shaded curve) showing positive expression of CD105, CD73, and negative for HLA–DR, CD34, CD45, and CD14.</p>
Full article ">Figure 2
<p>Transmission electron microscopy (TEM) characterization of USCs, and extracellular vesicles EVs morphology and Biogenesis; (<b>A</b>) USCs TEM where different cell structures can be visualized such as the nucleus (N), abundant mitochondria (M), rough endoplasmic reticulum (rER), lysosomes (ly), pools of dense granules (DG) and multivesicular body (MVB); magnified in (<b>B</b>), where biogenesis of exosomes can be observed in the three phases described, endosomes (Endo), MVB containing intraluminal vesicles of different sizes (Asterix) and fusing of MVB to the membrane with exosome release (Mic. Mag. ×13,000). The cell membrane exhibits multiple projections, membrane budding forming vesicles (solid square) and micro–vesicles (large dotted square, illustrated in (<b>C</b>) containing dense granules varying in size, shape, and content, (Mic. Mag. ×2900). (<b>C</b>–<b>E</b>) The pictures show the heterogenicity in sizes and density, varying dense granules containing larger vesicles between 150 nm–1µm, typical for micro–vesicles (<b>C</b>–<b>E</b>), and homogenous less granular vesicles &lt; 150 nm, typical for exosomes (<b>E</b>,<b>F</b>), (Mic. Mag. ×2900). (<b>F</b>) Vesicle enclosed by USC membrane, mimic endocytosis (Mic. Mag. ×18,500) (Lead citrate/Uranyl acetate stain).</p>
Full article ">Figure 3
<p>Characterization of USCs–derived EVs; (<b>A</b>,<b>B</b>) TEM negative staining of the isolated EVs showing cup–shaped vesicles with an average size of 110 nm visualized at lower magnification (<b>A</b>; Mic. Mag. ×23,000) and higher magnification (<b>B</b>; Mic. Mag. ×30,000) (2%Uranyl Acetate). (<b>C</b>) Histogram presentation of 19 EVs pellets collected from 1.6 × 10<sup>7</sup> USCs analyzed for size and concentration, showing an average size of 122 nm and an average concentration of 1.90925 × 10<sup>11</sup>. (<b>D</b>) Size distribution curve of USCs isolated EVs particle concentration (×10<sup>7</sup>) vs. particle size mode, measured by nanoparticle tracking analysis (NTA) showing the average of the six technical replicate measurements for each exosome isolation by NanosightS300. (<b>E</b>) Western blot of USCs cell lysate and isolated the EVs for CD63, CD81, TSG101 antibodies as positive markers for EVs and Cytochrome C as mitochondrial membrane marker (cellular Marker) and negative marker for EVs. (<b>F</b>) Immune gold staining of the isolated vesicles for CD63 using 6 nm gold nanoparticles (2% Uranyl Acetate, scale bar = 100nm, Mic. Mag. ×30,000 (upper left) and ×23,000 rest).</p>
Full article ">Figure 4
<p>The effect of USCs EVs on B cells; (<b>A</b>) CD69 (early activation marker) expression on B cells population showing increased expression with EV co-culture. (<b>B</b>,<b>C</b>) Proliferation assay of the % proliferating B cells as the result of EVs co-culture in (<b>B</b>) resting and (<b>C</b>) CpGB–stimulated conditions showing significant enhancement of proliferation in both conditions in response to EV co-culture (<span class="html-italic">n</span> = 5). (<b>D</b>–<b>F</b>) Confocal microscopy images of DAPI (4′,6-diamidino-2-phenylindole) stained purified B cells (<b>D</b>) co-cultured with labeled EVs (<b>E</b>) for 24 h, (<b>F</b>) showing the presence of labeled EVs inside the cytoplasm and in aggregates attached to the cell surface (Mic. Magnification ×600, scale bar = 10 µm). (<b>G</b>) A representative sample of flow cytometry analysis of CD69<sup>+</sup> B cell population for FITC labeled EVs uptake versus control. (<b>H</b>) Antibody analysis of the supernatant showing a significant increase of IgM with EV co-culture (<span class="html-italic">n</span> = 6). Values are presented as Mean ± SEM, <span class="html-italic">p</span>-values were determined by Paired Student’s <span class="html-italic">t</span>–tests. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 5
<p>Cytokines Analysis of the EVs; (<b>A</b>) Cytokines analysis of EV co-culture supernatant run by Biolegend plex assay shows a significant increase in IL6, CD40L, IL10, and TNFα in response to EV co-culture (<span class="html-italic">n</span> = 3). (<b>B</b>) Heat map of the EV lysates, PBMCs, and the co-culture for cytokines concentration generated by Bio–legendplex software. (<b>C</b>) Cytokines analysis of 10 µg/mL EV lysates shows the expression of considerable amounts of IL–6, BAFF, APRIL, CD40L, and traces of IL–10, TNFα, and IFN–γ (<span class="html-italic">n</span> = 3). (<b>D</b>) Immune gold staining of the isolated vesicles for CD40L using 6 nm gold nanoparticles (Mic. Mag. ×23,000, 2% Uranyl Acetate). (<b>E</b>) Confocal microscope image of DAPI (blue) stained USC after permeabilization for CD40L mAb (red) (Mic. Mag. ×600, scale bar = 20 µm). (<b>F</b>) Confocal microscope image of DAPI (blue) stained USCs after staining with BAFFR mAb (green) (Mic. Mag. ×600, scale bar = 50 µm). (<b>G</b>) Flow cytometry analysis of 10<sup>4</sup> USCs for BAFFR expression showing 99% positive cells (green shaded histogram) in relation to the isotype control (non–shaded histogram), representative sample. Values are presented as mean ± SEM and p–values were determined by paired Student’s <span class="html-italic">t</span>–tests. Cytokines data are prior to log treatment. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 6
<p>USCs EVs effect on T cells in response to anti–CD3/CD28 stimulation; (<b>A</b>) Effect of the EVs on the proliferation of T cells in the presence of anti–CD3/CD28–bead stimulation showing significant suppression of the T cell proliferation (represented by % proliferating cells, <span class="html-italic">n</span> = 5). Values are presented as mean ± SEM, and <span class="html-italic">p</span>-values were determined by paired <span class="html-italic">t</span>–test analysis. ** <span class="html-italic">p</span> &lt; 0.01. (<b>B</b>–<b>D</b>) A representative sample of the viability assay of CD3<sup>+</sup> (T cell) population where (<b>B</b>) control T cells viability, (<b>C</b>) viability in response to EVs co-culture. (<b>D</b>) Viability in response to EV co-culture in the presence of anti–CD3/CD28 beads. Note the decrease in the viable population in the activated T cells co-cultured with EVs.</p>
Full article ">
13 pages, 3227 KiB  
Article
Increasing Oxygen Partial Pressures Induce a Distinct Transcriptional Response in Human PBMC: A Pilot Study on the “Normobaric Oxygen Paradox”
by Deborah Fratantonio, Fabio Virgili, Alessandro Zucchi, Kate Lambrechts, Tiziana Latronico, Pierre Lafère, Peter Germonpré and Costantino Balestra
Int. J. Mol. Sci. 2021, 22(1), 458; https://doi.org/10.3390/ijms22010458 - 5 Jan 2021
Cited by 44 | Viewed by 4457
Abstract
The term “normobaric oxygen paradox” (NOP), describes the response to the return to normoxia after a hyperoxic event, sensed by tissues as oxygen shortage, and resulting in up-regulation of the Hypoxia-inducible factor 1α (HIF-1α) transcription factor activity. The molecular characteristics of this response [...] Read more.
The term “normobaric oxygen paradox” (NOP), describes the response to the return to normoxia after a hyperoxic event, sensed by tissues as oxygen shortage, and resulting in up-regulation of the Hypoxia-inducible factor 1α (HIF-1α) transcription factor activity. The molecular characteristics of this response have not been yet fully characterized. Herein, we report the activation time trend of oxygen-sensitive transcription factors in human peripheral blood mononuclear cells (PBMCs) obtained from healthy subjects after one hour of exposure to mild (MH), high (HH) and very high (VHH) hyperoxia, corresponding to 30%, 100%, 140% O2, respectively. Our observations confirm that MH is perceived as a hypoxic stress, characterized by the activation of HIF-1α and Nuclear factor (erythroid-derived 2)-like 2 (NRF2), but not Nuclear Factor kappa-light-chain-enhancer of activated B cells (NF-κB). Conversely, HH is associated to a progressive loss of NOP response and to an increase in oxidative stress leading to NRF2 and NF-kB activation, accompanied by the synthesis of glutathione (GSH). After VHH, HIF-1α activation is totally absent and oxidative stress response, accompanied by NF-κB activation, is prevalent. Intracellular GSH and Matrix metallopeptidase 9 (MMP-9) plasma levels parallel the transcription factors activation pattern and remain elevated throughout the observation time. In conclusion, our study confirms that, in vivo, the return to normoxia after MH is sensed as a hypoxic trigger characterized by HIF-1α activation. On the contrary, HH and VHH induce a shift toward an oxidative stress response, characterized by NRF2 and NF-κB activation in the first 24 h post exposure. Full article
(This article belongs to the Special Issue Cellular Oxygen Homeostasis)
Show Figures

Figure 1

Figure 1
<p>HIF-1α nuclear translocation following 1 h hyperoxia. (<b>a</b>) Mild hyperoxia (30% O<sub>2</sub>); (<b>b</b>) high hyperoxia (100% O<sub>2</sub>); (<b>c</b>) very high hyperoxia (140% O<sub>2</sub>) before and after the recovery to normoxic conditions. Above histograms, the picture shows a representative western blot analysis. Results are expressed as fold change (mean ± SEM) in comparison to baseline (0), which was set at 1. * <span class="html-italic">p</span> &lt; 0.05, <span class="html-italic">** p</span> &lt; 0,01; for one-way ANOVA followed by Dunnett’s post hoc test.</p>
Full article ">Figure 2
<p>NRF2 nuclear translocation following 1 h hyperoxia. (<b>a</b>) Mild hyperoxia (30% O<sub>2</sub>); (<b>b</b>) high hyperoxia (100% O<sub>2</sub>); (<b>c</b>) very high hyperoxia (140% O<sub>2</sub>) before and after the recovery to normoxic conditions. Above histograms, the picture shows a representative western blot analysis. Results are expressed as fold change (mean ± SEM) in comparison to baseline (0), which was set at 1. * <span class="html-italic">p</span> &lt; 0.05, <span class="html-italic">** p</span> &lt; 0,01; for one-way ANOVA followed by Dunnett’s post hoc test.</p>
Full article ">Figure 3
<p>NF-κB (p65 subunit) nuclear translocation following 1 h hyperoxia. (<b>a</b>) Mild hyperoxia (30% O<sub>2</sub>); (<b>b</b>) high hyperoxia (100% O<sub>2</sub>); (<b>c</b>) very high hyperoxia (140% O<sub>2</sub>) before and after the recovery to normoxic conditions. Above histograms, the picture shows a representative western blot analysis. Results are expressed as fold change (mean ± SEM) in comparison to baseline (0), which was set at 1. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0,01; for one-way ANOVA followed by Dunnett’s post hoc test.</p>
Full article ">Figure 4
<p>Changes in intracellular glutathione levels in human PBMC from different groups of healthy subjects after one hour of oxygen exposure in (<b>a</b>) mild (30% O<sub>2</sub>), (<b>b</b>) high (100% O<sub>2</sub>) and (<b>c</b>) very high hyperoxia (140% O<sub>2</sub>). The histograms display the nmole of GSH for mg of protein (mean ± SEM) in comparison to baseline (Time 0), which was set at 10. * <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 vs. base line before oxygen exposure (time 0), for one-way ANOVA followed by Dunnett’s post hoc test.</p>
Full article ">Figure 5
<p>Zymographic analysis of matrix metallo-proteinase (MMPs) in human plasma from different groups of healthy subjects after one hour of oxygen exposure in (<b>a</b>) mild (30% O<sub>2</sub>), (<b>b</b>) high (100% O<sub>2</sub>) and (<b>c</b>) very high hyperoxia (140% O<sub>2</sub>). The pictures show a representative zymograph gel for all the subjects submitted to the analysis. The histograms display the percentage (mean ± SEM) in comparison to baseline (Time 0), which was set at 10. * <span class="html-italic">p</span> &lt; 0.05 vs. base line before oxygen exposure (Time 0), for one-way ANOVA followed by Dunnett’s post hoc test.</p>
Full article ">
14 pages, 3814 KiB  
Article
Delivery of the Radionuclide 131I Using Cationic Fusogenic Liposomes as Nanocarriers
by Rejhana Kolašinac, Dirk Bier, Laura Schmitt, Andriy Yabluchanskiy, Bernd Neumaier, Rudolf Merkel and Agnes Csiszár
Int. J. Mol. Sci. 2021, 22(1), 457; https://doi.org/10.3390/ijms22010457 - 5 Jan 2021
Cited by 10 | Viewed by 3801
Abstract
Liposomes are highly biocompatible and versatile drug carriers with an increasing number of applications in the field of nuclear medicine and diagnostics. So far, only negatively charged liposomes with intercalated radiometals, e.g., 64Cu, 99mTc, have been reported. However, the process of [...] Read more.
Liposomes are highly biocompatible and versatile drug carriers with an increasing number of applications in the field of nuclear medicine and diagnostics. So far, only negatively charged liposomes with intercalated radiometals, e.g., 64Cu, 99mTc, have been reported. However, the process of cellular uptake of liposomes by endocytosis is rather slow. Cellular uptake can be accelerated by recently developed cationic liposomes, which exhibit extraordinarily high membrane fusion ability. The aim of the present study was the development of the formulation and the characterization of such cationic fusogenic liposomes with intercalated radioactive [131I]I for potential use in therapeutic applications. The epithelial human breast cancer cell line MDA-MB-231 was used as a model for invasive cancer cells and cellular uptake of [131I]I was monitored in vitro. Delivery efficiencies of cationic and neutral liposomes were compared with uptake of free iodide. The best cargo delivery efficiency (~10%) was achieved using cationic fusogenic liposomes due to their special delivery pathway of membrane fusion. Additionally, human blood cells were also incubated with cationic control liposomes and free [131I]I. In these cases, iodide delivery efficiencies remained below 3%. Full article
(This article belongs to the Special Issue Functionalized Liposomes)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Chemical structures of the lipid molecules forming cationic fusogenic (<b>A</b>) and control endocytic liposomes (<b>B</b>). Preparation of iodide-loaded liposomes (<b>C</b>). After evaporation of the organic solvent, the dry lipid mixture was rehydrated in an iodide-containing buffer whereby liposomes containing iodide isotopes formed spontaneously between the lamellas. Iodide loading efficiency was determined on liposomes separated from the free solution by centrifugation.</p>
Full article ">Figure 2
<p>Representative curves of the hydrodynamic diameter (<b>A</b>) and zeta potential distributions (<b>B</b>) of cationic fusogenic (FL) and control endocytic liposomes (EL) with and without [<sup>127</sup>I]I<sup>−</sup> incubation. Hypothetical structures of the same liposomes based on the physicochemical characterization (<b>C</b>). Note that the same liposomal structures are hypothesized in the presence of both [<sup>131</sup>I]I<sup>−</sup> as well as [<sup>127</sup>I]I<sup>−</sup>.</p>
Full article ">Figure 3
<p>Absolute activity of free [<sup>131</sup>I]I<sup>−</sup> radionuclide solution, as well as the activity of [<sup>131</sup>I]I<sup>−</sup> intercalated into cationic fusogenic (FL) and control endocytic liposomal (EL) pellet (index p) and the activity of free [<sup>131</sup>I]I<sup>−</sup> in the supernatant (<b>A</b>). The uptake efficiency of free [<sup>131</sup>I]I<sup>−</sup> was determined using two different iodide concentrations, the initial one (<sup>131</sup>I<sup>−</sup>), and one quarter of this concentration (<sup>131</sup>I<sup>−</sup>/4), which was approximately identical to the amount of intercalated iodide in FLs. Activity changes of [<sup>131</sup>I]I<sup>−</sup> in phosphate buffered saline (PBS) and liposomal supernatants (<b>B</b>) as well as in liposomal pellets (<b>C</b>) were monitored over 16 days. Statistical significances were considered as follows: <span class="html-italic">p</span> &lt; 0.001 (***), <span class="html-italic">p</span> &lt; 0.01 (**).</p>
Full article ">Figure 4
<p>Delivery efficiencies of [<sup>131</sup>I]I<sup>−</sup> using liposomes as carrier particles and free [<sup>131</sup>I]I<sup>-</sup> from supernatants and neat solutions as controls to MDA MB-231 cancer cell line (<b>A</b>), human blood cells (<b>B</b>), primary neuronal cells (<b>C</b>), and Chinese hamster ovary (CHO) cells (<b>D</b>). Data are presented as mean (SD). N = 3. Statistical significances were considered as follows: <span class="html-italic">p</span> &lt; 0.001 (***), <span class="html-italic">p</span> &lt; 0.01 (**), <span class="html-italic">p</span> &lt; 0.05 (*).</p>
Full article ">Figure 5
<p>Monitoring the incorporation of cationic fusogenic (FL) and control endocytic liposomal (EL) pellet (index p) and supernatant (index s) containing [<sup>127</sup>I]I<sup>−</sup> into MDA MB-231 cancer cells and human blood cells by microscopy. The fluorescence signal of the lipid tracer DiR (red) and the cell nuclei staining of DAPI (blue) are shown in overlays. Phase contrast images reveal healthy cell morphologies. Scale bars, 20 µm.</p>
Full article ">
29 pages, 1356 KiB  
Review
Emerging Options for the Diagnosis of Bacterial Infections and the Characterization of Antimicrobial Resistance
by Simone Rentschler, Lars Kaiser and Hans-Peter Deigner
Int. J. Mol. Sci. 2021, 22(1), 456; https://doi.org/10.3390/ijms22010456 - 5 Jan 2021
Cited by 35 | Viewed by 10219
Abstract
Precise and rapid identification and characterization of pathogens and antimicrobial resistance patterns are critical for the adequate treatment of infections, which represent an increasing problem in intensive care medicine. The current situation remains far from satisfactory in terms of turnaround times and overall [...] Read more.
Precise and rapid identification and characterization of pathogens and antimicrobial resistance patterns are critical for the adequate treatment of infections, which represent an increasing problem in intensive care medicine. The current situation remains far from satisfactory in terms of turnaround times and overall efficacy. Application of an ineffective antimicrobial agent or the unnecessary use of broad-spectrum antibiotics worsens the patient prognosis and further accelerates the generation of resistant mutants. Here, we provide an overview that includes an evaluation and comparison of existing tools used to diagnose bacterial infections, together with a consideration of the underlying molecular principles and technologies. Special emphasis is placed on emerging developments that may lead to significant improvements in point of care detection and diagnosis of multi-resistant pathogens, and new directions that may be used to guide antibiotic therapy. Full article
Show Figures

Figure 1

Figure 1
<p>Typical timeframes required for techniques in current use for the diagnosis of bacterial infections. The classical cultivation of biological specimens, in combination with biochemical characterization, requires ~42 h as a minimum estimate. Replacement of biochemical methods with MALDI-TOF mass spectrometry (MS) reduces this timeframe significantly. The use of nucleic acid testing (NAT) bypasses the initial cultivation of clinical specimens, and thus reduces the timeframes to fewer than 4 h. However, NAT methods are sequence-dependent and involve only a limited number of primer-combinations; as such, these methods require a priori knowledge of the suspected pathogen(s). The use of NGS-based methods eliminates the need for any a priori knowledge of a suspected pathogen, although typical timeframes are increased.</p>
Full article ">Figure 2
<p>Possible shortcuts for bacterial pathogen identification from patient specimens, using MALDI-TOF MS. Typically, specimens are cultivated and identified via biochemical approaches. By MALDI-TOF-based analysis of single colonies obtained during the first round of cultivation, long turnaround times obligatory for biochemical characterization, can be bypassed. Furthermore, patient specimens can be directly analyzed by purification and concentration of bacterial cells, further shortening turnaround time. However, in case of polymicrobial infections, single bacteria need to be further separated, increasing the time to result. As polymicrobial infections cannot be ruled out in most clinical cases, a further separation of single species should be integrated into routine workflows.</p>
Full article ">Figure 3
<p>Illustration of different technologies for pathogen identification, suitable for point-of-care (POC) application. General approaches can be divided into antigen detection (<b>top</b>) and nucleic acid testing (<b>bottom</b>). Lateral Flow Immunoassays, for example, are readily applicable for detection of pathogen specific antigens, multiplexing-approaches for identification of several different species can be realized by incorporating different fluorescence labels (e.g., quantum dots). Furthermore, Plasmonic Biosensors show a great potential for POC applications, as sensitivity can be drastically increased and sample treatment can be avoided [<a href="#B105-ijms-22-00456" class="html-bibr">105</a>]. Another emerging technology for POC application is the Whispering Gallery Mode sensor technology, attracting much attention over the past decade. Here, the binding of molecules to the resonators surface can be detected as a change of the effective refractive index. Although the WGM technology displays a promising candidate, there are currently still several challenges hindering transformation into the clinical environment [<a href="#B78-ijms-22-00456" class="html-bibr">78</a>]. In case of approaches for POC applicable nucleic acid testing, several variable approaches are present. In general, shorter time periods during the amplification step can be achieved via implementation of paper-based (e.g., isothermal amplification [<a href="#B19-ijms-22-00456" class="html-bibr">19</a>]) or micro-fluidic- based (e.g., micro-fluidic PCR [<a href="#B106-ijms-22-00456" class="html-bibr">106</a>]) approaches. For detection of the amplicon, several different technologies can readily be used, including Nucleic Acid Lateral Flow Assays or intercalating dyes.</p>
Full article ">Figure 4
<p>Overview of current and promising technologies suitable for antimicrobial resistance profiling. Beside classical culture-based approaches (not included), currently available commercial solutions include PCR multiplex panels (e.g., Biofire<sup>®</sup>-FilmArray® panels (biomérieux)), MALDI-TOF MS (MTB-STAR Assays (Bruker Daltonics, Inc.)), biochemical tests (e.g., RAPIDEC® CARBA NP test (biomérieux)) and protein marker tests (e.g., RESIST-3 O.K.N. <span class="html-italic">K</span>-SeT (Coris BioConcept, Gembloux, Belgium)). Further promising technological advances have been made in the field of Next-generation Sequencing (e.g., [<a href="#B174-ijms-22-00456" class="html-bibr">174</a>]) adaption of biochemical tests to faster electrochemical formats (e.g., [<a href="#B167-ijms-22-00456" class="html-bibr">167</a>]) or development of electrochemical sensors for bacterial growth (e.g., [<a href="#B154-ijms-22-00456" class="html-bibr">154</a>]). Several of these approaches are suspected to result in commercially available solutions for antimicrobial resistance profiling in the future.</p>
Full article ">
22 pages, 5055 KiB  
Review
FUT8 Alpha-(1,6)-Fucosyltransferase in Cancer
by Kayla Bastian, Emma Scott, David J. Elliott and Jennifer Munkley
Int. J. Mol. Sci. 2021, 22(1), 455; https://doi.org/10.3390/ijms22010455 - 5 Jan 2021
Cited by 84 | Viewed by 10466
Abstract
Aberrant glycosylation is a universal feature of cancer cells that can impact all steps in tumour progression from malignant transformation to metastasis and immune evasion. One key change in tumour glycosylation is altered core fucosylation. Core fucosylation is driven by fucosyltransferase 8 (FUT8), [...] Read more.
Aberrant glycosylation is a universal feature of cancer cells that can impact all steps in tumour progression from malignant transformation to metastasis and immune evasion. One key change in tumour glycosylation is altered core fucosylation. Core fucosylation is driven by fucosyltransferase 8 (FUT8), which catalyses the addition of α1,6-fucose to the innermost GlcNAc residue of N-glycans. FUT8 is frequently upregulated in cancer, and plays a critical role in immune evasion, antibody-dependent cellular cytotoxicity (ADCC), and the regulation of TGF-β, EGF, α3β1 integrin and E-Cadherin. Here, we summarise the role of FUT8 in various cancers (including lung, liver, colorectal, ovarian, prostate, breast, melanoma, thyroid, and pancreatic), discuss the potential mechanisms involved, and outline opportunities to exploit FUT8 as a critical factor in cancer therapeutics in the future. Full article
(This article belongs to the Special Issue Glycobiology of Cancer)
Show Figures

Figure 1

Figure 1
<p>The reaction catalysed by FUT8. FUT8 transfers an <span class="html-small-caps">l</span>-fucose reside from GDP-β-<span class="html-small-caps">l</span>-fucose (GDP-Fuc) onto the innermost GlcNAc of an N-glycan to form an α-1,6 linkage.</p>
Full article ">Figure 2
<p>The sythetic pathway for GDP-fucose. GDP-fucose, an essential component of core fucosyaltion is produced by two different pathways within the cell. The predominat de novo pathway relies on GMD and FX proteins. Adapted from [<a href="#B32-ijms-22-00455" class="html-bibr">32</a>].</p>
Full article ">Figure 3
<p>Structure of FUT8. Published by [<a href="#B43-ijms-22-00455" class="html-bibr">43</a>] and reproduced with permission. (<b>a</b>) Ribbon structure of HsFUT8 with orange carbon atoms representing GDP and green carbon atoms representing a bi-antennary complex N-glycan (G0). The coiled-coil domain is colored in gray, the catalytic domain in red, and the SH3 domain in orange. The interdomain α3 and loop β10–β11 are colored in blue and aquamarine, respectively. Yellow sulfur atoms indicate disulfide bridges. The C-terminal loop is colored in black. Electron density maps are F<sub>O</sub>–F<sub>C</sub> (blue) contoured at 2.2σ for GDP and G0. (<b>b</b>) Surface representation of the HsFUT8-GDP-G0 complex.</p>
Full article ">
20 pages, 7599 KiB  
Article
RNA-Seq Provides New Insights into the Molecular Events Involved in “Ball-Skin versus Bladder Effect” on Fruit Cracking in Litchi
by Jun Wang, Xiao Fang Wu, Yong Tang, Jian Guo Li and Ming Lei Zhao
Int. J. Mol. Sci. 2021, 22(1), 454; https://doi.org/10.3390/ijms22010454 - 5 Jan 2021
Cited by 17 | Viewed by 3537
Abstract
Fruit cracking is a disorder of fruit development in response to internal or external cues, which causes a loss in the economic value of fruit. Therefore, exploring the mechanism underlying fruit cracking is of great significance to increase the economic yield of fruit [...] Read more.
Fruit cracking is a disorder of fruit development in response to internal or external cues, which causes a loss in the economic value of fruit. Therefore, exploring the mechanism underlying fruit cracking is of great significance to increase the economic yield of fruit trees. However, the molecular mechanism underlying fruit cracking is still poorly understood. Litchi, as an important tropical and subtropical fruit crop, contributes significantly to the gross agricultural product in Southeast Asia. One important agricultural concern in the litchi industry is that some famous varieties with high economic value such as ‘Nuomici’ are susceptible to fruit cracking. Here, the cracking-susceptible cultivar ‘Nuomici’ and cracking-resistant cultivar ‘Huaizhi’ were selected, and the samples including pericarp and aril during fruit development and cracking were collected for RNA-Seq analysis. Based on weighted gene co-expression network analysis (WGCNA) and the “ball-skin versus bladder effect” theory (fruit cracking occurs upon the aril expanding pressure exceeds the pericarp strength), it was found that seven co-expression modules genes (1733 candidate genes) were closely associated with fruit cracking in ‘Nuomici’. Importantly, we propose that the low expression level of genes related to plant hormones (Auxin, Gibberellins, Ethylene), transcription factors, calcium transport and signaling, and lipid synthesis might decrease the mechanical strength of pericarp in ‘Nuomici’, while high expression level of genes associated with plant hormones (Auxin and abscisic acid), transcription factors, starch/sucrose metabolism, and sugar/water transport might increase the aril expanding pressure, thereby resulting in fruit cracking in ‘Nuomici’. In conclusion, our results provide comprehensive molecular events involved in the “ball-skin versus bladder effect” on fruit cracking in litchi. Full article
(This article belongs to the Section Molecular Plant Sciences)
Show Figures

Figure 1

Figure 1
<p>The fruit phenotype at different stages of development (<b>A</b>), daily cracking rate (<b>B</b>), and the cumulative fruit cracking rate (<b>C</b>) between ‘Nuomici (NMC)’ and ‘Huaizhi (HZ)’ litchi. Vertical bars represent the standard error of three biological replicates, significant differences at the 0.01 level are indicated with two asterisks (**) according to the Student <span class="html-italic">t</span>-test.</p>
Full article ">Figure 2
<p>A Venn diagram showing the number of genes annotated by different databases including NR (RefSeq non-redundant proteins), SWiss-Prot (Swiss-Prot Protein Sequence Database), COG (cluster of orthologous group), and KEGG (Kyoto Encyclopedia of Genes and Genomes).</p>
Full article ">Figure 3
<p>Network construction and hierarchical clustering of differentially expressed genes (DEGs) in pericarp and aril. Genes with similar expression patterns were clustered. Each branch in the (<b>A</b>) and (<b>B</b>) represents one gene, and the color below each branch represents the co-expression module. Twelve and six modules in (<b>C</b>) and (<b>D</b>) are grouped by genes with the same color. The asterisk (*) indicates that the module has a significant correlation coefficient (<span class="html-italic">p</span> &lt; 0.05) with DCR. The number above (or below) the asterisk represents the gene numbers in this module.</p>
Full article ">Figure 4
<p>Identification of DEGs during fruit cracking in pericarp and aril. The DEGs in the blue part has a significant correlation coefficient (<span class="html-italic">p</span> &lt; 0.05) with DCR, and the DEGs in the green part are genes that showed differential expression in pericarp (<b>A</b>) and aril (<b>B</b>) between ‘Nuomici’ and ‘Huaizhi’ during fruit cracking.</p>
Full article ">Figure 5
<p>SOTA (Self-Organizing Tree Algorithm) clustering of gene expression patterns. Cluster analysis of DEGs in ‘Huaizhi’ and ‘Nuomici’ based on the Z-score normalized method. The light red dotted and light blue dotted lines represent the expression profiles of ‘Huaizhi’ and ‘Nuomici’ in each cluster, respectively. The red and blue solid line represent the average profile of ‘Huaizhi’ and ‘Nuomici’ in each cluster, respectively.</p>
Full article ">Figure 6
<p>Functional annotation of DEGs with BLAST searches.</p>
Full article ">Figure 7
<p>Heatmaps of DEGs in pericarp involved in specific pathways. Heatmaps showing the DEGs that are involved in calcium, lipid metabolism, development, hormone metabolism, cell wall, and transcription factors in pericarp. The ‘N’ and ‘H’ represent ‘Nuomici’ and ‘Huaizhi’, respectively. The numbers together with ‘N’ and ‘H’ indicate the days after anthesis. Fold changes in the ‘Nuomic’ relative to ‘Huaizhi’ after log2 conversion (FPKM values were converted to FPKM+1) are shown in a red-purple color scale: red, upregulated; purple, downregulated; white, unchanged.</p>
Full article ">Figure 8
<p>Heatmaps of DEGs in aril involved in specific pathways. Heatmaps showing the DEGs that are involved in hormone metabolism, transport, starch/sucrose metabolism, and transcription factors.</p>
Full article ">Figure 9
<p>qRT-PCR (Quantitative Real-time Polymerase Chain Reaction) analysis of 40 randomly selected genes (<b>A</b>) pericarp; (<b>B</b>) aril. <span class="html-italic">Lc-EF-1α</span> and <span class="html-italic">Lc-Atcin</span> were used as reference genes for normalization of gene-expression data. Red and blue bars represent the data yielded by qRT-PCR in ‘Nuomici’ and ‘Huaizhi’, respectively. Red and blue lines represent the data yielded by RNA sequencing in ‘Nuomici’ and ‘Huaizhi’, respectively. Error bars indicate standard errors of the means (<span class="html-italic">n</span> = 3). The full name of each gene abbreviation is <span class="html-italic">MLO</span>: <span class="html-italic">Mildew resistance locus O</span>, <span class="html-italic">CML</span>: <span class="html-italic">Calmodulin-like protein</span>, <span class="html-italic">PIP</span>: <span class="html-italic">Aquaporin</span>, <span class="html-italic">ACP</span>: <span class="html-italic">Acyl carrier protein</span>; <span class="html-italic">ARF</span>: <span class="html-italic">Auxin response factor, CIPK</span>: <span class="html-italic">CBL-interacting serine/threonine-protein kinase, CGNC</span>: <span class="html-italic">Cyclic nucleotide-gated ion channel</span>, <span class="html-italic">PAEI</span>: <span class="html-italic">Pectinesterase inhibitors</span>, <span class="html-italic">PPR</span>: <span class="html-italic">Pentatricopeptide repeat-containing protein</span>, <span class="html-italic">BAK</span>: <span class="html-italic">brassinosteroid-insensitive associated receptor kinase</span>, <span class="html-italic">STK</span>: <span class="html-italic">Serine/threonine-protein kinase</span>, <span class="html-italic">TPS</span>: <span class="html-italic">Trehalose-6-phosphate synthase</span>, <span class="html-italic">TPP</span>: <span class="html-italic">Trehalose-phosphate phosphatase</span>, <span class="html-italic">XPT</span>: <span class="html-italic">Xylulose 5-phosphate translocate, EG: endo-1,4-β-glucanase</span>.</p>
Full article ">Figure 10
<p>Coefficient analysis between gene expression patterns between qRT-PCR and RNA-Seq data. Extremely significant difference levels (<span class="html-italic">p</span> &lt; 0.01) are indicated with two asterisks (**).</p>
Full article ">Figure 11
<p>A preliminary gene network involved in litchi fruit cracking. In pericarp, the expression level of genes related to hormones IAA, GA, and ETH is low, which might directly repress the pericarp growth or indirectly downregulate specific transcription factors to repress the expression of growth-promoted genes, as a consequence, the pericarp might provide a limited space for aril expanding. In addition, the transcripts of genes related to calcium transport and signaling, and wax synthesis is also low in pericarp, while the transcript level of genes related to cell wall remodeling is high in pericarp. As a result, the pericarp of ‘Nuomici’ might be both developmental retardation and weak mechanical strength. During the aril growth, the transcripts of genes related to IAA and ABA is high, which might directly promote aril expanding or indirectly upregulate transcription factors such as <span class="html-italic">WRKY</span>, <span class="html-italic">bHLH</span>, <span class="html-italic">DOF</span>, <span class="html-italic">NAC</span>, and <span class="html-italic">MYB</span> to increase growth-promoted gene expression. Additionally, the transcript level of genes involved in starch/sucrose metabolism and transport is high, which could cause high osmotic potential. Notably, the expression level of one <span class="html-italic">PIP</span> is also high in aril, which together imply that the fruit cracking could be accelerated when fruit encounter a heavy rain since the aril might possess a strong capacity of water absorption to enhance the aril expanding pressure. Collectively, these DEGs might weaken the pericarp strength, but enhance the aril expanding pressure, thereby making the ‘Nuomici’ fruit susceptible to cracking.</p>
Full article ">
13 pages, 928 KiB  
Review
Macrophage Autophagy and Silicosis: Current Perspective and Latest Insights
by Shiyi Tan and Shi Chen
Int. J. Mol. Sci. 2021, 22(1), 453; https://doi.org/10.3390/ijms22010453 - 5 Jan 2021
Cited by 59 | Viewed by 6416
Abstract
Silicosis is an urgent public health problem in many countries. Alveolar macrophage (AM) plays an important role in silicosis progression. Autophagy is a balanced mechanism for regulating the cycle of synthesis and degradation of cellular components. Our previous study has shown that silica [...] Read more.
Silicosis is an urgent public health problem in many countries. Alveolar macrophage (AM) plays an important role in silicosis progression. Autophagy is a balanced mechanism for regulating the cycle of synthesis and degradation of cellular components. Our previous study has shown that silica engulfment results in lysosomal rupture, which may lead to the accumulation of autophagosomes in AMs of human silicosis. The excessive accumulation of autophagosomes may lead to apoptosis in AMs. Herein, we addressed some assumptions concerning the complex function of autophagy-related proteins on the silicosis pathogenesis. We also recapped the molecular mechanism of several critical proteins targeting macrophage autophagy in the process of silicosis fibrosis. Furthermore, we summarized several exogenous chemicals that may cause an aggravation or alleviation for silica-induced pulmonary fibrosis by regulating AM autophagy. For example, lipopolysaccharides or nicotine may have a detrimental effect combined together with silica dust via exacerbating the blockade of AM autophagic degradation. Simultaneously, some natural product ingredients such as atractylenolide III, dioscin, or trehalose may be the potential AM autophagy regulators, protecting against silicosis fibrosis. In conclusion, the deeper molecular mechanism of these autophagy targets should be explored in order to provide feasible clues for silicosis therapy in the clinical setting. Full article
(This article belongs to the Section Molecular Pathology, Diagnostics, and Therapeutics)
Show Figures

Figure 1

Figure 1
<p>The general silicosis pathological mechanism. The attack of silica dust triggers the phagocytosis of alveolar macrophages (AMs). However, the lysosomal membrane of AM will be disrupted by the H-bonding reaction, causing AM apoptosis. Notably, AM apoptosis may be regulated by the mitochondria apoptotic pathway, Fas apoptotic pathway, nuclear factor kappa-B (NF-κB) apoptotic pathway, and p53 apoptotic pathway. Apoptotic AM secrets a series of inflammatory factors. Eventually, the proliferation, activation, and migration of fibroblasts synthesize and release collagen, resulting in silicosis fibrosis.</p>
Full article ">Figure 2
<p>Determined mechanism of silicosis pathogenesis mediated by macrophage autophagy. (<b>a</b>) Normally, the pre-autophagosomal structure (PAS) begins to engulf cytosolic components, following which the signal of autophagy is activated. Subsequently, the autophagosomal membrane expands, matures, and closes gradually. Once closed, the autophagosome will fuse with the lysosome to form the autophagolysosome, degrading its wrapped contents. Several autophagy-related proteins act critically, indicating effect during the process of autophagy. Sequestosome 1 (p62/SQSTM1) interacts with ubiquitinated substrates, participating in the subsequent process of autophagic degradation. Microtubule-associated protein 1A/1B-light chain 3-I (LC3-I) conjugates with phosphatidylethanolamine to form LC3-II, involved in the expansion, maturation, and closure of autophagosome. Additionally, lysosome-associated membrane protein (LAMP) is attached to the surface of the lysosome. (<b>b</b>) When invading macrophages (especially alveolar macrophages) excessively, silica dust will cause autophagosomes to accumulate and lysosomes to disrupt (i.e., the dysfunction of the autophagy-lysosomal system). (<b>c</b>) Silica dust invades macrophages (especially alveolar macrophages) via a class A scavenger receptor (SR-A) and causes the blockade of autophagic degradation. The impairment of autophagy function caused by silica further leads to excessive macrophage apoptosis by decreasing the level of BCL2 and increasing the level of BCL2-Associated X (Bax). The inflammation and subsequent silicosis fibrosis occur eventually. Notably, nuclear transfer of nuclear factor kappa-B (NF-κB) may be an important link between autophagy and apoptosis in silicosis. In this pathological progression, disruption of the autophagy-lysosomal system induces macrophage apoptosis through the activation of NACHT-, LRR-, and PYD domain-containing protein 3 (NALP3). NALP3 also increases macrophage apoptosis via the nuclear transfer of NF-κB. Meanwhile, activation of BCL2-binding component 3 (BBC3) or Monocyte chemoattractant protein-1-induced protein 1 (MCPIP1) promotes macrophage apoptosis through exacerbating the autophagic degradation. However, the functional role of Toll-like receptor 4 (TLR4) is still controversial.</p>
Full article ">
12 pages, 3317 KiB  
Article
Intracerebroventricular Treatment with 2-Hydroxypropyl-β-Cyclodextrin Decreased Cerebellar and Hepatic Glycoprotein Nonmetastatic Melanoma Protein B (GPNMB) Expression in Niemann–Pick Disease Type C Model Mice
by Madoka Fukaura, Yoichi Ishitsuka, Seiichi Shirakawa, Naoki Ushihama, Yusei Yamada, Yuki Kondo, Toru Takeo, Naomi Nakagata, Keiichi Motoyama, Taishi Higashi, Hidetoshi Arima, Yuki Kurauchi, Takahiro Seki, Hiroshi Katsuki, Katsumi Higaki, Muneaki Matsuo and Tetsumi Irie
Int. J. Mol. Sci. 2021, 22(1), 452; https://doi.org/10.3390/ijms22010452 - 5 Jan 2021
Cited by 20 | Viewed by 4584
Abstract
Niemann–Pick disease type C (NPC) is a recessive hereditary disease caused by mutation of the NPC1 or NPC2 gene. It is characterized by abnormality of cellular cholesterol trafficking with severe neuronal and hepatic injury. In this study, we investigated the potential of glycoprotein [...] Read more.
Niemann–Pick disease type C (NPC) is a recessive hereditary disease caused by mutation of the NPC1 or NPC2 gene. It is characterized by abnormality of cellular cholesterol trafficking with severe neuronal and hepatic injury. In this study, we investigated the potential of glycoprotein nonmetastatic melanoma protein B (GPNMB) to act as a biomarker reflecting the therapeutic effect of 2-hydroxypropyl-β-cyclodextrin (HP-β-CD) in an NPC mouse model. We measured serum, brain, and liver expression levels of GPNMB, and evaluated their therapeutic effects on NPC manifestations in the brain and liver after the intracerebroventricular administration of HP-β-CD in Npc1 gene-deficient (Npc1−/−) mice. Intracerebroventricular HP-β-CD inhibited cerebellar Purkinje cell damage in Npc1−/− mice and significantly reduced serum and cerebellar GPNMB levels. Interestingly, we also observed that the intracerebral administration significantly reduced hepatic GPNMB expression and elevated serum ALT in Npc1−/− mice. Repeated doses of intracerebroventricular HP-β-CD (30 mg/kg, started at 4 weeks of age and repeated every 2 weeks) drastically extended the lifespan of Npc1−/− mice compared with saline treatment. In summary, our results suggest that GPNMB level in serum is a potential biomarker for evaluating the attenuation of NPC pathophysiology by intracerebroventricular HP-β-CD treatment. Full article
Show Figures

Figure 1

Figure 1
<p>Potential of the intracerebroventricular administration of 2-hydroxypropyl-β-cyclodextrin (HP-β-CD) to protect against reductions in body weight (<b>A</b>) and cerebellar Purkinje cells (B and C) in <span class="html-italic">Npc1</span><sup>−/−</sup> mice. HP-β-CD was administered at 21.4 μmol/kg in mice at 4 and 6 weeks of age. We recorded the changes in body weight once a week from 4 to 8 weeks old. We collected samples from the whole brain, liver, and blood at 8 weeks and 3 days of age and analyzed them. Histological images of the mouse cerebellum (<b>B</b>). Scale bar: 100 μm. Quantitative evaluation of calbindin-positive cells (<b>C</b>). Data are presented as the mean ± S.E. (<span class="html-italic">n</span> = 8–11). *** <span class="html-italic">p</span> &lt; 0.001 compared with the wild-type group. ### <span class="html-italic">p</span> &lt; 0.001 compared with the saline group.</p>
Full article ">Figure 2
<p>Suppression of serum soluble glycoprotein nonmetastatic melanoma protein B (sGPNMB) levels in <span class="html-italic">Npc1</span><sup>−/−</sup> mice treated with 2-hydroxypropyl-β-cyclodextrin (HP-β-CD) intracerebroventricularly. Levels of sGPNMB in mouse serum were assayed by ELISA. Data are presented as the mean ± S.E. (<span class="html-italic">n</span> = 8–11). *** <span class="html-italic">p</span> &lt; 0.001 compared with the wild-type group. ### <span class="html-italic">p</span> &lt; 0.001 compared with the saline group.</p>
Full article ">Figure 3
<p>Intracerebroventricular treatment with 2-hydroxypropyl-β-cyclodextrin (HP-β-CD) attenuated the abnormal increase of glycoprotein nonmetastatic melanoma protein B (GPNMB) expression in <span class="html-italic">Npc1</span><sup>−/−</sup> mice. Assays of GPNMB levels in brain using ELISA (<b>A</b>) and immunohistochemical staining of GPNMB (<b>B</b>) were performed. Scale bar: 100 μm. Data are presented as the mean ± S.E. (<span class="html-italic">n</span> = 4–7). *** <span class="html-italic">p</span> &lt; 0.01 compared with the wild-type group. ### <span class="html-italic">p</span> &lt; 0.001 compared with the saline group.</p>
Full article ">Figure 4
<p>Attenuating effects of intracerebroventricular 2-hydroxypropyl-β-cyclodextrin (HP-β-CD) treatment on liver manifestations in <span class="html-italic">Npc1</span><sup>−/−</sup> mice. Glycoprotein nonmetastatic melanoma protein B (GPNMB) levels in liver homogenate (<b>A</b>), GPNMB immunohistochemistry in liver (<b>B</b>), serum alanine aminotransferase (ALT) levels (<b>C</b>), and hematoxylin-and-eosin (H&amp;E)-stained hepatic histology (<b>D</b>) are shown. Scale bar: 100 μm for GPNMB and H&amp;E staining. Each bar represents the mean ± S.E.M. (<span class="html-italic">n</span> = 8–11). *** <span class="html-italic">p</span> &lt; 0.001 compared with the wild-type group. # <span class="html-italic">p</span> &lt; 0.05, ### <span class="html-italic">p</span> &lt; 0.001 compared with the saline group.</p>
Full article ">Figure 5
<p>Life-prolonging effect of intracerebroventricular administration of 2-hydroxypropyl-β-cyclodextrin (HP-β-CD) in <span class="html-italic">Npc1</span><sup>−/−</sup> mice. HP-β-CD treatment (21.4 μmol/kg) was started at 4 weeks of age, with repeated injection into the mice every 2 weeks. Kaplan–Meier survival curves for the <span class="html-italic">Npc1</span><sup>−/−</sup> mice are shown (<span class="html-italic">n</span> = 4 for the saline-treated group (circle) and <span class="html-italic">n</span> = 7 for the HP-β-CD (biweekly) group (square)). A significant difference between the two groups was identified by the log-rank test (<span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">
15 pages, 1802 KiB  
Article
Impairment of IGF-1 Signaling and Antioxidant Response Are Associated with Radiation Sensitivity and Mortality
by Saeed Y. Aghdam, Doreswamy Kenchegowda, Gregory P. Holmes-Hampton, Maria Moroni and Sanchita P. Ghosh
Int. J. Mol. Sci. 2021, 22(1), 451; https://doi.org/10.3390/ijms22010451 - 5 Jan 2021
Cited by 5 | Viewed by 2578
Abstract
Following exposure to high doses of ionizing radiation, diverse strains of vertebrate species will manifest varying levels of radiation sensitivity. To understand the inter-strain cellular and molecular mechanisms of radiation sensitivity, two mouse strains with varying radiosensitivity (C3H/HeN, and CD2F1), were exposed to [...] Read more.
Following exposure to high doses of ionizing radiation, diverse strains of vertebrate species will manifest varying levels of radiation sensitivity. To understand the inter-strain cellular and molecular mechanisms of radiation sensitivity, two mouse strains with varying radiosensitivity (C3H/HeN, and CD2F1), were exposed to total body irradiation (TBI). Since Insulin-like Growth Factor-1 (IGF-1) signaling pathway is associated with radiosensitivity, we investigated the link between systemic or tissue-specific IGF-1 signaling and radiosensitivity. Adult male C3H/HeN and CD2F1 mice were irradiated using gamma photons at Lethal Dose-70/30 (LD70/30), 7.8 and 9.35 Gy doses, respectively. Those mice that survived up to 30 days post-irradiation, were termed the survivors. Mice that were euthanized prior to 30 days post-irradiation due to deteriorated health were termed decedents. The analysis of non-irradiated and irradiated survivor and decedent mice showed that inter-strain radiosensitivity and post-irradiation survival outcomes are associated with activation status of tissue and systemic IGF-1 signaling, nuclear factor erythroid 2–related factor 2 (Nrf2) activation, and the gene expression profile of cardiac mitochondrial energy metabolism pathways. Our findings link radiosensitivity with dysregulation of IGF-1 signaling, and highlight the role of antioxidant gene response and mitochondrial function in radiation sensitivity. Full article
(This article belongs to the Section Molecular Toxicology)
Show Figures

Figure 1

Figure 1
<p>Irradiation scheme for the CD2F1 and C3H/HeN adult male mice. Animals were irradiated at LD<sub>70/30</sub> dose. The day of irradiation was considered as day 0; animals that underwent unscheduled euthanasia due to severe health deterioration before day 30 post-irradiation were considered as decedents. Animals that survived from radiation exposure were euthanized on day 30 post-irradiation (scheduled euthanasia).</p>
Full article ">Figure 2
<p>Serum IGF-1 and Nitric Oxide (NO) levels in sham and LD<sub>70/30</sub> irradiated CD2F1 and C3H/HeN adult male mice. (<b>A</b>) ELISA analysis of serum IGF-1 levels in sham, irradiated decedent (Decd.) and irradiated survivor (Surv.) mice. <span class="html-italic">n</span> = 7 for sham animals of both strains, <span class="html-italic">n</span> = 8 for decedent animals of both strains, <span class="html-italic">n</span> = 3 for survivors of both strains. (<b>B</b>) Serum nitric oxide levels in sham, irradiated decedent (Decd.) and irradiated survivor (Surv.) mice. <span class="html-italic">n</span> = 8 for sham animals of both strains, <span class="html-italic">n</span> = 8 for CD2F1 decedents, <span class="html-italic">n</span> = 10 for C3H/HeN decedents, <span class="html-italic">n</span> = 3 for survivors of both strains. Data analyzed by student’s <span class="html-italic">t</span>-test. Results presented as mean ± SEM, *** <span class="html-italic">p</span> &lt; 0.001 and **** <span class="html-italic">p</span> &lt; 0.0001, <span class="html-italic">p</span> &gt; 0.05 considered as ‘not significant’ (ns).</p>
Full article ">Figure 3
<p>Western blot analysis of IGF-1R, Akt and Nrf2 activation in sham and LD<sub>70/30</sub> irradiated CD2F1 and C3H/HeN adult male mice. (<b>A</b>,<b>B</b>) Analysis of phosphorylated IGF-1 receptor (p.IGF1R, Tyr1135/1136), total IGF-1R, phosphorylated Akt (p.Akt, Ser473), total Akt, phosphorylated Nrf2 (p.Nrf2, Ser40 residue), β–Tubulin and β–Actin in heart protein extracts from sham, irradiated decedent (Decd.) and irradiated survivor (Surv.) mice. (<b>C</b>,<b>D</b>) Quantification of the IGF-1R and Nrf2 activation. The Y axis in graphs represents the ratios of the respective measured pixel intensities of Western blot bands. Total number of analyzed samples from both CD2F1 and C3H strains: sham—<span class="html-italic">n</span> = 6; decedent—<span class="html-italic">n</span> = 6; survivor <span class="html-italic">n</span> = 3. Data analyzed by Student’s <span class="html-italic">t</span>-test and are presented as mean ± 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 and **** <span class="html-italic">p</span> &lt; 0.0001, <span class="html-italic">p</span> &gt; 0.05 considered as ‘not significant’ (ns).</p>
Full article ">Figure 4
<p>Peroxide and ATP levels in sham and LD<sub>70/30</sub> irradiated survivor and decedent heart samples from CD2F1 and C3H/HeN mice. Graphs representing the heart peroxide (<b>A</b>) and ATP (<b>B</b>) levels measured in sham and irradiated CD2F1 and C3H/HeN mice. Total number of analyzed samples: <span class="html-italic">n</span> = 8 for sham and irradiated decedent mice and <span class="html-italic">n</span> = 3 for irradiated survivors. Data analyzed by Student’s <span class="html-italic">t</span>-test, presented as mean ± SEM, * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001 and <span class="html-italic">p</span> &gt; 0.05, ns—not significant.</p>
Full article ">Figure 5
<p>PCR array analysis of genes involved in nitric oxide (NO), oxidative stress and mitochondrial energy metabolism in heart samples from sham and LD<sub>70/30</sub> irradiated CD2F1 and C3H/HeN mice. Quantitative PCR array analysis showing the fold changes in the expression of NO (<b>A</b>,<b>B</b>) oxidative stress (<b>C</b>,<b>D</b>) and mitochondrial energy metabolism (<b>E</b>,<b>F</b>) mRNA levels in heart tissues from sham, irradiated decedent and irradiated survivor mice. The Y axis shows the relative fold changes in expression of the genes in irradiated survivors and decedents compared with the sham. Only those genes showing changes of three-fold or higher, either in the survivors or decedents are shown. The dotted line indicates 3-fold upregulation or downregulation in the graphs. Number of samples for both CD2F1 and C3H strains: Sham (control)—<span class="html-italic">n</span> = 7; decedent—<span class="html-italic">n</span> = 8; survivors—<span class="html-italic">n</span> = 3.</p>
Full article ">Figure 6
<p>Western blot analysis of IGF-1R, Akt and Nrf2 in lung and kidney samples from sham and LD<sub>70/30</sub> irradiated CD2F1 and C3H/HeN adult male mice. (<b>A</b>) Analysis of phosphorylated IGF-1 receptor (p.IGF1R, Tyr1135/1136), total IGF-1R, phosphorylated Akt (p.Akt, Ser473), total Akt and (<b>C</b>) phosphorylated Nrf2 (p.Nrf2, Ser40) in lung protein extracts from sham, irradiated decedent (Decd.) and irradiated survivor (Surv.) mice. β–Actin used as loading control. (<b>B</b>,<b>D</b>) Quantification of IGF-1R and Nrf2 activation in the lung samples. IGF-1R activation is determined by calculating the ratio of the pixel intensity of p.IGF-1R to total IGF-1R levels. Nrf2 activation is determined by calculating the ratio of pixel intensity of the p.Nrf2 to respective β–Actin band. (<b>E</b>) Western blot analysis of phosphorylated IGF-1 receptor (p.IGF1R, Tyr1135/1136), total IGF-1R and total Akt in kidney protein extracts from sham, irradiated decedent (Decd.) and irradiated survivor (Surv.) mice. GAPDH is used as loading control. Total number of analyzed samples for both strains: <span class="html-italic">n</span> = 6 for sham, <span class="html-italic">n</span> = 6 for decedents, <span class="html-italic">n</span> = 3 for survivors. Results analyzed by Student’s <span class="html-italic">t</span>-test; presented as mean ± SEM, * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001 and <span class="html-italic">p</span> &gt; 0.05 is considered as ‘not significant’ (ns).</p>
Full article ">
16 pages, 4667 KiB  
Article
CHID1 Is a Novel Prognostic Marker of Non-Small Cell Lung Cancer
by Olga V. Kovaleva, Madina A. Rashidova, Daria V. Samoilova, Polina A. Podlesnaya, Rasul M. Tabiev, Valeria V. Mochalnikova and Alexei Gratchev
Int. J. Mol. Sci. 2021, 22(1), 450; https://doi.org/10.3390/ijms22010450 - 5 Jan 2021
Cited by 8 | Viewed by 4029
Abstract
There is an urgent need for identification of new prognostic markers and therapeutic targets for non-small cell lung cancer (NSCLC). In this study, we evaluated immune cells markers in 100 NSCLC specimens. Immunohistochemical analysis revealed no prognostic value for the markers studied, except [...] Read more.
There is an urgent need for identification of new prognostic markers and therapeutic targets for non-small cell lung cancer (NSCLC). In this study, we evaluated immune cells markers in 100 NSCLC specimens. Immunohistochemical analysis revealed no prognostic value for the markers studied, except CD163 and CD206. At the same time, macrophage markers iNOS and CHID1 were found to be expressed in tumor cells and associated with prognosis. We showed that high iNOS expression is a marker of favorable prognosis for squamous cell lung carcinoma (SCC), and NSCLC in general. Similarly, high CHID1 expression is a marker of good prognosis in adenocarcinoma and in NSCLC in general. Analysis of prognostic significance of a high CHID1/iNOS expression combination showed favorable prognosis with 20 months overall survival of patients from the low CHID1/iNOS expression group. For the first time, we demonstrated that CHID1 can be expressed by NSCLC cells and its high expression is a marker of good prognosis for adenocarcinoma and NSCLC in general. At the same time, high expression of iNOS in tumor cells is a marker of good prognosis in SCC. When used in combination, CHID1 and iNOS show a very good prognostic capacity for NSCLC. We suggest that in the case of lung cancer, tumor-associated macrophages are likely ineffective as a therapeutic target. At the same time, macrophage markers expressed by tumor cells may be considered as targets for anti-tumor therapy or, as in the case of CHID1, as potential anti-tumor agents. Full article
(This article belongs to the Special Issue The New Molecular Strategies under Development in Thoracic Oncology)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>). Mann–Whitney analysis of CD3, CD8, and FOXP3 expression association with the clinicopathological properties of the analyzed tumors. * indicates <span class="html-italic">p</span> &lt; 0.05, *** indicates <span class="html-italic">p</span> &lt; 0.001. (<b>B</b>). Kaplan–Meier curves of overall survival (OS) in non-small cell lung cancer (NSCLC) based on CD3, CD8, and FOXP3 expression.</p>
Full article ">Figure 2
<p>Immunohistochemical analysis of CD206, CD204, CD163, and CD68 in NSCLC samples.</p>
Full article ">Figure 3
<p>Mann–Whitney analysis of CD68, CD163, CD206, CD204, and CHID1 expression association with clinicopathological properties of the analyzed tumors. * indicates <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 4
<p>Kaplan–Meier curves of overall survival (OS) in different subgroups of NSCLC based on CD68, CD163, CD206, CD204, and CHID1 expression in macrophages.</p>
Full article ">Figure 5
<p>Immunohistochemical analysis of CHID1, iNOS, IDO1, and PD-L1 expression in NSCLC samples. AC—adenocarcinoma, SCC—squamous cell carcinoma. Magnification—100×.</p>
Full article ">Figure 6
<p>Mann–Whitney analysis of CHID1, iNOS, IDO1 and PD-L1 expression association with clinicopathologic properties of analyzed tumors. * indicates <span class="html-italic">p</span> &lt; 0.05, *** indicates <span class="html-italic">p</span> &lt; 0.001, **** indicates <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 7
<p>Kaplan–Meier curves of overall survival (OS) in different subgroups of NSCLC based on CHID1, iNOS, IDO1, and PD-L1 expression in tumor cells.</p>
Full article ">Figure 8
<p>Kaplan–Meier curves of overall survival (OS) of NSCLC based on combined CHID1/iNOS expression analysis.</p>
Full article ">
13 pages, 2481 KiB  
Article
Tranexamic Acid Promotes Murine Bone Marrow-Derived Osteoblast Proliferation and Inhibits Osteoclast Formation In Vitro
by Anke Baranowsky, Jessika Appelt, Kristina Tseneva, Shan Jiang, Denise Jahn, Serafeim Tsitsilonis, Karl-Heinz Frosch and Johannes Keller
Int. J. Mol. Sci. 2021, 22(1), 449; https://doi.org/10.3390/ijms22010449 - 5 Jan 2021
Cited by 8 | Viewed by 2722
Abstract
Despite modern surgical trauma care, bleeding contributes to one-third of trauma-related death. A significant improvement was obtained through the introduction of tranexamic acid (TXA), which today is widely used in emergency and elective orthopedic surgery to control bleeding. However, concerns remain regarding potential [...] Read more.
Despite modern surgical trauma care, bleeding contributes to one-third of trauma-related death. A significant improvement was obtained through the introduction of tranexamic acid (TXA), which today is widely used in emergency and elective orthopedic surgery to control bleeding. However, concerns remain regarding potential adverse effects on bone turnover and regeneration. Therefore, we employed standardized cell culture systems including primary osteoblasts, osteoclasts, and macrophages to evaluate potential effects of TXA on murine bone cells. While osteoblasts derived from calvarial digestion were not affected, TXA increased cell proliferation and matrix mineralization in bone marrow-derived osteoblasts. Short-term TXA treatment (6 h) failed to alter the expression of osteoblast markers; however, long-term TXA stimulation (10 days) was associated with the increased expression of genes involved in osteoblast differentiation and extracellular matrix synthesis. Similarly, whereas short-term TXA treatment did not affect gene expression in terminally differentiated osteoclasts, long-term TXA stimulation resulted in the potent inhibition of osteoclastogenesis. Finally, in bone marrow-derived macrophages activated with LPS, simultaneous TXA treatment led to a reduced expression of inflammatory cytokines and chemokines. Collectively, our study demonstrates a differential action of TXA on bone cells including osteoanabolic, anti-resorptive, and anti-inflammatory effects in vitro which suggests novel treatment applications. Full article
Show Figures

Figure 1

Figure 1
<p>TXA does not affect the differentiation or function of calvaria-derived osteoblasts. (<b>a</b>) Representative alizarin red stainings of calvaria-derived osteoblasts from WT mice differentiated in the presence of indicated concentrations of TXA for 10 days in osteogenic medium. Scale bar 10 mm. The quantification of extracellular matrix mineralization is indicated below. (<b>b</b>) qRT-PCR expression analysis for the indicated genes in calvaria-derived osteoblasts at day 10 of osteogenic differentiation, stimulated with TXA (1 mg/mL) during the entire course of cell differentiation. (<b>c</b>) qRT-PCR expression analysis for the indicated genes in the same samples. (<b>d</b>) MTT proliferation assay of calvaria-derived osteoblasts stimulated with 1 mg/mL TXA for 6 h at the indicated days of differentiation. For (<b>a</b>–<b>d</b>), <span class="html-italic">n</span> = 4–6 independent cultures per group were used, as indicated by individual data points. Data presented are means ± SD. Gene abbreviations: runt-related transcription factor 2 (<span class="html-italic">Runx2</span>), osterix (<span class="html-italic">Sp7</span>), alkaline phosphatase (<span class="html-italic">Alpl</span>), alpha-1 type I collagen (<span class="html-italic">Col1a1</span>), osteocalcin (<span class="html-italic">Bglap</span>), sclerostin (<span class="html-italic">Sost</span>), Rankl (<span class="html-italic">Tnfsf11</span>), osteoprotegerin (<span class="html-italic">Tnfrsf11b</span>).</p>
Full article ">Figure 2
<p>TXA promotes extracellular matrix mineralization. (<b>a</b>) Representative alizarin red stainings of bone marrow-derived osteoblasts from WT mice differentiated in the presence of indicated concentrations of TXA for 10 days in osteogenic medium. Scale bar 10 mm. The quantification of extracellular matrix mineralization is indicated below. (<b>b</b>) qRT-PCR expression analysis for the indicated genes in bone marrow-derived osteoblasts at day 10 of osteogenic differentiation, stimulated with TXA (1 mg/mL) during the entire course of cell differentiation. (<b>c</b>,<b>d</b>) qRT-PCR expression analysis for the indicated genes in bone marrow-derived osteoblasts at day 2 (<b>c</b>) and day 10 (<b>d</b>) of osteogenic differentiation, stimulated with TXA for 6 h at the indicated concentrations after serum starvation overnight. (<b>e</b>) qRT-PCR expression analysis for the indicated genes in the same cells at day 2 or (<b>f</b>) day 10 of differentiation. For (<b>a</b>–<b>f</b>), <span class="html-italic">n</span> = 3–4 independent cultures per group were used. Data presented are means ± SD. Gene abbreviations: runt-related transcription factor 2 (<span class="html-italic">Runx2</span>), osterix (<span class="html-italic">Sp7</span>), alkaline phosphatase (<span class="html-italic">Alpl</span>), alpha-1 type I collagen (<span class="html-italic">Col1a1</span>), osteocalcin (<span class="html-italic">Bglap</span>), sclerostin (<span class="html-italic">Sost</span>), Rankl (<span class="html-italic">Tnfsf11</span>), osteoprotegerin (<span class="html-italic">Tnfrsf11b</span>).</p>
Full article ">Figure 3
<p>TXA enhances cell proliferation of bone marrow-derived osteoblasts. MTT proliferation assay of bone marrow-derived osteoblasts stimulated with 1 mg/mL TXA for 6 h at the indicated time points; <span class="html-italic">n</span> = 3–5 independent cultures per group were used. Data presented are means ± SD.</p>
Full article ">Figure 4
<p>Inhibition of early osteoclastogenesis through TXA. (<b>a</b>) TRAP activity staining of WT bone marrow cells at day 7 of differentiation, cultured in the presence of M-CSF and RANKL and continuous exposure to vehicle or TXA at the indicated concentrations. Scale bars = 50 μm. The quantification of osteoclast numbers per viewing field is depicted on the left (Ocl.N./VF). (<b>b</b>) TRAP activity staining of WT bone marrow cells at day 7 of differentiation, cultured in the presence of M-CSF and RANKL, and exposure to vehicle or TXA at the indicated concentrations only at day 1 of osteoclast differentiation. The quantification of osteoclast numbers per viewing field is depicted on the left (Ocl.N./VF). (<b>c</b>) TRAP activity staining of WT osteoclasts at day 7 of differentiation, cultured in the presence of M-CSF and RANKL and exposed to vehicle or TXA at the indicated concentrations for 24 h. (<b>d</b>) MTT proliferation assay of WT osteoclasts at day 7 of differentiation stimulated with the indicated concentrations of TXA for 6 h. (<b>e</b>) qRT-PCR expression analysis for the indicated genes in bone marrow-derived osteoclasts at day 7, stimulated with TXA (1 mg/mL) during the entire course of cell differentiation. (<b>f</b>) qRT-PCR expression analysis for the indicated genes in bone marrow-derived osteoclasts at the indicated stages of cell differentiation, stimulated with TXA (1 mg/mL) for 6 h. (<b>g</b>) qRT-PCR expression analysis for the indicated genes in bone marrow-derived osteoclasts at day 5 of differentiation, stimulated with TXA (1 mg/mL) during the entire course of cell differentiation. For (<b>a</b>–<b>g</b>), <span class="html-italic">n</span> = 3–4 independent cultures per group were used. Data presented are means ± SD. Gene abbreviations: chloride channel 7 (<span class="html-italic">Clcn7</span>), nuclear factor kappa B subunit 1 (<span class="html-italic">Nfkb1</span>), receptor activator of nf-κb (<span class="html-italic">Tnfrsf11a</span>), tartrate-resistant acid phosphatase (<span class="html-italic">Acp5</span>), cathepsin K (<span class="html-italic">Ctsk</span>).</p>
Full article ">Figure 5
<p>Modulation of cytokine responses through TXA in resting and activated macrophages. (<b>a</b>) qRT-PCR expression analysis for the indicated genes in bone marrow-derived macrophages at day 7 of differentiation, stimulated with TXA (1 mg/mL) for 6 h. (<b>b</b>) qRT-PCR expression analysis for the indicated genes in bone marrow-derived macrophages at day 7 of differentiation, activated with 1 μg/mL LPS and simultaneously co-stimulated with TXA (1 mg/mL) for 6 h. (<b>c</b>) MTT proliferation assay of bone marrow-derived macrophages stimulated with the indicated concentrations of TXA. (<b>d</b>) Representative images of macrophage migration assays using Ibidi chambers of the indicated groups and time points. Scale bars = 50 μm. Blue lines indicate spreading cell fronts. The quantification of cell migration is shown on the right. For (<b>a</b>–<b>d</b>), <span class="html-italic">n</span> = 3–5 independent macrophage cultures per group were used as indicated with individual data points. Data presented are means ± SD. Gene abbreviations: interleukin-1 alpha (<span class="html-italic">Il1a</span>), interleukin-1 beta (<span class="html-italic">Il1b</span>), interleukin-4 (<span class="html-italic">Il4</span>), interleukin-6 (<span class="html-italic">Il6</span>), interleukin-10 (<span class="html-italic">Il10)</span>, tumor necrosis factor alpha (<span class="html-italic">Tnfa</span>), CC-chemokine ligand 5 (<span class="html-italic">Ccl5</span>), cluster of differentiation 14 (<span class="html-italic">Cd14</span>), interleukin-1 receptor antagonist (<span class="html-italic">Il1ra</span>), inducible NO synthase (<span class="html-italic">iNos</span>), toll-like receptor 4 (<span class="html-italic">Tlr4</span>), transforming growth factor beta (<span class="html-italic">Tgfb</span>).</p>
Full article ">
13 pages, 1860 KiB  
Article
KRAS Promoter G-Quadruplexes from Sequences of Different Length: A Physicochemical Study
by Federica D’Aria, Bruno Pagano, Luigi Petraccone and Concetta Giancola
Int. J. Mol. Sci. 2021, 22(1), 448; https://doi.org/10.3390/ijms22010448 - 5 Jan 2021
Cited by 8 | Viewed by 3563
Abstract
DNA G-quadruplexes (G4s) form in relevant genomic regions and intervene in several biological processes, including the modulation of oncogenes expression, and are potential anticancer drug targets. The human KRAS proto-oncogene promoter region contains guanine-rich sequences able to fold into G4 structures. Here, by [...] Read more.
DNA G-quadruplexes (G4s) form in relevant genomic regions and intervene in several biological processes, including the modulation of oncogenes expression, and are potential anticancer drug targets. The human KRAS proto-oncogene promoter region contains guanine-rich sequences able to fold into G4 structures. Here, by using circular dichroism and differential scanning calorimetry as complementary physicochemical methodologies, we compared the thermodynamic stability of the G4s formed by a shorter and a longer version of the KRAS promoter sequence, namely 5′-AGGGCGGTGTGGGAATAGGGAA-3′ (KRAS 22RT) and 5′-AGGGCGGTGTGGGAAGAGGGAAGAGGGGGAGG-3′ (KRAS 32R). Our results show that the unfolding mechanism of KRAS 32R is more complex than that of KRAS 22RT. The different thermodynamic stability is discussed based on the recently determined NMR structures. The binding properties of TMPyP4 and BRACO-19, two well-known G4-targeting anticancer compounds, to the KRAS G4s were also investigated. The present physicochemical study aims to help in choosing the best G4 target for potential anticancer drugs. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic representation of the Nuclease-Hypersensitive Element (NHE) region of the <span class="html-italic">KRAS</span> gene promoter.</p>
Full article ">Figure 2
<p>(<b>a</b>) Circular dichroism (CD) spectra of <span class="html-italic">KRAS</span> 22RT (black line) and <span class="html-italic">KRAS</span> 32R (red line). CD melting (black line) and annealing (red line) profiles of (<b>b</b>) <span class="html-italic">KRAS</span> 22RT and (<b>c</b>) <span class="html-italic">KRAS</span> 32R at 0.5 °C min<sup>−1</sup>. The buffer solution was 20 mM KH<sub>2</sub>PO<sub>4</sub>/K<sub>2</sub>HPO<sub>4</sub> with 60 mM KCl and 0.1 mM EDTA at pH 7.0.</p>
Full article ">Figure 3
<p>Experimental differential scanning calorimetry (DSC) profiles (black line) and van ’t Hoff calculated curves based on the two-states model (red line) for (<b>a</b>) <span class="html-italic">KRAS</span> 22RT and (<b>b</b>) <span class="html-italic">KRAS</span> 32R. The buffer solution was 20 mM KH<sub>2</sub>PO<sub>4</sub>/K<sub>2</sub>HPO<sub>4</sub> with 60 mM KCl and 0.1 mM EDTA at pH 7.0.</p>
Full article ">Figure 4
<p>DSC profile of <span class="html-italic">KRAS</span> 32R (black line) and the best fitting curve (red line) obtained by the thermodynamic model described in <a href="#sec4dot4dot1-ijms-22-00448" class="html-sec">Section 4.4.1</a>.</p>
Full article ">Figure 5
<p>CD melting curves of (<b>a</b>) <span class="html-italic">KRAS</span> 22RT and (<b>c</b>) <span class="html-italic">KRAS</span> 32R in the absence (black line) and in the presence of 1 (red line) and 2 equivalents of TMPyP4 (blue line). CD melting curves of (<b>b</b>) <span class="html-italic">KRAS</span> 22RT and (<b>d</b>) <span class="html-italic">KRAS</span> 32R in the absence (black line) and in the presence of 1 (red line) and 2 equivalents (blue line) of BRACO-19. The buffer solution was 20 mM KH<sub>2</sub>PO<sub>4</sub>/K<sub>2</sub>HPO<sub>4</sub> with 60 mM KCl and 0.1 mM EDTA at pH 7.0.</p>
Full article ">Figure 6
<p>DSC profiles of <span class="html-italic">KRAS</span> 22RT (<b>a</b>) and <span class="html-italic">KRAS</span> 32R (<b>c</b>) in the absence (black line) and presence (red line) of 1 equivalent of TMPyP4. DSC profiles of <span class="html-italic">KRAS</span> 22RT (<b>b</b>) and <span class="html-italic">KRAS</span> 32R (<b>d</b>) in the absence (black line) and presence (red line) of 1 equivalent of BRACO-19. The buffer solution was 20 mM KH<sub>2</sub>PO<sub>4</sub>/K<sub>2</sub>HPO<sub>4</sub> with 60 mM KCl and 0.1 mM EDTA at pH 7.0.</p>
Full article ">
18 pages, 5717 KiB  
Article
Evolution of A bHLH Interaction Motif
by Peter S. Millard, Birthe B. Kragelund and Meike Burow
Int. J. Mol. Sci. 2021, 22(1), 447; https://doi.org/10.3390/ijms22010447 - 5 Jan 2021
Cited by 5 | Viewed by 4095
Abstract
Intrinsically disordered proteins and regions with their associated short linear motifs play key roles in transcriptional regulation. The disordered MYC-interaction motif (MIM) mediates interactions between MYC and MYB transcription factors in Arabidopsis thaliana that are critical for constitutive and induced glucosinolate (GLS) biosynthesis. [...] Read more.
Intrinsically disordered proteins and regions with their associated short linear motifs play key roles in transcriptional regulation. The disordered MYC-interaction motif (MIM) mediates interactions between MYC and MYB transcription factors in Arabidopsis thaliana that are critical for constitutive and induced glucosinolate (GLS) biosynthesis. GLSs comprise a class of plant defense compounds that evolved in the ancestor of the Brassicales order. We used a diverse set of search strategies to discover additional occurrences of the MIM in other proteins and in other organisms and evaluate the findings by means of structural predictions, interaction assays, and biophysical experiments. Our search revealed numerous MIM instances spread throughout the angiosperm lineage. Experiments verify that several of the newly discovered MIM-containing proteins interact with MYC TFs. Only hits found within the same transcription factor family and having similar characteristics could be validated, indicating that structural predictions and sequence similarity are good indicators of whether the presence of a MIM mediates interaction. The experimentally validated MIMs are found in organisms outside the Brassicales order, showing that MIM function is broader than regulating GLS biosynthesis. Full article
(This article belongs to the Special Issue Protein Intrinsic Disorder in Plant Biology)
Show Figures

Figure 1

Figure 1
<p>Examples of putative MYC-interaction motifs (MIMs) identified throughout the angiosperms. (<b>a</b>) Alignment of MIM-containing MYB transcription factors (TFs) from <span class="html-italic">A. thaliana</span>, with numbering according to AtMYB29, and with the MIM highlighted by a green box. (<b>b</b>) Phylogenetic tree showing the evolutionary relationship of orders with putative MIM-containing proteins. The tree was built with the NCBI Taxonomy Common Tree tool (<a href="https://www.ncbi.nlm.nih.gov/taxonomy" target="_blank">https://www.ncbi.nlm.nih.gov/taxonomy</a>) and is not to scale. (<b>c</b>) Domain organization and localization of each individual protein with a putative MIM. Domain information was obtained from PROSITE [<a href="#B41-ijms-22-00447" class="html-bibr">41</a>]. The location of the putative motif is shown in green, and the core motif residues (xLNxxA) in bold. Numbers (#) refer to each putative MIM-containing protein in <a href="#ijms-22-00447-t001" class="html-table">Table 1</a>.</p>
Full article ">Figure 2
<p>Disorder prediction and IDDomainSpotter profiles for putative MIM-containing proteins. Top panels show sequence-specific disorder predictions using IUPred2 [<a href="#B49-ijms-22-00447" class="html-bibr">49</a>], DISOPRED3.1 [<a href="#B50-ijms-22-00447" class="html-bibr">50</a>], PONDR VSL2 [<a href="#B51-ijms-22-00447" class="html-bibr">51</a>,<a href="#B52-ijms-22-00447" class="html-bibr">52</a>], and ODiNPred [<a href="#B53-ijms-22-00447" class="html-bibr">53</a>]. Bottom panels show IDDomainSpotter profiles: Each curve shows the fraction of a different set of residues within sliding windows of 15 residues [<a href="#B48-ijms-22-00447" class="html-bibr">48</a>]. Domain information was obtained from PROSITE [<a href="#B41-ijms-22-00447" class="html-bibr">41</a>]. The putative motif is shown by a green bar.</p>
Full article ">Figure 3
<p>Testing interaction between <span class="html-italic">A. thaliana</span> (At) MYC3 and MYC4 against AtMYB28, AtMYB29, AtMYB75, AtMYB95, the MIM-containing trihelix MYB-like TF from <span class="html-italic">A. thaliana</span> (AtASR3), and MIM-containing R2R3 MYB TF from <span class="html-italic">Macleaya cordata</span> (McMYB). Growth on SD-LW medium confirms successful co-transformation with bait and prey construct and growth on SD-AHLW + Rapa medium indicates that bait and prey proteins are expressed and present in the same cellular compartment. The SD-AHLW-Rapa medium is selective for physical interaction of bait and prey protein. Rapa, rapamycin.</p>
Full article ">Figure 4
<p>Structure elucidation of MYB29-MIM and ASR3-MIM. <sup>13</sup>C-HSQC spectra of (<b>a</b>) MYB29-MIM and (<b>c</b>) ASR3-MIM with assignments. SCS of (<b>b</b>) MYB29-MIM and (<b>d</b>) ASR3-MIM. (<b>e</b>) Far-UV CD spectra and (<b>f</b>) HT voltage of MYB29-MIM and ASR3-MIM. (<b>g</b>) Sequence alignment of MYB29-MIM and ASR3-MIM (core motif residues in bold).</p>
Full article ">Figure 5
<p>Alignments of (<b>a</b>) trihelix TFs and (<b>b</b>) R2R3 MYB TFs colored according to conservation, performed using BoxShade (<a href="https://embnet.vital-it.ch/software/BOX_form.html" target="_blank">https://embnet.vital-it.ch/software/BOX_form.html</a>). For each position where a consensus (&gt;50%) could be defined, residues identical to this consensus were colored black, while those similar were colored gray. A green border and box highlight the putative motifs. Vv: <span class="html-italic">Vitis vinifera</span>, At: <span class="html-italic">A. thaliana</span>, Tc: <span class="html-italic">Theobroma cacao</span>, Coc: <span class="html-italic">Corchorus capsularis</span>, Gr: <span class="html-italic">Gossypium raimondii</span>, Eug: <span class="html-italic">Eucalyptus grandis</span>, Jr: <span class="html-italic">Juglans regia</span>, Qs: <span class="html-italic">Quercus suber</span>, To: <span class="html-italic">Trema orientale</span>, Mn: <span class="html-italic">Morus notabilis</span>, Cs: <span class="html-italic">Cucumis sativus</span>, Pv: <span class="html-italic">Phaseolus vulgaris</span>, Gm: <span class="html-italic">Glycine max</span>, Erg: <span class="html-italic">Erythranthe guttata</span>, Sp: <span class="html-italic">Solanum pennellii</span>, Cac: <span class="html-italic">Capsicum chinense</span>, Mc: <span class="html-italic">Macleaya cordata</span>, Ta: <span class="html-italic">Triticum aestivum</span>, Zm: <span class="html-italic">Zea mays</span>, Bd: <span class="html-italic">Brachypodium distachyon</span>, Ac: <span class="html-italic">Ananas comosus</span>, Hv: <span class="html-italic">Hordeum vulgare</span>, Hi: <span class="html-italic">Handroanthus impetiginosus</span>.</p>
Full article ">
12 pages, 1361 KiB  
Article
A Theoretical and Experimental Study on the Potential Luminescent and Biological Activities of Diaminodicyanoquinodimethane Derivatives
by Edison Rafael Jiménez, Manuel Caetano, Nelson Santiago, F. Javier Torres, Thibault Terencio and Hortensia Rodríguez
Int. J. Mol. Sci. 2021, 22(1), 446; https://doi.org/10.3390/ijms22010446 - 5 Jan 2021
Cited by 2 | Viewed by 3412
Abstract
Recently, several studies have demonstrated that diaminodicyanoquinone derivatives (DADQs) could present interesting fluorescence properties. Furthermore, some DADQs under the solid state are capable of showing quantum yields that can reach values of 90%. Besides, the diaminodiacyanoquinone core represents a versatile building block propense [...] Read more.
Recently, several studies have demonstrated that diaminodicyanoquinone derivatives (DADQs) could present interesting fluorescence properties. Furthermore, some DADQs under the solid state are capable of showing quantum yields that can reach values of 90%. Besides, the diaminodiacyanoquinone core represents a versatile building block propense either to modification or integration into different systems to obtain and provide them unique photophysical features. Herein, we carried out a theoretical study on the fluorescence properties of three different diaminodicyanoquinodimethane systems. Therefore, time-dependent density functional theory (TD-DFT) was used to obtain the values associated with the dipole moments, oscillator strengths, and the conformational energies between the ground and the first excited states of each molecule. The results suggest that only two of the three studied systems possess significant luminescent properties. In a further stage, the theoretical insights were confirmed by means of experimental measurements, which not only retrieved the photoluminescence of the DADQs, but also suggest a preliminary and promising antibacterial activity of these systems. Full article
(This article belongs to the Section Molecular Biophysics)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Molecular structure of the target molecules.</p>
Full article ">Figure 2
<p>Comparison between the optimized ground state (blue) and excited state (red) structures. (<b>a</b>) frontal view of molecule 1; (<b>b</b>) lateral view of molecule 1; (<b>c</b>) frontal view of molecule 2; (<b>d</b>) lateral view of molecule 2; (<b>e</b>) frontal view of molecule 3; (<b>f</b>) lateral view of molecule 3.</p>
Full article ">Figure 3
<p>Synthetic route starting with the activation of TCNQ, to obtain molecules <b>1</b> and <b>2</b>.</p>
Full article ">Figure 4
<p>Recorded emission spectra in all the visible range using a diode laser class 3R of 405 nm. (<b>a</b>) Emission spectra and its proper deconvolution of molecule <b>1</b>; (<b>b</b>) Emission spectra and its proper deconvolution of molecule <b>2</b>.</p>
Full article ">Figure 5
<p>Growth curves for the OD<sub>600</sub> tests performed to five different test tubes, blank (growth curve under standard conditions); Solvent (growth curve under the presence of the used solvent 5:5 water/DMSO); Comp.b (growth curve under the presence of PTCNQ) ); Comp.1 (growth curve under the presence of compound <b>1</b>); Comp.2 (growth curve under the presence of compound <b>2</b>); Comp c (growth curve under the presence of compound N-(4-nitrophenyl)ethylenediamine)).</p>
Full article ">
14 pages, 1821 KiB  
Article
Switching between Ultrafast Pathways Enables a Green-Red Emission Ratiometric Fluorescent-Protein-Based Ca2+ Biosensor
by Longteng Tang, Shuce Zhang, Yufeng Zhao, Nikita D. Rozanov, Liangdong Zhu, Jiahui Wu, Robert E. Campbell and Chong Fang
Int. J. Mol. Sci. 2021, 22(1), 445; https://doi.org/10.3390/ijms22010445 - 5 Jan 2021
Cited by 11 | Viewed by 4654
Abstract
Ratiometric indicators with long emission wavelengths are highly preferred in modern bioimaging and life sciences. Herein, we elucidated the working mechanism of a standalone red fluorescent protein (FP)-based Ca2+ biosensor, REX-GECO1, using a series of spectroscopic and computational methods. Upon 480 nm [...] Read more.
Ratiometric indicators with long emission wavelengths are highly preferred in modern bioimaging and life sciences. Herein, we elucidated the working mechanism of a standalone red fluorescent protein (FP)-based Ca2+ biosensor, REX-GECO1, using a series of spectroscopic and computational methods. Upon 480 nm photoexcitation, the Ca2+-free biosensor chromophore becomes trapped in an excited dark state. Binding with Ca2+ switches the route to ultrafast excited-state proton transfer through a short hydrogen bond to an adjacent Glu80 residue, which is key for the biosensor’s functionality. Inspired by the 2D-fluorescence map, REX-GECO1 for Ca2+ imaging in the ionomycin-treated human HeLa cells was achieved for the first time with a red/green emission ratio change (ΔR/R0) of ~300%, outperforming many FRET- and single FP-based indicators. These spectroscopy-driven discoveries enable targeted design for the next-generation biosensors with larger dynamic range and longer emission wavelengths. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Steady-state electronic spectra of REX-GECO1. The pH-dependent (pH = 3.5–11) UV/Visible absorption spectra of the Ca<sup>2+</sup> (<b>a</b>) free and (<b>b</b>) bound biosensors exhibit different patterns. The horizontal dashed line in (<b>b</b>) highlights spectral changes across pH = 5. (<b>c</b>) Modeled chromophore (CRO) and key residues inside the Ca<sup>2+</sup>-bound REX-GECO1 biosensor. 2D-fluorescence spectra for the Ca<sup>2+</sup> (<b>d</b>) free and (<b>e</b>) bound REX-GECO1 in pH = 7 buffer show considerable differences. The ground-state electronic absorption spectra are displayed vertically in white curves.</p>
Full article ">Figure 2
<p>Femtosecond transient absorption (fs-TA) spectroscopy elucidates the excited-state pathways of REX-GECO1. 2D-contour plots for the Ca<sup>2+</sup> (<b>a</b>) free and (<b>b</b>) bound biosensors in pH = 7 buffer upon 480 nm excitation show prominent electronic bands that rise and decay within hundreds of picoseconds. Intensity dynamics of representative 10-nm band regions are correspondingly plotted in (<b>c</b>,<b>d</b>), respectively.</p>
Full article ">Figure 3
<p>Quantum calculations (<b>a</b>) and molecular dynamics simulations (<b>b</b>) indicate a more twisted structure of the Ca<sup>2+</sup> free than bound REX-GECO1 in the excited and ground states. Two key dihedral angles of the chromophore are shown in the (<b>a</b>) inset. On the density contour plots in (<b>b</b>), angles from the Ca<sup>2+</sup>-bound R-GECO1 (progenitor of REX-GECO1) crystal structure are marked (black crosses).</p>
Full article ">Figure 4
<p>Emission-ratiometric imaging of the REX-GECO1 biosensor. (<b>a</b>) Confocal images of HeLa cells expressing REX-GECO1 before (low Ca<sup>2+</sup>) and after (high Ca<sup>2+</sup>) ionomycin treatment. Scale bar = 100 μm. (<b>b</b>) Time-lapse Ca<sup>2+</sup> imaging shown by the red/green fluorescence intensity ratio change (ΔR/R<sub>0</sub>) after 480 nm excitation. Gray lines: single-cell traces. Red line with pink shade: mean ± s.e.m.</p>
Full article ">
11 pages, 2405 KiB  
Article
PACAP for Retinal Health: Model for Cellular Aging and Rescue
by Etelka Pöstyéni, Andrea Kovács-Valasek, Viktória Dénes, Adrienn Mester, György Sétáló, Jr. and Róbert Gábriel
Int. J. Mol. Sci. 2021, 22(1), 444; https://doi.org/10.3390/ijms22010444 - 5 Jan 2021
Cited by 12 | Viewed by 2598
Abstract
Retinal aging is the result of accumulating molecular and cellular damage with a manifest decline in visual functions. Somatostatin (SST) and pituitary adenylate cyclase-activating polypeptide (PACAP) have been implicated in neuroprotection through regulating disparate aspects of neuronal activity (survival, proliferation and renewal). The [...] Read more.
Retinal aging is the result of accumulating molecular and cellular damage with a manifest decline in visual functions. Somatostatin (SST) and pituitary adenylate cyclase-activating polypeptide (PACAP) have been implicated in neuroprotection through regulating disparate aspects of neuronal activity (survival, proliferation and renewal). The aim of the present study was to validate a transgenic model for SST-expressing amacrine cells and to investigate the chronic effect of PACAP on the aging of SSTergic and dopaminergic cells of the retina. SST-tdTomato transgenic mice that were 6, 12 and 18 months old were treated intravitreally with 100 pmol of PACAP every 3 months. The density of SST and dopaminergic amacrine cells was assessed in whole-mounted retinas. Cells displaying the transgenic red fluorescence were identified as SST-immunopositive amacrine cells. By comparing the three age groups. PACAP treatment was shown to induce a moderate elevation of cell densities in both the SST and dopaminergic cell populations in the 12- and 18-month-old animals. By contrast, the control untreated and saline-treated retinas showed a minor cell loss. In conclusion, we report a reliable transgenic model for examining SSTergic amacrine cells. The fundamental novelty of this study is that PACAP could increase the cell density in matured retinal tissue, anticipating new therapeutic potential in age-related pathological processes. Full article
(This article belongs to the Special Issue Peptides for Health Benefits 2020)
Show Figures

Figure 1

Figure 1
<p>SST and TH immunohistochemistry in retinal whole mounts and sections in 6-month-old mouse retina. tdTomato-expressing cells (<b>a</b>) and TH-immunopositive neurons (<b>b</b>) in retinal whole mount. Co-localization of anti-SST antibody (labeled with AF 488-green) and red autofluorescent cells appeared as yellow in the merged image (<b>c</b>). Co-localization was not observed between SST cells (red) and TH immunopositive cells (green) (<b>d</b>). INL—Inner nuclear layer; IPL—Inner plexiform layer. Scale bar in (<b>a</b>) is valid for (<b>b</b>).</p>
Full article ">Figure 2
<p>Distributions of SST (red) and TH (green) cells are shown in central (<b>a</b>,<b>b</b>) and peripheral (<b>c</b>,<b>d</b>) retina of 6-month-old mice as well as in saline-treated (<b>a</b>,<b>c</b>) and PACAP-treated (<b>b</b>,<b>d</b>) whole mounts. Scale bar is identical for all pictures.</p>
Full article ">Figure 3
<p>Distributions of SST (red) and TH (green) cells are shown in central (<b>a</b>,<b>b</b>) and peripheral (<b>c</b>,<b>d</b>) retina of 12-month-old mice. An increase in the density of both cell populations is clearly seen in the PACAP-treated (<b>b</b>,<b>d</b>) whole mount compared to saline-treated (<b>a</b>,<b>c</b>) retina. Scale bar is identical for all pictures.</p>
Full article ">Figure 4
<p>Distributions of SST (red) and TH (green) cells are shown in central (<b>a</b>,<b>b</b>) and peripheral (<b>c</b>,<b>d</b>) retina of 18-month-old mice. The increase in the SST and TH cell density is even more enhanced in the PACAP-treated (<b>b</b>,<b>d</b>) whole mount compared to saline-treated (<b>a</b>,<b>c</b>) retina. The arrows point to tdTomato-expressing glia-like cells. Scale bar is identical for all pictures.</p>
Full article ">Figure 5
<p>The SST (<b>a</b>,<b>c</b>)- and TH (<b>b</b>,<b>d</b>)-positive cell densities in central (<b>a</b>,<b>b</b>) and peripheral retinas (<b>c</b>,<b>d</b>) in different groups. Data presented as mean ± SEM, where * means <span class="html-italic">p</span> &lt; 0.05 and ** means <span class="html-italic">p</span> &lt; 0.01, compared to the cell densities of the 6-month-old PACAP-treated group.</p>
Full article ">
16 pages, 2248 KiB  
Article
Temporomandibular Joint Osteoarthritis: Regenerative Treatment by a Stem Cell Containing Advanced Therapy Medicinal Product (ATMP)—An In Vivo Animal Trial
by Robert Köhnke, Marcus Oliver Ahlers, Moritz Alexander Birkelbach, Florian Ewald, Michael Krueger, Imke Fiedler, Björn Busse, Max Heiland, Tobias Vollkommer, Martin Gosau, Ralf Smeets and Rico Rutkowski
Int. J. Mol. Sci. 2021, 22(1), 443; https://doi.org/10.3390/ijms22010443 - 5 Jan 2021
Cited by 22 | Viewed by 5796
Abstract
Temporomandibular joint osteoarthritis (TMJ-OA) is a chronic degenerative disease that is often characterized by progressive impairment of the temporomandibular functional unit. The aim of this randomized controlled animal trial was a comparative analysis regarding the chondroregenerative potency of intra-articular stem/stromal cell therapy. Four [...] Read more.
Temporomandibular joint osteoarthritis (TMJ-OA) is a chronic degenerative disease that is often characterized by progressive impairment of the temporomandibular functional unit. The aim of this randomized controlled animal trial was a comparative analysis regarding the chondroregenerative potency of intra-articular stem/stromal cell therapy. Four weeks after combined mechanical and biochemical osteoarthritis induction in 28 rabbits, therapy was initiated by a single intra-articular injection, randomized into the following groups: Group 1: AB Serum (ABS); Group 2: Hyaluronic acid (HA); Group 3: Mesenchymal stromal cells (STx.); Group 4: Mesenchymal stromal cells in hyaluronic acid (HA + STx.). After another 4 weeks, the animals were euthanized, followed by histological examination of the removed joints. The histological analysis showed a significant increase in cartilage thickness in the stromal cell treated groups (HA + STx. vs. ABS, p = 0.028; HA + ST.x vs. HA, p = 0.042; STx. vs. ABS, p = 0.036). Scanning electron microscopy detected a similar heterogeneity of mineralization and tissue porosity in the subchondral zone in all groups. The single intra-articular injection of a stem cell containing, GMP-compliant advanced therapy medicinal product for the treatment of iatrogen induced osteoarthritis of the temporomandibular joint shows a chondroregenerative effect. Full article
Show Figures

Figure 1

Figure 1
<p>To objectify the cartilage thickness measurements, a line grid (black) was used to define random measuring points on the cartilage surface. At these points the thickness measurement (yellow) was finally taken, each right angled to an imaginary tangent (dotted).</p>
Full article ">Figure 2
<p>Boxplot analysis showing the underlying distribution of final cartilage thickness vs. the four treatment groups. Circles represent outliers with more than 1.5 times interquartile range. Significant group differences are marked (***).</p>
Full article ">Figure 3
<p>(<b>A</b>–<b>D</b>). Histological analysis 4 weeks after intra-articular injection therapy (1 = Safranin-O, 2 = Picrosirius red, 3 = Picrosirius red (polarized)). (<b>A1</b>–<b>A3</b>) representing group 1 (ABS), (<b>B1</b>–<b>B3</b>) group 2 (HA), (<b>C1</b>–<b>C3</b>) group 3 (STx.), and (<b>D1</b>–<b>D3</b>) group 4 (STx. + HA). Collagen I on cartilage surface could be visualized much better in group 3 (STx.) than in the other groups (Picrosirius red: red; Picrosirius red (polarized): yellow-orange), indicating a comparatively higher cartilage regeneration. A statistical significance analysis was not performed. Overall, no significant increased or decreased GAG accumulation could be detected in any of the groups using Safranin-O staining.</p>
Full article ">Figure 4
<p>Representative back-scattered images of temporomandibular joint (TMJ) specimens of all groups. (<b>A</b>–<b>D</b>) The images show a cross-sectional view of the TMJ, whereby darker pixels correspond to a low degree of mineralization and brighter pixels correspond to higher degree of mineralization. The articular cartilage (AC) is located on top of the subchondral mineralized zone (SMZ) which is composed of calcified cartilage containing chondrocyte lacunae and subchondral bone containing osteocyte lacunae. (<b>E</b>) In the calcified cartilage zone, the ratio of mineralized area per tissue area, an inverse measure of tissue porosity, was similar between the groups. (<b>F</b>) The heterogeneity of mineralization, assessed as the full-width-half-maximum of the gray value histogram, was similar between groups.</p>
Full article ">
Previous Issue
Back to TopTop