[go: up one dir, main page]

Next Issue
Volume 21, October-2
Previous Issue
Volume 21, September-2
 
 
ijms-logo

Journal Browser

Journal Browser

Int. J. Mol. Sci., Volume 21, Issue 19 (October-1 2020) – 454 articles

Cover Story (view full-size image): Extracellular vesicles (EVs) represent a new reality for many physiological and pathological functions as an alternative mode of intercellular communication. Here, we provide an overview of studies showing that EVs released from blood–brain barrier (BBB) endothelial cells, platelets, leukocytes, myeloid cells, astrocytes, and oligodendrocytes are involved in the pathogenesis of MS and of its rodent model experimental autoimmune encephalomyelitis. Most of the information points to an impact of EVs on BBB damage, on spreading pro-inflammatory signals, and altering neuronal functions, but EVs reparative function of brain damage deserves attention. Finally, we will describe recent advances about EVs as potential therapeutic targets and tools for therapeutic intervention in MS. View this paper
  • 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, 5016 KiB  
Article
The Discovery of Highly Potent THP Derivatives as OCTN2 Inhibitors: From Structure-Based Virtual Screening to In Vivo Biological Activity
by Francesca Di Cristo, Anna Calarco, Filomena Anna Digilio, Maria Stefania Sinicropi, Camillo Rosano, Umberto Galderisi, Mariarosa Anna Beatrice Melone, Carmela Saturnino and Gianfranco Peluso
Int. J. Mol. Sci. 2020, 21(19), 7431; https://doi.org/10.3390/ijms21197431 - 8 Oct 2020
Cited by 8 | Viewed by 3093
Abstract
A mismatch between β-oxidation and the tricarboxylic acid cycle (TCA) cycle flux in mitochondria produces an accumulation of lipid metabolic intermediates, resulting in both blunted metabolic flexibility and decreased glucose utilization in the affected cells. The ability of the cell to switch to [...] Read more.
A mismatch between β-oxidation and the tricarboxylic acid cycle (TCA) cycle flux in mitochondria produces an accumulation of lipid metabolic intermediates, resulting in both blunted metabolic flexibility and decreased glucose utilization in the affected cells. The ability of the cell to switch to glucose as an energy substrate can be restored by reducing the reliance of the cell on fatty acid oxidation. The inhibition of the carnitine system, limiting the carnitine shuttle to the oxidation of lipids in the mitochondria, allows cells to develop a high plasticity to metabolic rewiring with a decrease in fatty acid oxidation and a parallel increase in glucose oxidation. We found that 3-(2,2,2-trimethylhydrazine)propionate (THP), which is able to reduce cellular carnitine levels by blocking both carnitine biosynthesis and the cell membrane carnitine/organic cation transporter (OCTN2), was reported to improve mitochondrial dysfunction in several diseases, such as Huntington’s disease (HD). Here, new THP-derived carnitine-lowering agents (TCL), characterized by a high affinity for the OCTN2 with a minimal effect on carnitine synthesis, were developed, and their biological activities were evaluated in both in vitro and in vivo HD models. Certain compounds showed promising biological activities: reducing protein aggregates in HD cells, ameliorating motility defects, and increasing the lifespan of HD Drosophila melanogaster. Full article
(This article belongs to the Section Molecular Biology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>A</b>) Binding modes of <span class="html-small-caps">l</span>-carnitine (white sticks) THP (yellow sticks) resulting from docking simulations. Residues involved in protein‒ligand interactions are drawn as balls and sticks; the receptor hOCTN2 is reported in tan ribbons; (<b>B</b>) structural modifications carried out on THP; (<b>C</b>) synthesis of TCL <b>4a</b>–<b>t</b>. (a) HCOH 37%, HCOOH, 60 °C, 16 h; (b) for <b>3a</b>-<b>j</b> and <b>3l</b>-<b>t</b>: amine, DCC, HOBT, Et<sub>3</sub>N, CH<sub>2</sub>Cl<sub>2</sub>, or THF, RT, 24 h, for <b>3k</b>: 4-nitroaniline, PCl<sub>3</sub>, pyridine, 3 h, 40 °C; (c) CH<sub>3</sub>I, acetone, 18 h, RT. Abbreviations: 3-(2,2,2-trimethylhydrazine)propionate, THP; THP structurally related compounds, TCL; hydroxymethylene, HCOH; formic acid, HCOOH; <span class="html-italic">N</span>,<span class="html-italic">N</span>’- dicyclohexylcarbodiimide, DCC; hydroxybenzotriazole, HOBT; <span class="html-italic">N</span>,<span class="html-italic">N</span>-diethylethanamine, Et<sub>3</sub>N; dichloromethane, CH<sub>2</sub>Cl<sub>2</sub>; tetrahydrofurane, THF; room temperature, RT; phosphorus trichloride, iodomethane, PCl<sub>3</sub>.</p>
Full article ">Figure 2
<p>(<b>a</b>,<b>b</b>) Western blot analysis of hOCTN2 expression in STHdh<sup>Q7/7</sup>, STHdh<sup>Q111/111</sup>, and transfected STHdh<sup>Q111/111</sup> (STHdh<sup>Q111/111</sup>/hOCTN2). <sup>§§</sup> <span class="html-italic">p</span> &lt; 0.01 vs. STHdh<sup>Q7/7</sup>. (<b>c</b>) The total accumulation of carnitine (10 µM), THP, or TCL (50 μM) in STHdh<sup>Q111/111</sup> cells with or without the heterologous expression of hOCTN2. (<b>d</b>) The efficiency of the transport of drugs by hOCTN2 in STHdh<sup>Q111/111</sup> cells with the heterologous expression of hOCTN2. (<b>e</b>) Accumulation at 37 °C or at 4 °C of carnitine, THP, or TCL in STHdh<sup>Q111/111</sup> cells with or without the heterologous expression of hOCTN2. (<b>f</b>) Carnitine uptake in OCTN2-transfected STHdh<sup>Q111/111</sup> cells in the presence of 50 μM THP, and the newly synthesized compound. The bars represent the mean ± standard deviation (<span class="html-italic">n</span> = 3). Statistical significance: * <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 vs. CTL.</p>
Full article ">Figure 3
<p>(<b>a</b>,<b>b</b>) STHdh<sup>Q111/111</sup> treated with 50 μM of THP or selected TCL; (<b>c</b>,<b>d</b>) Bioenergetics studies of STHdh<sup>Q111/111</sup> cultured in the presence of pyruvate as a specific energetic substrate. The percent capacity to use pyruvate for respiration as well as the percent dependence on pyruvate was calculated as described in the Materials and Methods <a href="#sec4dot9-ijms-21-07431" class="html-sec">Section 4.9</a>. Each experiment was independently repeated at least three times, in separate 24-well plates with each treatment in 4–12 replicate wells.</p>
Full article ">Figure 4
<p>The binding modes of the different ligands tested, resulting from docking simulations. The receptor hOCTN2 is depicted by tan ribbons, and all the ligands are sticks. The residues involved in protein‒ligand interactions are drawn as shapes and sticks. (<b>a</b>,<b>b</b>) The binding mode for <b>4i</b> (pink sticks) and <b>4j</b> (light blue sticks); <b>4k</b> is in green in (<b>c</b>); (<b>d</b>) molecule <b>4t</b> is shown in orange.</p>
Full article ">Figure 5
<p>Representative western blot of mHtt aggregates in the STHdh<sup>Q111/111</sup> cell line incubated for 72 h in the presence of of 50 μM <b>4i</b>, <b>4j</b>, <b>4k</b>, or <b>4t</b>. Densitometric quantification was performed on three different experiments, and the results are expressed as the mean of the values obtained (mean ± SD). ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001 versus control.</p>
Full article ">Figure 6
<p>The effects of THP-derived molecules <b>4i</b>, <b>4j</b>, <b>4k</b>, and <b>4t</b> on fly lifespan and motor dysfunction. (<b>a</b>–<b>d</b>) Lifespan curves of transgenic UAS-Q128HD-FL/Elav-GAL4 flies fed with diets supplemented with the tested molecules. For each molecule, the comparison of age-dependent survival curves was done both with respect to the solvent alone and to THP. All treated groups had increased lifespans compared to both the control group and the THP group. (<b>a</b>) A longer extension in survival with respect to THP was observed in response to <b>4i</b>, (<span class="html-italic">p</span> &lt; 0.001), with a 44.95% increase in the mean lifespan. (<b>b</b>) A dramatic comparable increase was obtained after treatment with <b>4j</b> (<span class="html-italic">p</span> &lt; 0.001), with an increase in the mean lifespan of 25.3%. (<b>c</b>) A moderate but highly significant increase in survival was observed in response to <b>4k</b> (<span class="html-italic">p</span> &lt; 0.001; mean lifespan +16.96%). (<b>d</b>) A minor increase in survival, although highly significant, was observed in response to <b>4t</b> (<span class="html-italic">p</span> &lt; 0.01; mean lifespan +9.18%). <span class="html-italic">n</span> = 100 flies. (<b>e</b>) The climbing ability of UAS-Q128HD-FL/Elav-GAL4 transgenic flies fed with different media supplemented with one of four different derived-THP molecules, with THP, or with vehicle alone for control, was evaluated. All the tested THP-derived molecules improved the climbing ability of transgenic flies. The data are given as the mean ± SD. <span class="html-italic">n</span> = 60. <span class="html-italic">p</span> &lt; 0.001 compared with THP for <b>4j</b>, <b>4i</b>, <b>4k</b>, and <b>4t</b>.</p>
Full article ">
16 pages, 1153 KiB  
Review
Helicobacter pylori Virulence Factor Cytotoxin-Associated Gene A (CagA)-Mediated Gastric Pathogenicity
by Shamshul Ansari and Yoshio Yamaoka
Int. J. Mol. Sci. 2020, 21(19), 7430; https://doi.org/10.3390/ijms21197430 - 8 Oct 2020
Cited by 63 | Viewed by 7460
Abstract
Helicobacter pylori causes persistent infection in the gastric epithelium of more than half of the world’s population, leading to the development of severe complications such as peptic ulcer diseases, gastric cancer, and gastric mucosa-associated lymphoid tissue (MALT) lymphoma. Several virulence factors, including cytotoxin-associated [...] Read more.
Helicobacter pylori causes persistent infection in the gastric epithelium of more than half of the world’s population, leading to the development of severe complications such as peptic ulcer diseases, gastric cancer, and gastric mucosa-associated lymphoid tissue (MALT) lymphoma. Several virulence factors, including cytotoxin-associated gene A (CagA), which is translocated into the gastric epithelium via the type 4 secretory system (T4SS), have been indicated to play a vital role in disease development. Although infection with strains harboring the East Asian type of CagA possessing the EPIYA-A, -B, and -D sequences has been found to potentiate cell proliferation and disease pathogenicity, the exact mechanism of CagA involvement in disease severity still remains to be elucidated. Therefore, we discuss the possible role of CagA in gastric pathogenicity. Full article
(This article belongs to the Special Issue Microbial Virulence Factors 2.0)
Show Figures

Figure 1

Figure 1
<p>Function of different Cag proteins. The respective function of the Cag protein component is shown by a black circle [<a href="#B16-ijms-21-07430" class="html-bibr">16</a>,<a href="#B18-ijms-21-07430" class="html-bibr">18</a>].</p>
Full article ">Figure 2
<p>CagA after its synthesis is translocated into the gastric epithelial cells via T4SS, which forms a syringe-like structure [<a href="#B16-ijms-21-07430" class="html-bibr">16</a>,<a href="#B18-ijms-21-07430" class="html-bibr">18</a>]. The functions of various Cag proteins have been described in <a href="#ijms-21-07430-f001" class="html-fig">Figure 1</a>.</p>
Full article ">
17 pages, 3604 KiB  
Article
Vitamin E Is Superior to Vitamin C in Delaying Seedling Senescence and Improving Resistance in Arabidopsis Deficient in Macro-Elements
by Zhong-Wei Zhang, Xin-Yue Yang, Xiao-Jian Zheng, Yu-Fan Fu, Ting Lan, Xiao-Yan Tang, Chang-Quan Wang, Guang-Deng Chen, Jian Zeng and Shu Yuan
Int. J. Mol. Sci. 2020, 21(19), 7429; https://doi.org/10.3390/ijms21197429 - 8 Oct 2020
Cited by 7 | Viewed by 3042
Abstract
Nitrogen (N), phosphorus (P), and potassium (K) are three essential macro-elements for plant growth and development. Used to improve yield in agricultural production, the excessive use of chemical fertilizers often leads to increased production costs and ecological environmental pollution. Vitamins C and E [...] Read more.
Nitrogen (N), phosphorus (P), and potassium (K) are three essential macro-elements for plant growth and development. Used to improve yield in agricultural production, the excessive use of chemical fertilizers often leads to increased production costs and ecological environmental pollution. Vitamins C and E are antioxidants that play an important role in alleviating abiotic stress. However, there are few studies on alleviating oxidative stress caused by macro-element deficiency. Here, we used Arabidopsis vitamin E synthesis-deficient mutant vte4 and vitamin C synthesis-deficient mutant vtc1 on which exogenous vitamin E and vitamin C, respectively, were applied at the bolting stage. In the deficiency of macro-elements, the Arabidopsis chlorophyll content decreased, malondialdehyde (MDA) content and relative electric conductivity increased, and reactive oxygen species (ROS) accumulated. The mutants vtc1 and vte4 are more severely stressed than the wild-type plants. Adding exogenous vitamin E was found to better alleviate stress than adding vitamin C. Vitamin C barely affected and vitamin E significantly inhibited the synthesis of ethylene (ETH) and jasmonic acid (JA) genes, thereby reducing the accumulation of ETH and JA that alleviated the senescence caused by macro-element deficiency at the later stage of bolting in Arabidopsis. A deficiency of macro-elements also reduced the yield and germination rate of the seeds, which were more apparent in vtc1 and vte4, and adding exogenous vitamin C and vitamin E, respectively, could restore them. This study reported, for the first time, that vitamin E is better than vitamin C in delaying seedling senescence caused by macro-element deficiency in Arabidopsis. Full article
Show Figures

Figure 1

Figure 1
<p>Leaf phenotypes and chlorophyll contents of <span class="html-italic">Arabidopsis thaliana</span> at the 35th day (<b>A</b>) and the 45th day (<b>B</b>) after germination. CK, DN, DP, DK represent normal, deficiency of nitrogen, deficiency of phosphorus, and deficiency of potassium treatments, respectively. <span class="html-italic">vtc1</span> and <span class="html-italic">vte4</span> are vitamin C and vitamin E synthetic deletion mutants, respectively. WT+VC and WT+VE represent treatments of 5 mM exogenous vitamin C and vitamin E, respectively, to the wild-type (WT) plants. Bar = 1 cm. FW, fresh weight. The data represent average values ± SEM (<span class="html-italic">n</span> = 3). Different small letters show significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>Seed pod phenotypes and chlorophyll contents of <span class="html-italic">Arabidopsis thaliana</span> at the 35th day (<b>A</b>) and the 45th day (<b>B</b>) after germination. CK, DN, DP, DK represent normal, deficiency of nitrogen, deficiency of phosphorus, and deficiency of potassium treatments, respectively. <span class="html-italic">vtc1</span> and <span class="html-italic">vte4</span> are vitamin C and vitamin E synthetic deletion mutants, respectively. WT+VC and WT+VE represent treatments of 5 mM exogenous vitamin C and vitamin E, respectively, to the wild-type (WT) plants. Bar = 1 cm. FW, fresh weight. The data represent average values ± SEM (<span class="html-italic">n</span> = 3). Different small letters show significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>Malondialdehyde (MDA) content (<b>A</b>) and relative electric conductivity (<b>B</b>) of 45-day-old seedlings (20 days of nutrient deficiency treatments). CK, DN, DP, DK represent normal, deficiency of nitrogen, deficiency of phosphorus, and deficiency of potassium treatments, respectively. <span class="html-italic">vtc1</span> and <span class="html-italic">vte4</span> are vitamin C and vitamin E synthetic deletion mutants, respectively. WT+VC and WT+VE represent treatments of 5 mM exogenous vitamin C and vitamin E, respectively, to the wild-type (WT) plants. FW, fresh weight. The data represent average values ± SEM (<span class="html-italic">n</span> = 3). Different small letters show significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>O<sub>2</sub><sup>−</sup> (<b>A</b>) and H<sub>2</sub>O<sub>2</sub> (<b>B</b>) accumulation levels in 45-day-old seedlings (20 days of nutrient deficiency treatments). Histochemical assays for superoxide anion radicals (O<sub>2</sub><sup>−</sup>) and H<sub>2</sub>O<sub>2</sub> were performed by nitro-blue tetrazolium (NBT) and 3,3’-diamino-benzidine (DAB) staining, respectively. Quantitative data are presented below the staining images. CK, DN, DP, DK represent normal, deficiency of nitrogen, deficiency of phosphorus, and deficiency of potassium treatments, respectively. <span class="html-italic">vtc1</span> and <span class="html-italic">vte4</span> are vitamin C and vitamin E synthetic deletion mutants, respectively. WT+VC and WT+VE represent treatments of 5 mM exogenous vitamin C and vitamin E, respectively, to the wild-type (WT) plants. FW, fresh weight. The data represent average values ± SEM (<span class="html-italic">n</span> = 3). Different small letters show significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Relative expression of vitamin C synthetic genes <span class="html-italic">VTC1</span> (<b>A</b>) and <span class="html-italic">VTC2</span> (<b>B</b>) and vitamin E synthetic genes <span class="html-italic">VTE4</span> (<b>C</b>) and <span class="html-italic">GGR</span> (<b>D</b>) in 45-day-old seedlings (20 days of nutrient deficiency treatments). CK, DN, DP, DK represent normal, deficiency of nitrogen, deficiency of phosphorus, and deficiency of potassium treatments, respectively. <span class="html-italic">vtc1</span> and <span class="html-italic">vte4</span> are vitamin C and vitamin E synthetic deletion mutants, respectively. WT+VC and WT+VE represent treatments of 5 mM exogenous vitamin C and vitamin E, respectively, to the wild-type (WT) plants. VTC1, Vitamin C defective 1; VTC2, Vitamin C defective 2; VTE4, Vitamin E defective 4; GGR, Geranylgeranyl Reductase. The specific gene expression levels are represented as the percentages relatively to <span class="html-italic">ACTIN7</span> expression levels. The expression level of WT in the control sample (CK) was normalized into “1.0”. The data represent average values ± SEM (<span class="html-italic">n</span> = 3). Different small letters show significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 6
<p>Relative expression of ethylene synthetic genes <span class="html-italic">ACS2</span> (<b>A</b>) and <span class="html-italic">ACO1</span> (<b>B</b>) and ethylene-signaling genes <span class="html-italic">EIN3</span> (<b>C</b>) and <span class="html-italic">ERF1</span> (<b>D</b>) and ethylene contents (<b>E</b>) in 45-day-old seedlings. ACS2, 1-Aminocyclopropane-1 -Carboxylic acid Synthase 2; ACO1, 1-Aminocyclopropane-1-Carboxylate Oxidase 1; EIN3, Ethylene Insensitive 3; ERF1, Ethylene Response Factor 1. The expression level of WT in the control sample (CK) was normalized into “1.0”. The data represent average values ± SEM (<span class="html-italic">n</span> = 3). Different small letters show significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 7
<p>Relative expression of jasmonic acid synthetic gene <span class="html-italic">COI1</span> (<b>A</b>) and jasmonic acid-signaling gene <span class="html-italic">PDF1.2</span> (<b>B</b>) and jasmonic acid contents (<b>C</b>) in 45-day-old seedlings (20 days of nutrient deficiency treatments). COL1, CONSTANS-Like 1; PDF1.2, Plant Defensin 1.2. The expression level of WT in the control sample (CK) was normalized into “1.0”. The data represent average values ± SEM (<span class="html-italic">n</span> = 3). Different small letters show significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 8
<p>Yield per plant (<b>A</b>), 1000 seed weight (<b>B</b>), pod number per plant (<b>C</b>), length of pod (<b>D</b>), and seed germination rate (<b>E</b>) of Arabidopsis plants after macro-element deficiency treatments. CK, DN, DP, DK represent normal, deficiency of nitrogen, deficiency of phosphorus, and deficiency of potassium treatments, respectively. <span class="html-italic">vtc1</span> and <span class="html-italic">vte4</span> are vitamin C and vitamin E synthetic deletion mutants, respectively. WT+VC and WT+VE represent treatments of 5 mM exogenous vitamin C and vitamin E, respectively, to the wild-type (WT) plants. The data represent average values ± SEM (<span class="html-italic">n</span> = 3). Different small letters show significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
18 pages, 1909 KiB  
Article
The Paralogue of the Intrinsically Disordered Nuclear Protein 1 Has a Nuclear Localization Sequence that Binds to Human Importin α3
by José L. Neira, Bruno Rizzuti, Ana Jiménez-Alesanco, Olga Abián, Adrián Velázquez-Campoy and Juan L. Iovanna
Int. J. Mol. Sci. 2020, 21(19), 7428; https://doi.org/10.3390/ijms21197428 - 8 Oct 2020
Cited by 7 | Viewed by 2536
Abstract
Numerous carrier proteins intervene in protein transport from the cytoplasm to the nucleus in eukaryotic cells. One of those is importin α, with several human isoforms; among them, importin α3 (Impα3) features a particularly high flexibility. The protein NUPR1L is an intrinsically disordered [...] Read more.
Numerous carrier proteins intervene in protein transport from the cytoplasm to the nucleus in eukaryotic cells. One of those is importin α, with several human isoforms; among them, importin α3 (Impα3) features a particularly high flexibility. The protein NUPR1L is an intrinsically disordered protein (IDP), evolved as a paralogue of nuclear protein 1 (NUPR1), which is involved in chromatin remodeling and DNA repair. It is predicted that NUPR1L has a nuclear localization sequence (NLS) from residues Arg51 to Gln74, in order to allow for nuclear translocation. We studied in this work the ability of intact NUPR1L to bind Impα3 and its depleted species, ∆Impα3, without the importin binding domain (IBB), using fluorescence, isothermal titration calorimetry (ITC), circular dichroism (CD), nuclear magnetic resonance (NMR), and molecular docking techniques. Furthermore, the binding of the peptide matching the isolated NLS region of NUPR1L (NLS-NUPR1L) was also studied using the same methods. Our results show that NUPR1L was bound to Imp α3 with a low micromolar affinity (~5 μM). Furthermore, a similar affinity value was observed for the binding of NLS-NUPR1L. These findings indicate that the NLS region, which was unfolded in isolation in solution, was essentially responsible for the binding of NUPR1L to both importin species. This result was also confirmed by our in silico modeling. The binding reaction of NLS-NUPR1L to ∆Impα3 showed a larger affinity (i.e., lower dissociation constant) compared with that of Impα3, confirming that the IBB could act as an auto-inhibition region of Impα3. Taken together, our findings pinpoint the theoretical predictions of the NLS region in NUPR1L and, more importantly, suggest that this IDP relies on an importin for its nuclear translocation. Full article
Show Figures

Figure 1

Figure 1
<p>Binding of intact nuclear protein 1 NUPR1L to ∆Impα3 monitored by spectroscopic techniques: (<b>A</b>) Fluorescence spectrum obtained by excitation at 295 nm of the complex between ∆Impα3 and intact NUPR1L, and addition spectrum obtained by the sum of the spectra of both isolated macromolecules. (<b>B</b>) Far-UV circular dichroism (CD) spectrum of the complex between ∆Impα3 and NUPR1L and the addition spectrum obtained by the sum of the spectra of both isolated macromolecules. (<b>C</b>) Thermal denaturations of ∆Impα3 in the presence and absence of NUPR1L followed by the changes in ellipticity at 222 nm. All experiments were carried out in phosphate buffer (50 mM, pH 7.0).</p>
Full article ">Figure 2
<p>Interaction of intact NUPR1L with both importin species as measured by fluorescence: (<b>A</b>) Titration curve monitoring the changes of Impα3 fluorescence at 330 nm in the presence of NUPR1L, after excitation at 280 nm. (<b>B</b>) Titration curve monitoring the changes of ∆Impα3 fluorescence at 330 nm in the presence of NUPR1L, after excitation at 280 nm. All experiments were carried out in phosphate buffer (50 mM, pH 7.0).</p>
Full article ">Figure 3
<p>Conformational features of isolated nuclear localization sequence (NLS)-NUPR1L in solution: (<b>A</b>) Far-UV CD spectrum of NLS-NUPR1L at 298 K in phosphate buffer (50 mM, pH 7.0). (<b>B</b>) 1D-<sup>1</sup>H-nuclear magnetic resonance (NMR) spectrum of isolated NLS-NUPR1L at 283 K and pH 7.2 (50 mM, Tris buffer). (<b>C</b>) NOE (Nuclear Overhuaser effect) diagram of isolated NLS-NUPR1L at 283 K: NOEs are classified into strong, medium, or weak, as represented by the height of the bar underneath the sequence; signal intensity was judged by visual inspection from the NOESY (Nuclear Overhuaser effect spectrosocopy) experiments. The corresponding H<sub>α</sub> NOEs with the H<sub>δ</sub> of the following proline residue are indicated by an open bar in the row corresponding to the sequential αN contacts. The dotted lines indicate NOE contacts that could not be unambiguously assigned owing to signal overlap. The numbering of residues corresponds to that of the sequence of intact NUPR1L. The symbols αN, βN, γN, and NN correspond to the sequential contacts (that is, for instance, the αN corresponds to the αN (i,i + 1) contacts).</p>
Full article ">Figure 4
<p>Binding of NLS-NUNPR1L to both importin species: (<b>A</b>) Titration curve monitoring the changes of Impα3 fluorescence at 340 nm in the presence of NLS-NUPR1L, after excitation at 280 nm. (<b>B</b>) Titration curve monitoring the changes of ∆Impα3 fluorescence at 340 nm in the presence of NLS-NUPR1L, after excitation at 280 nm. All experiments were carried out in phosphate buffer (50 mM, pH 7.0). (<b>C</b>) Calorimetric binding isotherms (ligand normalized heat effect per injection as a function of the ligand/protein molar ratio) for the interaction of NLS-NUPR1L with Impα3 (left) and ∆Impα3 (right) are shown, with the thermogram (raw thermal power data as a function of time) at the top of each panel. Binding parameters were estimated by non-linear least squares regression data analysis of the interaction isotherms applying a single ligand binding site model implemented in Origin 7.0. (OriginLab, Northampton, MA, USA).</p>
Full article ">Figure 5
<p>Binding energy of 8-residue-long fragments of NLS-NUPR1L peptide to importin: Affinity of each fragment is shown in correspondence of the two residues at the centre of each 8-residue sequence. (Inset) Difference in the binding energy between the docking pose with the highest affinity found at increasing exhaustiveness in the search versus the best pose found at any exhaustiveness value, for three 8-residue-long fragments that together span the whole sequence of the twenty-residue-long NLS-NUPR1L: (light grey) fragment RTRREQAL, (dark grey) RTNWPAPG, and (black) GHERKVAQ.</p>
Full article ">Figure 6
<p>Predicted docking poses for NUPR1L peptide on importin. (<b>A</b>) Bound conformation of the capped fragment PAPGGHER: backbone (–N–C<sup>α</sup>–C– atoms) representation of the best five docking poses on ∆Impα3. (<b>B</b>) Most favorable binding pose of the same fragment (cyan), compared with the crystallographic conformation [<a href="#B37-ijms-21-07428" class="html-bibr">37</a>] of the NLS of the Epstein–Barr virus EBNA-LP protein (purple). For clarity, H atoms and backbone O atoms are omitted. The tryptophan residues (orange) in the major NLS-binding site of importin are labeled. All images were created with PyMol [<a href="#B38-ijms-21-07428" class="html-bibr">38</a>].</p>
Full article ">
14 pages, 971 KiB  
Article
Circulating miR-99a-5p Expression in Plasma: A Potential Biomarker for Early Diagnosis of Breast Cancer
by Iris Garrido-Cano, Vera Constâncio, Anna Adam-Artigues, Ana Lameirinhas, Soraya Simón, Belen Ortega, María Teresa Martínez, Cristina Hernando, Begoña Bermejo, Ana Lluch, Paula Lopes, Rui Henrique, Carmen Jerónimo, Juan Miguel Cejalvo and Pilar Eroles
Int. J. Mol. Sci. 2020, 21(19), 7427; https://doi.org/10.3390/ijms21197427 - 8 Oct 2020
Cited by 28 | Viewed by 3579
Abstract
MicroRNAs have emerged as new diagnostic and therapeutic biomarkers for breast cancer. Herein, we analysed miR-99a-5p expression levels in primary tumours and plasma of breast cancer patients to evaluate its usefulness as a minimally invasive diagnostic biomarker. MiR-99a-5p expression levels were determined by [...] Read more.
MicroRNAs have emerged as new diagnostic and therapeutic biomarkers for breast cancer. Herein, we analysed miR-99a-5p expression levels in primary tumours and plasma of breast cancer patients to evaluate its usefulness as a minimally invasive diagnostic biomarker. MiR-99a-5p expression levels were determined by quantitative real-time PCR in three independent cohorts of patients: (I) Discovery cohort: breast cancer tissues (n = 103) and healthy breast tissues (n = 26); (II) Testing cohort: plasma samples from 105 patients and 98 healthy donors; (III) Validation cohort: plasma samples from 89 patients and 85 healthy donors. Our results demonstrated that miR-99a-5p was significantly downregulated in breast cancer tissues compared to healthy breast tissues. Conversely, miR-99a-5p levels were significantly higher in breast cancer patients than in healthy controls in plasma samples from both testing and validation cohorts, and ROC curve analysis revealed that miR-99a-5p has good diagnostic potential even to detect early breast cancer. In conclusion, miR-99a-5p’s deregulated expression distinguished healthy patients from breast cancer patients in two different types of samples (tissues and plasma). Interestingly, expression levels in plasma were significantly lower in healthy controls than in early-stage breast cancer patients. Our findings suggest circulating miR-99a-5p as a novel promising non-invasive biomarker for breast cancer detection. Full article
(This article belongs to the Section Biochemistry)
Show Figures

Figure 1

Figure 1
<p>Study design to develop a novel miRNA biomarker.</p>
Full article ">Figure 2
<p>(<b>A</b>) MiR-99a expression levels in breast cancer tissues from Cohort #1. Differential miR-99a expression levels in 103 breast cancer tissues were compared with 26 normal breast tissues. Red horizontal line: median with interquartile range. Mann–Whitney U, **** <span class="html-italic">p</span> &lt; 0.0001. (<b>B</b>) Receiver-operating characteristic (ROC) curve analysis for miR-99a expression levels in breast cancer tissue samples. (<b>C</b>) TCGA data for the expression of miR-99a-5p in normal solid tissue (<span class="html-italic">n</span> = 52) and breast primary tumour (<span class="html-italic">n</span> = 782). Expression is represented as reads per million miRNA mapped. Horizontal line: median with interquartile range. Mann–Whitney U, **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 3
<p>(<b>A</b>) Circulating miR-99a levels in cohort #2. Differential miR-99a levels in 105 plasma of breast cancer patients were compared with those of 98 healthy controls. Expression levels were significantly lower in healthy controls than in breast cancer patients. Red horizontal line: median with interquartile range. Mann–Whitney U, **** <span class="html-italic">p</span> &lt; 0.0001. (<b>B</b>) Receiver-operating characteristic (ROC) curve analysis for circulating miR-99a levels in cohort #2.</p>
Full article ">Figure 4
<p>(<b>A</b>) Circulating miR-99a levels in cohort #3. Differential miR-99a levels in plasma of 89 breast cancer patients were compared with those of 85 healthy controls. Expression levels were significantly lower in healthy controls than in breast cancer patients. Red horizontal line: median with interquartile range. Mann–Whitney U, **** <span class="html-italic">p</span> &lt; 0.0001. (<b>B</b>) Receiver-operating characteristic (ROC) curve analysis for circulating miR-99a levels in cohort #3.</p>
Full article ">Figure 5
<p>(<b>A</b>) Expression of miR-99a in early-stage breast cancer plasma. Distribution of circulating miR-99a levels in 125 plasma of early-stage breast cancer patients (stage I and II) and 193 healthy controls. Expression levels were significantly lower in healthy controls than in early-stage breast cancer patients. Horizontal line: median with interquartile range. Mann–Whitney U, **** <span class="html-italic">p</span> &lt; 0.0001. (<b>B</b>) Receiver-operating characteristic (ROC) curve analysis for circulating miR-99a levels in early-stage breast cancer patients.</p>
Full article ">
12 pages, 2915 KiB  
Article
Discovery of Novel Fetal Hemoglobin Inducers through Small Chemical Library Screening
by Giulia Breveglieri, Salvatore Pacifico, Cristina Zuccato, Lucia Carmela Cosenza, Shaiq Sultan, Elisabetta D’Aversa, Roberto Gambari, Delia Preti, Claudio Trapella, Remo Guerrini and Monica Borgatti
Int. J. Mol. Sci. 2020, 21(19), 7426; https://doi.org/10.3390/ijms21197426 - 8 Oct 2020
Cited by 1 | Viewed by 2853
Abstract
The screening of chemical libraries based on cellular biosensors is a useful approach to identify new hits for novel therapeutic targets involved in rare genetic pathologies, such as β-thalassemia and sickle cell disease. In particular, pharmacologically mediated stimulation of human γ-globin [...] Read more.
The screening of chemical libraries based on cellular biosensors is a useful approach to identify new hits for novel therapeutic targets involved in rare genetic pathologies, such as β-thalassemia and sickle cell disease. In particular, pharmacologically mediated stimulation of human γ-globin gene expression, and increase of fetal hemoglobin (HbF) production, have been suggested as potential therapeutic strategies for these hemoglobinopathies. In this article, we screened a small chemical library, constituted of 150 compounds, using the cellular biosensor K562.GR, carrying enhanced green fluorescence protein (EGFP) and red fluorescence protein (RFP) genes under the control of the human γ-globin and β-globin gene promoters, respectively. Then the identified compounds were analyzed as HbF inducers on primary cell cultures, obtained from β-thalassemia patients, confirming their activity as HbF inducers, and suggesting these molecules as lead compounds for further chemical and biological investigations. Full article
(This article belongs to the Section Molecular Pathology, Diagnostics, and Therapeutics)
Show Figures

Figure 1

Figure 1
<p>Representative examples of FACS analysis of K562.GR cells, untreated (bold solid lines) or treated (solid lines), with 150 µM HU (<b>A</b>,<b>B</b>), 10 µM compound 63 (<b>C</b>,<b>D</b>), and 10 µM compound <b>1</b> (<b>E</b>,<b>F</b>), respectively. The CheLiFe compounds were dissolved in DMSO. 30,000 cells were analyzed after five days of treatment, by detecting both enhanced green fluorescence protein (EGFP) (<b>A</b>,<b>C</b>,<b>E</b>) and RFP (<b>B</b>,<b>D</b>,<b>F</b>) fluorescence.</p>
Full article ">Figure 2
<p>Representative example of HPLC analysis after erythroid precursor cells treatment. Chromatograms obtained from HPLC assay of erythroid precursor cells of patient 6 (a <span class="html-italic">β</span>-thalassemia patient), untreated (<b>A</b>) or after 5 days of treatment with 10 µM compound 63 (<b>B</b>). The positions of fetal (HbF) and adult (HbA ed HbA<sub>2</sub>) hemoglobin peaks are indicated by arrows.</p>
Full article ">
15 pages, 2417 KiB  
Article
Phosphoglycerate Mutase 1 Prevents Neuronal Death from Ischemic Damage by Reducing Neuroinflammation in the Rabbit Spinal Cord
by Hyo Young Jung, Hyun Jung Kwon, Woosuk Kim, Kyu Ri Hahn, Seung Myung Moon, Yeo Sung Yoon, Dae Won Kim and In Koo Hwang
Int. J. Mol. Sci. 2020, 21(19), 7425; https://doi.org/10.3390/ijms21197425 - 8 Oct 2020
Cited by 10 | Viewed by 2928
Abstract
Phosphoglycerate mutase 1 (PGAM1) is a glycolytic enzyme that increases glycolytic flux in the brain. In the present study, we examined the effects of PGAM1 in conditions of oxidative stress and ischemic damage in motor neuron-like (NSC34) cells and the rabbit spinal cord. [...] Read more.
Phosphoglycerate mutase 1 (PGAM1) is a glycolytic enzyme that increases glycolytic flux in the brain. In the present study, we examined the effects of PGAM1 in conditions of oxidative stress and ischemic damage in motor neuron-like (NSC34) cells and the rabbit spinal cord. A Tat-PGAM1 fusion protein was prepared to allow easy crossing of the blood-brain barrier, and Control-PGAM1 was synthesized without the Tat peptide protein transduction domain. Intracellular delivery of Tat-PGAM1, not Control-PGAM1, was achieved in a time- and concentration-dependent manner. Immunofluorescent staining confirmed the intracellular expression of Tat-PGAM1 in NSC34 cells. Tat-PGAM1, but not Control-PGAM1, significantly alleviated H2O2-induced oxidative stress, neuronal death, mitogen-activated protein kinase, and apoptosis-inducing factor expression in NSC34 cells. After ischemia induction in the spinal cord, Tat-PGAM1 treatment significantly improved ischemia-induced neurological impairments and ameliorated neuronal cell death in the ventral horn of the spinal cord 72 h after ischemia. Tat-PGAM1 treatment significantly mitigated the ischemia-induced increase in malondialdehyde and 8-iso-prostaglandin F2α production in the spinal cord. In addition, Tat-PGAM1, but not Control-PGAM1, significantly decreased microglial activation and secretion of pro-inflammatory cytokines, such as interleukin (IL)-1β, IL-6, and tumor necrosis factor (TNF)-α induced by ischemia in the ventral horn of the spinal cord. These results suggest that Tat-PGAM1 can be used as a therapeutic agent to reduce spinal cord ischemia-induced neuronal damage by lowering the oxidative stress, microglial activation, and secretion of pro-inflammatory cytokines, such as IL-1β, IL-6, and TNF-α. Full article
(This article belongs to the Section Molecular Neurobiology)
Show Figures

Figure 1

Figure 1
<p>Intracellular delivery of Trans-activator of transcription-phosphoglycerate mutase 1 (Tat-PGAM1) and Control-PGAM1 into NSC34 cells. To assess the concentration dependence, NSC34 cells were incubated with various concentrations (0.5–5.0 μM) of Tat-PGAM1 and Control-PGAM1 proteins for 1 h. To assess the time dependence, NSC34 cells were incubated with 3 μM each of Tat-PGAM1 and Control-PGAM1 for various durations (15–60 min). Delivery was confirmed by western blot analysis for polyhistidines. The delivery of Tat-PGAM1 and Control-PGAM1 was visualized by immunohistochemical staining for polyhistidine 60 min after protein (both 3 μM) incubation. Scale bar = 50 μm. The optical densities of the polyhistidine bands and immunoreactivities were measured and the data were analyzed using two-way ANOVA followed by Bonferroni’s post hoc test (<sup>a</sup><span class="html-italic">p</span> &lt; 0.05, significantly different from the control group; <sup>b</sup><span class="html-italic">p</span> &lt; 0.05, significantly different from the Control-PGAM1 group). The bar graph represents the mean ± standard deviation (SD).</p>
Full article ">Figure 2
<p>The protective effects of Tat-PGAM1 and Control-PGAM1 against oxidative damage in NSC34 cells. Oxidative stress was induced by treatment with 1 mM H<sub>2</sub>O<sub>2</sub> for 3 h (terminal deoxynucleotidyl transferase-mediated biotinylated deoxyuridine triphosphate (dUTP) nick end labeling (TUNEL) staining), 5 h (water-soluble tetrazolium salt-1 (WST-1) assay), 10 min (2,7-dichlorofluorescein (DCF) fluorescence), or 6 h (western blot). The cells were incubated with Tat-PGAM1 and Control-PGAM1 (both 3 μM for TUNEL and DCF staining) for 60 min before treatment with 1 mM H<sub>2</sub>O<sub>2</sub> and measured with each assay kit. The cell viability, DNA fragmentation, and reactive oxygen species (ROS) formation were measured using a WST-1 assay, TUNEL staining, and DCF fluorescence, respectively. Scale bar = 50 μm. Mitogen-activated protein kinase (MAPK) pathways were confirmed by western blot analysis for c-jun N-terminal kinase (JNK), extracellular-signal-regulated kinase (ERK), p38, and their phosphorylated antibodies. In addition, apoptosis inducing factor (AIF) levels were measured in nuclear and cytosolic fraction. The cell viability was measured and the data were analyzed using two-way ANOVA followed by Bonferroni’s post hoc test (<sup>a</sup><span class="html-italic">p</span> &lt; 0.05, significantly different from the control group; and <sup>b</sup><span class="html-italic">p</span> &lt; 0.05, significantly different from the Control-PGAM1 group). The intensities of TUNEL positive cells, the DCF fluorescence, and the optical densities of the protein bands were measured and the data were analyzed using one-way ANOVA followed by Bonferroni’s post hoc test (<sup>a</sup><span class="html-italic">p</span> &lt; 0.05, significantly different from the control group; <sup>b</sup><span class="html-italic">p</span> &lt; 0.05, significantly different from the only H<sub>2</sub>O<sub>2</sub> group; and <sup>c</sup><span class="html-italic">p</span> &lt; 0.05, significantly different from the Control-PGAM1 group). The bar graph represents the mean ± standard deviation (SD).</p>
Full article ">Figure 3
<p>The protective effects of Tat-PGAM1 and Control-PGAM1 against ischemic damage in the rabbit spinal cord. Neurological scores of the control, Tat peptide-, Control-PGAM1-, and Tat-PGAM1-treated groups were measured using the modified Tarlov’s criteria at 24 and 72 h after ischemia/reperfusion. The animals were sacrificed 3 days or 7 days after ischemia for NeuN immunohistochemical staining in the lumbar segments (L<sub>6</sub>-L<sub>7</sub>) of the spinal cord. Scale bar = 100 μm. The number of NeuN-positive cells were counted in the spinal cord per section for all the groups, and the data were analyzed using one-way ANOVA followed by Bonferroni’s post hoc test (<sup>a</sup><span class="html-italic">p</span> &lt; 0.05, significantly different from the control group; <sup>b</sup><span class="html-italic">p</span> &lt; 0.05, significantly different from the Tat peptide (3 d) group; <sup>c</sup><span class="html-italic">p</span> &lt; 0.05, significantly different from the Control-PGAM1 (3 d) group; and <sup>d</sup><span class="html-italic">p</span> &lt; 0.05, significantly different from the Tat peptide (7 d) group). Th bar graph represents the mean ± standard deviation (SD).</p>
Full article ">Figure 4
<p>The anti-oxidative and anti-inflammatory effects of Tat-PGAM1 and Control-PGAM1 against ischemic damage in the rabbit spinal cord. Iba-1 immunohistochemistry was conducted to visualize the morphology of microglia/macrophages in the ventral horn of the spinal cord. Scale bar = 100 μm, 25 μm (inset image). The relative optical densities were expressed as percentile values of Iba-1 immunoreactivity versus control group per section. The lipid peroxidation and oxidized arachidonic acid were determined by MDA and 8-iso-PGF2α assay, respectively, after ischemia/reperfusion. In addition, pro-inflammatory cytokines, such as Interleukin (IL)-1β, IL-6, and tumor necrosis factor (TNF)-α were measured in the spinal cord after ischemia/reperfusion. The data were analyzed using one- or two-way ANOVA followed by Bonferroni’s post hoc test (<sup>a</sup><span class="html-italic">p</span> &lt; 0.05, significantly different from the control group; <sup>b</sup><span class="html-italic">p</span> &lt; 0.05, significantly different from the Tat peptide group; and <sup>c</sup><span class="html-italic">p</span> &lt; 0.05, significantly different from the Control-PGAM1 group). The bar graph represents the mean ± standard deviation (SD).</p>
Full article ">
26 pages, 2997 KiB  
Review
T Cell Activation Machinery: Form and Function in Natural and Engineered Immune Receptors
by Nicholas J. Chandler, Melissa J. Call and Matthew E. Call
Int. J. Mol. Sci. 2020, 21(19), 7424; https://doi.org/10.3390/ijms21197424 - 8 Oct 2020
Cited by 10 | Viewed by 8519
Abstract
The impressive success of chimeric antigen receptor (CAR)-T cell therapies in treating advanced B-cell malignancies has spurred a frenzy of activity aimed at developing CAR-T therapies for other cancers, particularly solid tumors, and optimizing engineered T cells for maximum clinical benefit in many [...] Read more.
The impressive success of chimeric antigen receptor (CAR)-T cell therapies in treating advanced B-cell malignancies has spurred a frenzy of activity aimed at developing CAR-T therapies for other cancers, particularly solid tumors, and optimizing engineered T cells for maximum clinical benefit in many different disease contexts. A rapidly growing body of design work is examining every modular component of traditional single-chain CARs as well as expanding out into many new and innovative engineered immunoreceptor designs that depart from this template. New approaches to immune cell and receptor engineering are being reported with rapidly increasing frequency, and many recent high-quality reviews (including one in this special issue) provide comprehensive coverage of the history and current state of the art in CAR-T and related cellular immunotherapies. In this review, we step back to examine our current understanding of the structure-function relationships in natural and engineered lymphocyte-activating receptors, with an eye towards evaluating how well the current-generation CAR designs recapitulate the most desirable features of their natural counterparts. We identify key areas that we believe are under-studied and therefore represent opportunities to further improve our grasp of form and function in natural and engineered receptors and to rationally design better therapeutics. Full article
(This article belongs to the Special Issue Recent Advances in T Cell Immunity)
Show Figures

Figure 1

Figure 1
<p>T cell activation following TCR recognition of stimulatory pMHC requires sensitivity enhancing co-receptor engagement of MHC (CD4 or CD8αβ) as well as co-stimulatory signals from constitutively expressed CD28 and several TCR induced co-stimulatory molecules (4-1BB depicted here). Yellow boxes represent ITAMs, green boxes represent non-ITAM stimulatory motifs. (<b>A</b>) Co-receptors CD4/CD8αβ engage MHC, dramatically increasing TCR sensitivity. (<b>B</b>) Positively charged tails interact with negatively charged lipid head groups. (<b>C</b>) Stalk cysteines facilitate interchain disulfide crosslinking. (<b>D</b>) Homo/hetero-typic TM interactions are vital to immunoreceptor assembly and function. Protein data bank (PDB) codes of structures shown in this figure: CD8αβ 2ATP, CD4/pMHC/TCRαβ 3TOE, TCR 6XJR (TCRαβ from 3TOE aligned against TCRαβ chains in 6XJR using pymol, 3TOE TCRαβ chains not shown), CD28 1YJD, 4-1BB/4-1BBL 6CPR.</p>
Full article ">Figure 2
<p>2nd Generation CAR constructs: the native receptor sequences commonly incorporated and the benefits and liabilities of those domains with regard to CAR function. Structure of the scFv domain is from PDB code 3H3B.</p>
Full article ">Figure 3
<p>Novel design approaches to improving CAR function. 3rd Gen CARs use two or more co-stimulatory tails to improve signalling outcomes (OX40, ICOS, CD3ε also used). KIR-CARs and NKG2D CARs leverage native TM interactions to constitutively recruit endogenous signaling modules DAP12 and DAP10, respectively. TCR-CAR uses a CD3ε-scFv fusion to harness the high number of ITAMs and regulatory sequences within all six CD3 tails. PDB codes of structures shown are: scFv: 3H3B, NKG2D ECD: 1MPU, TCR: 6XJR.</p>
Full article ">
5 pages, 186 KiB  
Editorial
Wheat and Barley: Acclimatization to Abiotic and Biotic Stress
by Tomasz Hura
Int. J. Mol. Sci. 2020, 21(19), 7423; https://doi.org/10.3390/ijms21197423 - 8 Oct 2020
Cited by 15 | Viewed by 3366
Abstract
Twelve articles (ten research papers and two reviews) included in the Special Issue entitled “Wheat and Barley: Acclimatization to Abiotic and Biotic Stress” are summed up here to present the latest research on the molecular background of adaptation to environmental stresses in two [...] Read more.
Twelve articles (ten research papers and two reviews) included in the Special Issue entitled “Wheat and Barley: Acclimatization to Abiotic and Biotic Stress” are summed up here to present the latest research on the molecular background of adaptation to environmental stresses in two cereal species. Crucial research results were presented and discussed, as they may be of importance in breeding aimed at increasing wheat and barley tolerance to abiotic and biotic stresses. Full article
(This article belongs to the Special Issue Wheat and Barley: Acclimatization to Abiotic and Biotic Stress)
22 pages, 3309 KiB  
Article
Comprehensive Analysis of Antibodies Induced by Vaccination with 4 Kinds of Avian Influenza H5N1 Pre-Pandemic Vaccines
by Nobuko Ohshima, Yoshitaka Iba, Ritsuko Kubota-Koketsu, Ayami Yamasaki, Keiko Majima, Gene Kurosawa, Daisuke Hirano, Shunji Yoshida, Mototaka Sugiura, Yoshizo Asano, Yoshinobu Okuno and Yoshikazu Kurosawa
Int. J. Mol. Sci. 2020, 21(19), 7422; https://doi.org/10.3390/ijms21197422 - 8 Oct 2020
Cited by 2 | Viewed by 2383
Abstract
Four kinds of avian-derived H5N1 influenza virus, A/Vietnam/1194/2004 (Clade 1), A/Indonesia/5/2005 (Clade 2.1), A/Qinghai/1A/2005 (Clade 2.2), and A/Anhui/1/2005 (Clade 2.3), have been stocked in Japan for use as pre-pandemic vaccines. When a pandemic occurs, these viruses would be used as vaccines in the [...] Read more.
Four kinds of avian-derived H5N1 influenza virus, A/Vietnam/1194/2004 (Clade 1), A/Indonesia/5/2005 (Clade 2.1), A/Qinghai/1A/2005 (Clade 2.2), and A/Anhui/1/2005 (Clade 2.3), have been stocked in Japan for use as pre-pandemic vaccines. When a pandemic occurs, these viruses would be used as vaccines in the hope of inducing immunity against the pandemic virus. We analyzed the specificity of antibodies (Abs) produced by B lymphocytes present in the blood after immunization with these vaccines. Eighteen volunteers took part in this project. After libraries of Ab-encoding sequences were constructed using blood from subjects vaccinated with these viruses, a large number of clones that encoded Abs that bound to the virus particles used as vaccines were isolated. These clones were classified into two groups according to the hemagglutination inhibition (HI) activity of the encoded Abs. While two-thirds of the clones were HI positive, the encoded Abs exhibited only restricted strain specificity. On the other hand, half of the HI-negative clones encoded Abs that bound not only to the H5N1 virus but also to the H1N1 virus; with a few exceptions, these Abs appeared to be encoded by memory B cells present before vaccination. The HI-negative clones included those encoding broadly cross-reactive Abs, some of which were encoded by non-VH1-69 germline genes. However, although this work shows that various kinds of anti-H5N1 Abs are encoded by volunteers vaccinated with pre-pandemic vaccines, broad cross-reactivity was seen only in a minority of clones, raising concern regarding the utility of these H5N1 vaccine viruses for the prevention of H5N1 pandemics. Full article
(This article belongs to the Section Molecular Microbiology)
Show Figures

Figure 1

Figure 1
<p>Binding activity of HI-positive clones. The binding activity of Fab-PP (produced from sixty-two clones) to vaccines of four H5N1, H1N1, and H3N2 strains was examined by ELISA. Clones F081-007 (anti-H1 and -H5 HA (hemagglutinin) antibody) and F005-126 (anti-H3 HA antibody) were used as positive controls. Clone 2-3E (anti-rotavirus antibody) was used as a negative control. The binding activity is shown as the optical density at 450 nm (OD450).</p>
Full article ">Figure 2
<p>Neutralization activity of Fab-PP from HI-positive clones was measured by the focus reduction assay. Neutralization activity was tested at two antibody concentrations (250 µg/mL, blue bars; and 100 µg/mL, red bars). Three kinds of H5N1 viruses were tested, including V: Vie04 (Clade 1), I: Ind05 (Clade 2.1), and A: Anh05 (Clade 2.3); the strain used for each assay is indicated under the clone name. The neutralization activity is shown as the focus reduction rate (%) normalized to that observed in the absence of antibody.</p>
Full article ">Figure 3
<p>Binding activity of HI-positive clones to recombinant HA molecules derived from seven kinds of H5N1 viruses. Binding was measured by ELISA. F081-007 is an anti-H1 and -H5 HA antibody that was used as a positive control. F005-126 (an anti-H3 HA antibody) and F008-009 (an anti-influenza nucleoprotein antibody) were used as negative controls. Anti-FLAG antibody was used for detecting the FLAG tag that had been fused to the recombinant HA. The binding activity is shown as the optical density at 450 nm (OD450).</p>
Full article ">Figure 4
<p>Binding activity of HI-negative clones. Binding activities of thirty-three clones to four H5N1 vaccine strains, H1N1, H3N2, and two kinds of type-B viruses, were examined by ELISA. F081-007 (an anti-H1 and -H5 HA antibody), F005-126 (an anti-H3 HA antibody), and F114-191 (an anti-influenza type-B antibody) were used as positive controls, while 2-3E (anti-rotavirus antibody) was used as a negative control. The binding activity is shown as the optical density at 450 nm (OD450).</p>
Full article ">Figure 5
<p>Neutralization activity of Fab-PP from HI-negative clones was measured by focus reduction assay. Neutralization activity was tested at two antibody concentrations (250 µg/mL, blue bars; and 100 µg/mL, red bars). Three kinds of H5N1 viruses were tested, including V: Vie04 (Clade 1), I: Ind05 (Clade 2.1), and A: Anh05 (Clade 2.3); the strain used for each assay is indicated under the clone name. The neutralization activity is shown as the focus reduction rate (%) normalized to that observed in the absence of antibody.</p>
Full article ">Figure 6
<p>Binding activity of HI-negative clones to recombinant HA molecules derived from seven kinds of H5N1 viruses. Binding was measured by ELISA. F081-007 is an anti-H1 and -H5 HA antibody that was used as a positive control. F005-126 (an anti-H3 HA antibody) and F008-009 (an anti-influenza nucleoprotein antibody) were used as negative controls. Anti-FLAG antibody was used for detecting the FLAG tag that had been fused to the recombinant HA. The binding activity is shown as the optical density at 450 nm (OD450).</p>
Full article ">Figure 7
<p>A pie chart indicating the ratio of anti-H5 antibodies that bound only to H5N1 strains compared to broadly cross-reactive antibodies that bound to H5N1, H1N1, H3N2, and type-B strains. Each of the percentages indicates distribution of representative clones categorized based on their cross-reactivity among H5N1 and H1N1 strains. The bar graph shows the percentage of anti-H5 antibodies categorized into 4 groups based on their cross-reactivity among the four H5N1 vaccine strains used in the present study.</p>
Full article ">Figure 8
<p>Mutation frequency of amino acid residues in V<sub>H</sub>-encoding region when comparing between the representative clone and the corresponding germline gene. (<b>A</b>,<b>B</b>) The number of changed amino acid residues introduced by nucleic acid mutation in the V<sub>H</sub>-encoding region. The representative clones are classified into three categories based on their cross-reactivity profiles, as follows: (<b>A</b>) one vaccine (singular specificity against one H5N1 strain; blue); two, three, and four vaccines (multiple specificity against H5N1 strains; red); (<b>B</b>) Broad cross-reactivity against H1, H3, H5, and type-B viruses; green; (<b>C</b>) The degree of identity of amino acid residues in the protein encoded by the V<sub>H</sub> region of each clone compared to that encoded by the respective germline gene. The identity (%) of each clone shown in <a href="#ijms-21-07422-t003" class="html-table">Table 3</a> is presented in a dot plot format for each category described in (<b>A</b>). Blue: singular specificity against one H5N1 strain; red: multiple specificity against H5N1 strains; and green: broad cross-reactivity against H1, H3, H5, and type-B viruses.</p>
Full article ">Figure 9
<p>Amino acid sequences of antigenic sites on HAs of H5 and H1 strains used in this study. The sequence of the H5N1 Vie04 strain was used as the standard for comparison. The bars indicate identity to the amino acid in the standard. Amino acid sequences in sites Ca1, Ca2, Cb, Sa, and Sb are boxed in blue. Amino acid sequences in the 130-loop, 150-loop, 190-helix, and 220-loop are shaded in yellow-green. Accession numbers of amino acid sequences: EF541402 (Vie04); CY116646 (Ind05); DQ137873 (Qin05); DQ371928 (Anh05); AB621352 (Ind10); KJ522737 (Egy13); AB979487 (Vie14); FJ966974 (Cal09).</p>
Full article ">
19 pages, 13043 KiB  
Article
Intraarticular Administration Effect of Hydrogen Sulfide on an In Vivo Rat Model of Osteoarthritis
by Carlos Vaamonde-García, Elena F. Burguera, Ángela Vela-Anero, Tamara Hermida-Gómez, Purificación Filgueira-Fernández, Jennifer A. Fernández-Rodríguez, Rosa Meijide-Faílde and Francisco J. Blanco
Int. J. Mol. Sci. 2020, 21(19), 7421; https://doi.org/10.3390/ijms21197421 - 8 Oct 2020
Cited by 17 | Viewed by 4876
Abstract
Osteoarthritis (OA) is the most common articular chronic disease. However, its current treatment is limited and mostly symptomatic. Hydrogen sulfide (H2S) is an endogenous gas with recognized physiological activities. The purpose here was to evaluate the effects of the intraarticular administration [...] Read more.
Osteoarthritis (OA) is the most common articular chronic disease. However, its current treatment is limited and mostly symptomatic. Hydrogen sulfide (H2S) is an endogenous gas with recognized physiological activities. The purpose here was to evaluate the effects of the intraarticular administration of a slow-releasing H2S compound (GYY-4137) on an OA experimental model. OA was induced in Wistar rats by the transection of medial collateral ligament and the removal of the medial meniscus of the left joint. The animals were randomized into three groups: non-treated and intraarticularly injected with saline or GYY-4137. Joint destabilization induced articular thickening (≈5% increment), the loss of joint mobility and flexion (≈12-degree angle), and increased levels of pain (≈1.5 points on a scale of 0 to 3). Animals treated with GYY-4137 presented improved motor function of the joint, as well as lower pain levels (≈75% recovery). We also observed that cartilage deterioration was attenuated in the GYY-4137 group (≈30% compared with the saline group). Likewise, these animals showed a reduced presence of pro-inflammatory mediators (cyclooxygenase-2, inducible nitric oxide synthase, and metalloproteinase-13) and lower oxidative damage in the cartilage. The increment of the nuclear factor-erythroid 2-related factor 2 (Nrf-2) levels and Nrf-2-regulated gene expression (≈30%) in the GYY-4137 group seem to be underlying its chondroprotective effects. Our results suggest the beneficial impact of the intraarticular administration of H2S on experimental OA, showing a reduced cartilage destruction and oxidative damage, and supporting the use of slow H2S-producing molecules as a complementary treatment in OA. Full article
(This article belongs to the Special Issue Redox Signaling and Oxidative Stress in Bone Health and Disease)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Macroscopic and clinical evaluation. Macroscopic evaluation of the animals from the experimental groups (under non-treatment (C); intraarticular injection of saline (CI) or H<sub>2</sub>S donor (SI)) was performed. Firstly, the weight (<b>a</b>) and articular diameter (<b>b</b>) were monitored. Evaluation of the pain levels (<b>c</b>) was performed with an arbitrary scale, as previously indicated. Finally, the angle of joint flexion (<b>d</b>) and extension (<b>e</b>) over the course of the model were assessed by a protractor. Values are mean ± SEM (<span class="html-italic">n</span> = 6 independent animals for each condition). * <span class="html-italic">p</span> ≤ 0.05 vs. day 0; <sup><span>$</span></sup> <span class="html-italic">p</span> ≤ 0.05 vs. day 7; <sup>&amp;</sup> <span class="html-italic">p</span> ≤ 0.05 vs. the respective day in the control group; <sup>#</sup> <span class="html-italic">p</span> ≤ 0.05 vs. the respective day in the CI group.</p>
Full article ">Figure 2
<p>Analysis of the articular motor function with the Rotarod performance test. Animals from the three experimental groups were forced to move in the rotating cylinder for 300 s. The number of falls (<b>a</b>) during this period was monitored, as well as the time remaining on the rotarod (<b>b</b>). Values are mean ± SEM (<span class="html-italic">n</span> = 6 independent animals for each condition). * <span class="html-italic">p</span> ≤ 0.05 vs. day 0; <sup><span>$</span></sup> <span class="html-italic">p</span> ≤ 0.05 vs. day 7; <sup>&amp;</sup> <span class="html-italic">p</span> ≤ 0.05 vs. the respective day in the C group; <sup>#</sup> <span class="html-italic">p</span> ≤ 0.05 vs. the respective day in the CI group. C, non-treated; CI, control injection; SI, H<sub>2</sub>S donor injection.</p>
Full article ">Figure 3
<p>Histological analysis of articular cartilage. Cartilage lesions were evaluated by the semi-quantitative modified osteoarthritis Research Society International (OARSI) score, as previously indicated in the Materials and Methods section. (<b>a</b>) Representative images of the joint sections stained with Safranin-O-fast green from each group of the study, showing the cartilage of the medial compartment from the tibial plateau (MTP) and femoral condyle (MFC) in the right knee (sham surgery) and left knee (OA surgery). Arrows indicate areas with a loss of cartilage matrix and ★ indicates cartilage with the loss of Safranin staining (indicator of proteoglycan content). Analysis of the semi-quantitative score of the pathological alterations in the cartilage from MFC (<b>b</b>) and MTP (<b>c</b>). Values are mean ± SEM (<span class="html-italic">n</span> = 3 independent animals for each condition). * <span class="html-italic">p</span> ≤ 0.05 vs. the respective sham-operated joint; <sup>&amp;</sup> <span class="html-italic">p</span> ≤ 0.05 vs. the C group; <sup>#</sup> <span class="html-italic">p</span> ≤ 0.05 vs. the CI group. Control, non-treated; CI, control injection; SI, H<sub>2</sub>S donor injection. Scale bar = 500 µm.</p>
Full article ">Figure 4
<p>Histological analysis of synovial tissue. Pathological responses in the synovial tissue were evaluated by a semi-quantitative Krenn score, as previously indicated in the Materials and Methods section. (<b>a</b>) Representative images of synovial sections stained with hematoxylin and eosin from each group of the study, showing synovium in the medial compartment of the right knee (sham surgery) and left knee (OA surgery). Analysis of the semi-quantitative score of the following pathological alterations in the synovial tissue: number of lining cell layers (<b>b</b>), proliferation of the subintima tissue (<b>c</b>), and infiltration of inflammatory cells (<b>d</b>). (<b>e</b>) The sum up of the pathological changes in the tissue is shown. Values are mean ± SEM (<span class="html-italic">n</span> = 3 independent animals for each condition). * <span class="html-italic">p</span> ≤ 0.05 vs. the respective sham-operated joint. Control, non-treated; CI, control injection; SI, H<sub>2</sub>S donor injection. Scale bar = 50 µm.</p>
Full article ">Figure 5
<p>Presence of oxidative damage markers and antioxidant response in the articular cartilage. (<b>a</b>) Representative images of 8-hydroxy-2′-deoxyguanosine (8-oxo-dG), 4-hydroxy-2-nonenal (4-HNE), and nuclear factor erythroid-derived 2-like 2 (Nrf-2) immunohistochemistry in the cartilage from each group of the study. Magnification (2x) of the images are shown on its bottom-left corner. Quantitative analysis of 8-oxo-dG (<b>b</b>), 4-HNE (<b>c</b>), and Nrf-2 (<b>d</b>)-positive cells. Values are mean ± SEM (<span class="html-italic">n</span> = 3 independent animals for each condition). <sup>&amp;</sup> <span class="html-italic">p</span> ≤ 0.05 vs. the C group; <sup>#</sup> <span class="html-italic">p</span> ≤ 0.05 vs. the CI group. Control, non-treated; CI, control injection; SI, H<sub>2</sub>S donor injection. Scale bar = 20 µm.</p>
Full article ">Figure 6
<p>Presence of pro-catabolic mediators in the articular cartilage. (<b>a</b>) Representative images of metalloproteinase-13 (MMP-13), inducible nitric oxide synthase (iNOS), and cyclooxygenase 2 (COX-2) immunohistochemistry in the cartilage from each group of study. Magnifications (2x) of the images are shown on their bottom-left corners. Quantitative analysis of MMP-13 (<b>b</b>), iNOS (<b>c</b>), and COX-2 (<b>d</b>)-positive cells. Values are mean ± SEM (<span class="html-italic">n</span> = 3 independent animals for each condition). <sup>&amp;</sup> <span class="html-italic">p</span> ≤ 0.05 vs. the C group; <sup>#</sup> <span class="html-italic">p</span> ≤ 0.05 vs. the CI group. Control, non-treated; CI, control injection; SI, H<sub>2</sub>S donor injection. Scale bar = 50 µm.</p>
Full article ">Figure 7
<p>Measurement of the gene expression of antioxidant and pro-inflammatory markers. Relative expression levels of genes involved in antioxidant response (<b>a</b>) (nuclear factor erythroid-derived 2-like 2 (<span class="html-italic">Nfe2l2</span>), nicotinamide adenine dinucleotide phosphate hydrogen (NADPH):quinone oxidoreductase (<span class="html-italic">Nqo1</span>), and heme oxygenase-1 (<span class="html-italic">Hmox1</span>)) and in pro-inflammatory signaling (<b>b</b>) (interleukin-1β (<span class="html-italic">Il1b</span>), tumor necrosis factor-α (<span class="html-italic">Tnf</span>), cytokine-induced neutrophil chemoattractant-1 (<span class="html-italic">Cxcl1</span>), cyclooxygenase 2 (<span class="html-italic">Ptgse</span>)) were analyzed in blood samples from each group of the study, as previously indicated. Values are mean ± SEM (<span class="html-italic">n</span> = 3 independent animals for each condition). * <span class="html-italic">p</span> ≤ 0.05 vs. the CI group. CI, control injection; SI, H<sub>2</sub>S donor injection.</p>
Full article ">
20 pages, 5481 KiB  
Article
Prognostic Role of Survivin and Macrophage Infiltration Quantified on Protein and mRNA Level in Molecular Subtypes Determined by RT-qPCR of KRT5, KRT20, and ERBB2 in Muscle-Invasive Bladder Cancer Treated by Adjuvant Chemotherapy
by Thorsten H. Ecke, Adisch Kiani, Thorsten Schlomm, Frank Friedersdorff, Anja Rabien, Klaus Jung, Ergin Kilic, Peter Boström, Minna Tervahartiala, Pekka Taimen, Jan Gleichenhagen, Georg Johnen, Thomas Brüning, Stefan Koch, Jenny Roggisch and Ralph M. Wirtz
Int. J. Mol. Sci. 2020, 21(19), 7420; https://doi.org/10.3390/ijms21197420 - 8 Oct 2020
Cited by 2 | Viewed by 3339
Abstract
Objectives: Bladder cancer is a heterogeneous malignancy. Therefore, it is difficult to find single predictive markers. Moreover, most studies focus on either the immunohistochemical or molecular assessment of tumor tissues by next-generation sequencing (NGS) or PCR, while a combination of immunohistochemistry (IHC) and [...] Read more.
Objectives: Bladder cancer is a heterogeneous malignancy. Therefore, it is difficult to find single predictive markers. Moreover, most studies focus on either the immunohistochemical or molecular assessment of tumor tissues by next-generation sequencing (NGS) or PCR, while a combination of immunohistochemistry (IHC) and PCR for tumor marker assessment might have the strongest impact to predict outcome and select optimal therapies in real-world application. We investigated the role of proliferation survivin/BIRC5 and macrophage infiltration (CD68, MAC387, CLEVER-1) on the basis of molecular subtypes of bladder cancer (KRT5, KRT20, ERBB2) to predict outcomes of adjuvant treated muscle-invasive bladder cancer patients with regard to progression-free survival (PFS) and disease-specific survival (DSS). Materials and Methods: We used tissue microarrays (TMA) from n = 50 patients (38 males, 12 female) with muscle-invasive bladder cancer. All patients had been treated with radical cystectomy followed by adjuvant triple chemotherapy. Median follow-up time was 60.5 months. CD68, CLEVER-1, MAC387, and survivin protein were detected by immunostaining and subsequent visual inspection. BIRC5, KRT5, KRT20, ERBB2, and CD68 mRNAs were detected by standardized RT-qPCR after tissue dot RNA extraction using a novel stamp technology. All these markers were evaluated in three different centers of excellence. Results: Nuclear staining rather than cytoplasmic staining of survivin predicted DSS as a single marker with high levels of survivin being associated with better PFS and DSS upon adjuvant chemotherapy (p = 0.0138 and p = 0.001, respectively). These results were validated by the quantitation of BIRC5 mRNA by PCR (p = 0.0004 and p = 0.0508, respectively). Interestingly, nuclear staining of survivin protein was positively associated with BIRC5 mRNA, while cytoplasmic staining was inversely related, indicating that the translocation of survivin protein into the nucleus occurred at a discrete, higher level of its mRNA. Combining survivin/BIRC5 levels based on molecular subtype being assessed by KRT20 expression improved the predictive value, with tumors having low survivin/BIRC5 and KRT20 mRNA levels having the best survival (75% vs. 20% vs. 10% 5-year DSS, p = 0.0005), and these values were independent of grading, node status, and tumor stage in multivariate analysis (p = 0.0167). Macrophage infiltration dominated in basal tumors and was inversely related with the luminal subtype marker gene expression. The presence of macrophages in survivin-positive or ERBB2-positive tumors was associated with worse DSS. Conclusions: For muscle-invasive bladder cancer patients, the proliferative activity as determined by the nuclear staining of survivin or RT-qPCR on the basis of molecular subtype characteristics outperforms single marker detections and single technology approaches. Infiltration by macrophages detected by IHC or PCR is associated with worse outcome in defined subsets of tumors. The limitations of this study are the retrospective nature and the limited number of patients. However, the number of molecular markers has been restricted and based on predefined assumptions, which resulted in the dissection of muscle-invasive disease into tumor–biological axes of high prognostic relevance, which warrant further investigation and validation. Full article
Show Figures

Figure 1

Figure 1
<p>Remark diagram.</p>
Full article ">Figure 2
<p>(<b>a</b>) Data distribution of immunohistochemical staining of CD68, MAC387, and common lymphatic endothelial and vascular endothelial receptor-1 (CLEVER-1) by visual analysis and survivin by semi-quantitative assessment of cytoplasmic versus nuclear stain; (<b>b</b>) Data distribution and box and whisker plot of <span class="html-italic">KRT5</span>, <span class="html-italic">KRT20</span>, <span class="html-italic">ERBB2</span>, <span class="html-italic">BIRC5</span>, and <span class="html-italic">CD68</span> mRNA levels in the bladder cancer study cohort treated by adjuvant chemotherapy (<span class="html-italic">n</span> = 39). Normalized gene expression (40-DCT method) as well as quantile values are depicted in the y-axis. DCT: Delta Cycle Threshold.</p>
Full article ">Figure 3
<p>Spearman correlation of IHC staining and semi-quantitative assessment of survivin protein located in cytoplasmic versus nuclear localization with quantitative <span class="html-italic">BIRC5</span> (survivin) mRNA levels in the combined PCR and IHC cohort (<span class="html-italic">n</span> = 28). Graphical display of Spearman rho values and respective <span class="html-italic">p</span>-values are depicted.</p>
Full article ">Figure 4
<p>(<b>a</b>) Correlation of normalized <span class="html-italic">KRT5</span>, <span class="html-italic">KRT20</span>, <span class="html-italic">ERBB2</span>, <span class="html-italic">BIRC5</span>, and <span class="html-italic">CD68</span> mRNA levels in the PCR cohort (<span class="html-italic">n</span> = 39) of bladder cancer patients treated with adjuvant chemotherapy. Graphical display of Spearman rho values and respective p-values are depicted. * indicates statistically significant results; (<b>b</b>) Correlation of <span class="html-italic">KRT5</span>, <span class="html-italic">KRT20</span>, <span class="html-italic">ERBB2</span>, and <span class="html-italic">BIRC5</span> mRNA levels with protein levels of CD68, MAC387, and CLEVER-1 determined by IHC in the combined PCR and IHC cohort (<span class="html-italic">n</span> = 28) of bladder cancer patients treated with adjuvant chemotherapy. Graphical display of Spearman rho values and respective p-values are depicted. * indicates statistically significant results.</p>
Full article ">Figure 4 Cont.
<p>(<b>a</b>) Correlation of normalized <span class="html-italic">KRT5</span>, <span class="html-italic">KRT20</span>, <span class="html-italic">ERBB2</span>, <span class="html-italic">BIRC5</span>, and <span class="html-italic">CD68</span> mRNA levels in the PCR cohort (<span class="html-italic">n</span> = 39) of bladder cancer patients treated with adjuvant chemotherapy. Graphical display of Spearman rho values and respective p-values are depicted. * indicates statistically significant results; (<b>b</b>) Correlation of <span class="html-italic">KRT5</span>, <span class="html-italic">KRT20</span>, <span class="html-italic">ERBB2</span>, and <span class="html-italic">BIRC5</span> mRNA levels with protein levels of CD68, MAC387, and CLEVER-1 determined by IHC in the combined PCR and IHC cohort (<span class="html-italic">n</span> = 28) of bladder cancer patients treated with adjuvant chemotherapy. Graphical display of Spearman rho values and respective p-values are depicted. * indicates statistically significant results.</p>
Full article ">Figure 5
<p>(<b>a</b>) Disease-specific survival (DSS) of bladder cancer patients treated with adjuvant chemotherapy based on survivin nuclear stain in the PCR and IHC cohort. (<b>b</b>) DSS of bladder cancer patients treated with adjuvant chemotherapy based on <span class="html-italic">BIRC5</span> mRNA expression in the PCR cohort.</p>
Full article ">Figure 6
<p>(<b>a</b>) DSS of bladder cancer patients treated with adjuvant chemotherapy based on <span class="html-italic">KRT20</span> mRNA and survivin nuclear protein stain in the PCR and IHC cohort; (<b>b</b>) DSS of bladder cancer patients treated with adjuvant chemotherapy based on <span class="html-italic">KRT20</span> and <span class="html-italic">BIRC5</span> mRNA expression in the PCR cohort.</p>
Full article ">Figure 7
<p>(<b>a</b>) DSS of bladder cancer patients treated with adjuvant chemotherapy based on survivin nuclear protein staining and <span class="html-italic">CD68</span> mRNA in the PCR and IHC cohort. (<b>b</b>) DSS of bladder cancer patients treated with adjuvant chemotherapy based on survivin nuclear protein staining and MAC387 protein in the PCR and IHC cohort.</p>
Full article ">Figure 8
<p>DSS of bladder cancer patients treated with adjuvant chemotherapy based on <span class="html-italic">ERBB2</span>-positive tumors in relation to <span class="html-italic">CD68</span> mRNA levels in the PCR and IHC cohort.</p>
Full article ">
19 pages, 620 KiB  
Review
Recent Advances in Drosophila Models of Charcot-Marie-Tooth Disease
by Fukiko Kitani-Morii and Yu-ichi Noto
Int. J. Mol. Sci. 2020, 21(19), 7419; https://doi.org/10.3390/ijms21197419 - 8 Oct 2020
Cited by 8 | Viewed by 8866
Abstract
Charcot-Marie-Tooth disease (CMT) is one of the most common inherited peripheral neuropathies. CMT patients typically show slowly progressive muscle weakness and sensory loss in a distal dominant pattern in childhood. The diagnosis of CMT is based on clinical symptoms, electrophysiological examinations, and genetic [...] Read more.
Charcot-Marie-Tooth disease (CMT) is one of the most common inherited peripheral neuropathies. CMT patients typically show slowly progressive muscle weakness and sensory loss in a distal dominant pattern in childhood. The diagnosis of CMT is based on clinical symptoms, electrophysiological examinations, and genetic testing. Advances in genetic testing technology have revealed the genetic heterogeneity of CMT; more than 100 genes containing the disease causative mutations have been identified. Because a single genetic alteration in CMT leads to progressive neurodegeneration, studies of CMT patients and their respective models revealed the genotype-phenotype relationships of targeted genes. Conventionally, rodents and cell lines have often been used to study the pathogenesis of CMT. Recently, Drosophila has also attracted attention as a CMT model. In this review, we outline the clinical characteristics of CMT, describe the advantages and disadvantages of using Drosophila in CMT studies, and introduce recent advances in CMT research that successfully applied the use of Drosophila, in areas such as molecules associated with mitochondria, endosomes/lysosomes, transfer RNA, axonal transport, and glucose metabolism. Full article
(This article belongs to the Special Issue Role of Drosophila in Human Disease Research 2.0)
Show Figures

Figure 1

Figure 1
<p>Schematic summary showing various Charcot-Marie-Tooth disease (CMT)-related genes and pathways in the peripheral nerve. If each gene has multiple functions, the most representative one is described. The enlarged box shows a cross-sectional view of the peripheral nerve.</p>
Full article ">
27 pages, 9780 KiB  
Article
A Comparative Transcriptome Analysis, Conserved Regulatory Elements and Associated Transcription Factors Related to Accumulation of Fusariotoxins in Grain of Rye (Secale cereale L.) Hybrids
by Khalid Mahmood, Jihad Orabi, Peter Skov Kristensen, Pernille Sarup, Lise Nistrup Jørgensen and Ahmed Jahoor
Int. J. Mol. Sci. 2020, 21(19), 7418; https://doi.org/10.3390/ijms21197418 - 8 Oct 2020
Cited by 4 | Viewed by 3236
Abstract
Detoxification of fusariotoxin is a type V Fusarium head blight (FHB) resistance and is considered a component of type II resistance, which is related to the spread of infection within spikes. Understanding this type of resistance is vital for FHB resistance, but to [...] Read more.
Detoxification of fusariotoxin is a type V Fusarium head blight (FHB) resistance and is considered a component of type II resistance, which is related to the spread of infection within spikes. Understanding this type of resistance is vital for FHB resistance, but to date, nothing is known about candidate genes that confer this resistance in rye due to scarce genomic resources. In this study, we generated a transcriptomic resource. The molecular response was mined through a comprehensive transcriptomic analysis of two rye hybrids differing in the build-up of fusariotoxin contents in grain upon pathogen infection. Gene mining identified candidate genes and pathways contributing to the detoxification of fusariotoxins in rye. Moreover, we found cis regulatory elements in the promoters of identified genes and linked them to transcription factors. In the fusariotoxin analysis, we found that grain from the Nordic seed rye hybrid “Helltop” accumulated 4 times higher concentrations of deoxynivalenol (DON), 9 times higher nivalenol (NIV), and 28 times higher of zearalenone (ZEN) than that of the hybrid “DH372” after artificial inoculation under field conditions. In the transcriptome analysis, we identified 6675 and 5151 differentially expressed genes (DEGs) in DH372 and Helltop, respectively, compared to non-inoculated control plants. A Kyoto Encyclopedia of Genes and Genomes (KEGG) analysis revealed that DEGs were associated with glycolysis and the mechanistic target of rapamycin (mTOR) signaling pathway in Helltop, whereas carbon fixation in photosynthesis organisms were represented in DH372. The gene ontology (GO) enrichment and gene set enrichment analysis (GSEA) of DEGs lead to identification of the metabolic and biosynthetic processes of peptides and amides in DH372, whereas photosynthesis, negative regulation of catalytic activity, and protein-chromophore linkage were the significant pathways in Helltop. In the process of gene mining, we found four genes that were known to be involved in FHB resistance in wheat and that were differentially expressed after infection only in DH372 but not in Helltop. Based on our results, we assume that DH372 employed a specific response to pathogen infection that led to detoxification of fusariotoxin and prevented their accumulation in grain. Our results indicate that DH372 might resist the accumulation of fusariotoxin through activation of the glycolysis and drug metabolism via cytochrome P450. The identified genes in DH372 might be regulated by the WRKY family transcription factors as associated cis regulatory elements found in the in silico analysis. The results of this study will help rye breeders to develop strategies against type V FHB. Full article
(This article belongs to the Section Molecular Plant Sciences)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Graphical representation of the experimental setup and data analysis: this setup covers the fusariotoxin analysis, RNAseq experimental setup, RNA extraction, and subsequently next generation sequencing (NGS) data analysis; 5 DAI represents 5 days after inoculation.</p>
Full article ">Figure 2
<p>Distribution and annotations of differentially expressed genes (DEGs) of both hybrids after inoculation with <span class="html-italic">Fusarium</span> spp. through assigning enzyme classes and KEGG (Kyoto Encyclopedia of Genes and Genome) pathways: (<b>A</b>) the Venn diagram illustrates the number of differentially expressed genes in both hybrids and divides them into three categories. (<b>B</b>) The major classes of enzymes of DEGs associated to exclusive DH372, exclusive Helltop, and common category and (<b>C</b>) the top 12 most highly represented KEGG pathways belonging to each category are shown. Analysis was performed using the OmicsBox and the KEGG database.</p>
Full article ">Figure 3
<p>Gene enrichment analysis of differentially expressed genes (DEGs): this enrichment analysis was carried out using the upregulated genes after inoculation and belong to three different categories, i.e., (<b>A</b>) upregulated genes exclusive to DH372, (<b>B</b>) upregulated genes exclusive to Helltop, and (<b>C</b>) upregulated genes common in both hybrids. The top 15 significant gene-enriched gene ontology (GO) terms are shown. The test set represents the DEGs that were upregulated after inoculation, and the reference set is the total number of genes in rye known to be involved in that specific function.</p>
Full article ">Figure 4
<p>Gene enrichment analysis of differentially expressed genes (DEGs): this enrichment analysis was carried out using the downregulated genes after inoculation in both hybrids and belong to three different categories, i.e., (<b>A</b>) downregulated genes exclusive to DH372, (<b>B</b>) downregulated genes exclusive to Helltop, and (<b>C</b>) downregulated genes common in both hybrids. The top 15 significant gene-enriched gene ontology (GO) terms are shown. The test set represents the DEGs that were upregulated after inoculation, and the reference set is the total number of genes in rye known to be involved in that specific function.</p>
Full article ">Figure 5
<p>Gene set enrichment analysis (GSEA) of differentially expressed genes in both hybrids: (<b>A</b>) GSEA of DH372 and (<b>B</b>) GSEA of Helltop. The GSEA revealed that genes were recruited into different pathways in both hybrids after <span class="html-italic">Fusarium</span> spp. inoculation. The GSEA is presented in the form of a word cloud.</p>
Full article ">Figure 6
<p>Dissection and mining of differentially expressed genes (DEGs): (<b>A</b>) the overall comparison of all DEGs was divided into three different categories known as exclusively responding genes, common responding genes, and contrasting responding genes categories. The exclusively responsive genes are the ones that were differentially expressed in one hybrid but absent in other. The common responding genes are the ones that exhibited similar differential expression pattern in both hybrids, whereas contrasting responding genes are the one that exhibited opposite response. (<b>B</b>) Comparison of cytochrome P450s in both hybrids and their division into three categories as mentioned above and (<b>C</b>) comparison of glycosyltransferases (GTs) in both hybrids and their division into the categories: the pentagons, stars, and circle encompass the total number of genes present in different comparisons. Venn diagrams were plotted using online tools Venny 2.0 (<a href="http://bioinfogp.cnb.csic.es/tools/venny/" target="_blank">http://bioinfogp.cnb.csic.es/tools/venny/</a>).</p>
Full article ">Figure 7
<p>Gene expression patterns of wheat orthologs during different developmental stages and under various stresses: (<b>A</b>,<b>B</b>) these seven wheat orthologs belong to common responding genes, and (<b>C</b>,<b>D</b>) these four genes belong to exclusively responding genes in DH372. Developmental stage-specific expression patterns are shown in A and C. Our selected genes tend to show higher expression during anthesis. “HIGH,” “MEDIUM,” and “LOW” expressions were based on gene expression data found in GENEVESTIGATOR (<a href="http://www.genevestigator.com" target="_blank">http://www.genevestigator.com</a>). Heat map of expression of selected genes in response to various stresses (<b>B</b>,<b>D</b>) using Genevestigator perturbation tool: the relative expression of the genes was represented in a log2 ratio, and significant changes in expression were filtered out based on <span class="html-italic">p</span> &lt; 0.001. Expression of genes was strongly induced in response to biotic stresses including <span class="html-italic">Fusarium</span> spp.</p>
Full article ">
24 pages, 5532 KiB  
Article
Structural and Computational Insights into a Blebbistatin-Bound Myosin•ADP Complex with Characteristics of an ADP-Release Conformation along the Two-Step Myosin Power Stoke
by Wiebke Ewert, Peter Franz, Georgios Tsiavaliaris and Matthias Preller
Int. J. Mol. Sci. 2020, 21(19), 7417; https://doi.org/10.3390/ijms21197417 - 8 Oct 2020
Cited by 3 | Viewed by 4149
Abstract
The motor protein myosin drives a wide range of cellular and muscular functions by generating directed movement and force, fueled through adenosine triphosphate (ATP) hydrolysis. Release of the hydrolysis product adenosine diphosphate (ADP) is a fundamental and regulatory process during force production. However, [...] Read more.
The motor protein myosin drives a wide range of cellular and muscular functions by generating directed movement and force, fueled through adenosine triphosphate (ATP) hydrolysis. Release of the hydrolysis product adenosine diphosphate (ADP) is a fundamental and regulatory process during force production. However, details about the molecular mechanism accompanying ADP release are scarce due to the lack of representative structures. Here we solved a novel blebbistatin-bound myosin conformation with critical structural elements in positions between the myosin pre-power stroke and rigor states. ADP in this structure is repositioned towards the surface by the phosphate-sensing P-loop, and stabilized in a partially unbound conformation via a salt-bridge between Arg131 and Glu187. A 5 Å rotation separates the mechanical converter in this conformation from the rigor position. The crystallized myosin structure thus resembles a conformation towards the end of the two-step power stroke, associated with ADP release. Computationally reconstructing ADP release from myosin by means of molecular dynamics simulations further supported the existence of an equivalent conformation along the power stroke that shows the same major characteristics in the myosin motor domain as the resolved blebbistatin-bound myosin-II·ADP crystal structure, and identified a communication hub centered on Arg232 that mediates chemomechanical energy transduction. Full article
(This article belongs to the Section Molecular Biophysics)
Show Figures

Figure 1

Figure 1
<p>The actomyosin system. (<b>A</b>) Scheme of the actomyosin motor cycle. The myosin motor is subject to conformational changes of the subdomains upon interaction with both the actin filament and the nucleotide adenosine triphosphate (ATP), as well as during nucleotide hydrolysis, and release of its hydrolysis products inorganic phosphate (P<sub>i</sub>) and adenosine diphosphate (ADP) to eventually produce force in a two-step mechanism. Color code: subdomains N-terminal domain (pink), U50 kDa (blue), L50 kDa (green), converter (orange), lever arm (grey). (<b>B</b>) Structure of the nucleotide-free <span class="html-italic">Dd</span> myosin-II motor domain, possessing a closed actin-binding cleft, a twisted transducer (cyan), and the converter (orange)/lever arm (grey) in the down position. The active site with the nucleotide sensors P-loop (yellow), switch-1 (red), and switch-2 (brown) is empty and prepared for ATP binding.</p>
Full article ">Figure 2
<p>X-ray crystal structure of the <span class="html-italic">Dd</span> myosin-II motor domain in complex with ADP and blebbistatin. (<b>A</b>) Overview of the crystal structure in the blebbistatin-bound ADP-release conformation. The P-loop together with the bound ADP (grey) is shifted towards the surface. Blebbistatin (magenta) is buried in the known allosteric binding pocket at the apex of the large actin-binding cleft. (<b>B</b>) The 2F<sub>o</sub>-F<sub>c</sub> density map of ADP in its new binding position. The map was contoured at 1.0 σ. Note that the β-phosphate has rotated away from the P-loop. (<b>C</b>) 2F<sub>o</sub>-F<sub>c</sub> density map of blebbistatin at the apex of the actin-binding cleft and near the relay helix. The map was contoured at 1.0 σ.</p>
Full article ">Figure 3
<p>The blebbistatin-bound myosin-II·ADP structure (blue) feature characteristics of a myosin conformation between the pre-power stroke (black), and the rigor-like states (grey). (<b>A</b>) The transducer is partially twisted, particularly β-strands 1 to 3. (<b>B</b>) The relay helix and the converter approached the rigor-like position. The converter/lever arm must undergo a ~5° rotation to transition to the final rigor down position (grey). The N-terminal domain is markedly shifted towards its rigor position. (<b>C</b>) The large actin-binding cleft is found in a position between pre-power stroke (black) and rigor-like states (grey), thereby creating a new actin-binding interface. (<b>D</b>) The active site P-loop has moved ~9 Å from its pre-power stroke position in the blebbistatin-bound myosin-II·ADP conformation, but has not reached the rigor position.</p>
Full article ">Figure 4
<p>The myosin-II·ADP·blebbistatin crystal structure shows unique conformational features in the active site. (<b>A</b>) ADP interacts with its phosphate groups, primarily with the reoriented P-loop. The β-phosphate is partially unbound and oriented towards the surrounding water at the protein surface. The adenosine base of the nucleotide is nestled into a groove, formed by a salt-bridge between Arg131 and Glu187. (<b>B</b>) Close-up view on the active site in the myosin-II·ADP·blebbistatin structure shows the critical salt-bridge between switch-1 (Arg238) and switch-2 (Glu459), as well as the newly identified salt-bridge between Arg131 and Glu187 that stabilizes ADP in the release position. A third, complex salt-bridge between Arg232 (switch-1), Asp674 (SH2-helix), and Glu180 (P-loop) seems to be important for mediating the chemomechanical coupling between the active site, the actin-binding region and the mechanical converter/lever arm. (<b>C</b>) Close-up view on the active site in the rigor-like, nucleotide-free myosin-II structure. Due to slight adjustments to the active site, the interaction network around Arg232 is altered as compared to the myosin-II·ADP·blebbistatin structure, and Arg232 interacts with Asp674 (SH2-helix), and Ile460 (switch-2), while Glu180 (P-loop) formed a hydrogen-bond to Gly457 (switch-2) in this conformation.</p>
Full article ">Figure 5
<p>Binding assays of ADP to myosin-II using microscale thermophoresis. (<b>A</b>) The affinity of ADP for myosin-II in the absence of magnesium is reduced from K<sub>D</sub> = 37.5 ± 7.3 µM (10 mM Mg<sup>2+</sup>) to 145.8 ± 19.3 µM (0 mM Mg<sup>2+</sup>). (<b>B</b>) Addition of 100 µM blebbistatin increased the binding affinity of ADP for myosin-II, both in the presence (K<sub>D</sub> = 21.2 ± 3.4 µM) and absence of magnesium (K<sub>D</sub> = 40.6 ± 3.2 µM).</p>
Full article ">Figure 6
<p>Classical molecular dynamics (cMD) simulations support the stabilizing effect of blebbistatin on the crystallographically determined conformation of myosin-II. (<b>A</b>) Root mean square deviations of the protein backbone atoms, averaged over duplicate 500 ns cMD simulations of blebbistatin-bound (magenta) and -unbound (blue) myosin, confirm the stabilizing effect of blebbistatin on the crystallized conformation, while in the absence of blebbistatin, larger deviations from the start structure were observed. (<b>B</b>) Blebbistatin remains embedded in its binding pocket throughout the simulations, showing only negligible deviations from its crystallized binding pose. (<b>C</b>) Root mean square deviations of the converter indicate no movement of the converter while blebbistatin is bound (magenta); however, a smaller shift towards the pre-power stroke position of the converter is monitored in the absence of blebbistatin (blue). Shown in diagrams A-C are the averages of duplicate cMD simulations. (<b>D</b>) The positions of active site phosphate sensors are stabilized spatiotemporally along the cMD simulations in the presence of blebbistatin, with the critical salt-bridge formed. Binding of ADP in the crystallized conformation appeared loose during the cMD simulations with considerable fluctuations of the nucleotide. (<b>E</b>) During cMD simulations in the absence of blebbistatin, the P-loop moved towards switch-1 and -2 in the reversed direction of the power stroke. The nucleotide ADP was markedly shifted together with the P-loop inside into the active site.</p>
Full article ">Figure 7
<p>Pulling MD simulations using biasing forces on the converter domain visualize coupled conformational transitions in myosin and ADP dissociation. (<b>A</b>) Targeted molecular dynamics (TMD) simulations of the converter rotation from the up position of a strong-ADP-bound myosin conformation with closed actin-binding cleft and partially twisted transducer (light blue) towards the down position (dark blue) were coupled to rearrangements of the N-terminal domain. Structural changes are indicated by vectors. (<b>B</b>) The induced converter rotation during the TMD simulations led to a significant ~6 Å shift of the P-loop towards the front entrance of the active site. The bound ADP moved together with the P-loop towards the surface and lost interactions, which finally led to larger fluctuation of the nucleotide, and the population of a conformation similar to the myosin-II·ADP·blebbistatin crystal structure. (<b>C</b>) Visualization of the release vectors used to pull the nucleotide out of the crystallized position in the myosin motor domain. (<b>D</b>) Pulling forces required to dissociate ADP from myosin along the four different vectors. Independent of the release vector, a consistent set of interactions was involved in ADP dissociation. Shown are the averages of duplicate steered MD simulations.</p>
Full article ">Figure 8
<p>Unbiased MD simulations captured a potential ADP release pathway. (<b>A</b>) Molecular events during ADP dissociation along the unbiased MD trajectories reveal a guiding mechanism of Arg131 together with Thr186. Panels I through III show snapshots along the cMD trajectory. (<b>B</b>) Distance of the protein residues Arg131 and Thr186 to the nucleotide β-phosphate over the MD simulations illustrate the interplay between the two protein residues in the proposed ADP release mechanism. (<b>C</b>) Energy profile of the reconstructed ADP release pathway from a strongly bound ADP to nucleotide dissociation as determined by the molecular mechanics/generalized-born surface area (MM/GBSA) interaction energy between the protein and ADP, obtained by combining TMD and cMD trajectories. Both the myosin conformation with strongly bound ADP (orange dot) and the myosin conformation resembling our myosin-II·ADP·blebbistatin structure (blue dot) represent energy minima (low-energy conformations) along the reconstructed myosin power stroke.</p>
Full article ">
15 pages, 4074 KiB  
Article
PPAR-α Deletion Attenuates Cisplatin Nephrotoxicity by Modulating Renal Organic Transporters MATE-1 and OCT-2
by Leandro Ceotto Freitas-Lima, Alexandre Budu, Adriano Cleis Arruda, Mauro Sérgio Perilhão, Jonatan Barrera-Chimal, Ronaldo Carvalho Araujo and Gabriel Rufino Estrela
Int. J. Mol. Sci. 2020, 21(19), 7416; https://doi.org/10.3390/ijms21197416 - 8 Oct 2020
Cited by 29 | Viewed by 3383
Abstract
Cisplatin is a chemotherapy drug widely used in the treatment of solid tumors. However, nephrotoxicity has been reported in about one-third of patients undergoing cisplatin therapy. Proximal tubules are the main target of cisplatin toxicity and cellular uptake; elimination of this drug can [...] Read more.
Cisplatin is a chemotherapy drug widely used in the treatment of solid tumors. However, nephrotoxicity has been reported in about one-third of patients undergoing cisplatin therapy. Proximal tubules are the main target of cisplatin toxicity and cellular uptake; elimination of this drug can modulate renal damage. Organic transporters play an important role in the transport of cisplatin into the kidney and organic cations transporter 2 (OCT-2) has been shown to be one of the most important transporters to play this role. On the other hand, multidrug and toxin extrusion 1 (MATE-1) transporter is the main protein that mediates the extrusion of cisplatin into the urine. Cisplatin nephrotoxicity has been shown to be enhanced by increased OCT-2 and/or reduced MATE-1 activity. Peroxisome proliferator-activated receptor alpha (PPAR-α) is the transcription factor which controls lipid metabolism and glucose homeostasis; it is highly expressed in the kidneys and interacts with both MATE-1 and OCT-2. Considering the above, we treated wild-type and PPAR-α knockout mice with cisplatin in order to evaluate the severity of nephrotoxicity. Cisplatin induced renal dysfunction, renal inflammation, apoptosis and tubular injury in wild-type mice, whereas PPAR-α deletion protected against these alterations. Moreover, we observed that cisplatin induced down-regulation of organic transporters MATE-1 and OCT-2 and that PPAR-α deletion restored the expression of these transporters. In addition, PPAR-α knockout mice at basal state showed increased MATE-1 expression and reduced OCT-2 levels. Here, we show for the first time that PPAR-α deletion protects against cisplatin nephrotoxicity and that this protection is via modulation of the organic transporters MATE-1 and OCT-2. Full article
Show Figures

Figure 1

Figure 1
<p>PPAR-α deletion attenuates cisplatin-induced increased pro-inflammatory cytokines and apoptosis-related genes. Cisplatin treatment (CP) increased mRNA levels of pro-inflammatory cytokines, (<b>A</b>) <span class="html-italic">TNF-</span><span class="html-italic">α</span>, (<b>B</b>) <span class="html-italic">IL-1</span><span class="html-italic">β</span> and (<b>C</b>) <span class="html-italic">IL-6</span> in renal tissue; PPAR-α knockout mice (CP PPARKO) prevented this increase. Apoptosis-related genes (<b>D</b>) <span class="html-italic">TNFR-2</span> and (<b>E</b>) <span class="html-italic">Bax/Bcl-2</span> ratio were also increased by cisplatin (CP) and PPAR-α deletion (CP PPARKO) avoided this increase. <span class="html-italic">n</span> = 5–6 per group. One-way ANOVA followed by post hoc Tukey’s test. * <span class="html-italic">p</span> &lt; 0.05 compared to the VEH group. # <span class="html-italic">p</span> &lt; 0.05; ## <span class="html-italic">p</span> &lt; 0.01 compared to the CP group.</p>
Full article ">Figure 2
<p>PPAR-α deletion attenuates tubular injury and apoptosis induced by cisplatin after 96 h. Representative photomicrography of H&amp;E staining. (<b>A</b>) CP treatment increases tubular injury while PPAR-α deletion attenuates it. (<b>B</b>) Immunofluorescence was performed to assess apoptosis. CP increases cleaved caspase-3 staining and CP PPARKO reverses this increase. In arrows is indicated tubules with the tubular lumen obstructed by the tubular casts and cell detachment from the tubular basement membrane. G to indicate glomeruli and a T for examples of tubules with normal structure, no cell detachment and free tubular lumen. <span class="html-italic">n</span> = 5 per group. Scale bar = 100 µm. One-way ANOVA followed by post hoc Tukey’s test. ** <span class="html-italic">p</span> &lt; 0.01, **** <span class="html-italic">p</span> &lt; 0.0001. compared to the VEH group. ## <span class="html-italic">p</span> &lt; 0.01, #### <span class="html-italic">p</span> &lt; 0.0001; compared to the CP group.</p>
Full article ">Figure 3
<p>PPAR-α knockout mice mitigate the decreased mRNA and protein expression by immunofluorescence of MATE-1. Ninety-six hours after cisplatin treatment (CP) downregulates (<b>A</b>) mRNA and (<b>B</b>) protein levels of MATE-1. PPAR-α knockout mice (CP PPARKO) prevented this downregulation. G to indicate glomeruli and a T to indicate tubules. <span class="html-italic">n</span> = 5 per group. One-way ANOVA followed by post hoc Tukey´s test. Scale bar = 100 µm. * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001 compared to the VEH group. # <span class="html-italic">p</span> &lt; 0.05, ### <span class="html-italic">p</span> &lt; 0.001 compared to the CP group.</p>
Full article ">Figure 4
<p>PPAR-α ablation attenuates downregulation of mRNA and protein expression by immunofluorescence of organic cations transporter 2 (OCT-2). Ninety-six hours after cisplatin treatment (CP) downregulates (<b>A</b>) mRNA and (<b>B</b>) protein (levels of OCT-2. PPAR-α knockout mice (CP PPARKO) attenuated this downregulation. G to indicate glomeruli and a T to indicate tubules. <span class="html-italic">n</span> = 5 per group. One-way ANOVA followed by post hoc Tukey´s test. Scale bar = 100 µm. * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001 compared to the VEH group. # <span class="html-italic">p</span> &lt; 0.05, ## <span class="html-italic">p</span> &lt; 0.01; compared to the CP group.</p>
Full article ">Figure 5
<p>PPAR-α absence enhances protein expression by immunofluorescence of multidrug and toxin extrusion 1 (MATE-1). (<b>A</b>) No differences between WT and PPARKO mice were found in <span class="html-italic">MATE-1</span> mRNA levels. (<b>B</b>) However, PPAR-α knockout mice enhanced MATE-1 protein levels. G to indicate glomeruli and a T to indicate tubules. <span class="html-italic">n</span> = 5 per group. Scale bar = 100 µm. Two-tailed Student´s t-test. * <span class="html-italic">p</span> &lt; 0.05, compared to the WT group.</p>
Full article ">Figure 6
<p>PPAR-α absence decreases mRNA and protein expression by immunofluorescence of organic cations transporter 2 (OCT-2). PPAR-α knockout mice presented reduced (<b>A</b>) mRNA and (<b>B</b>) protein levels of renal OCT-2. G to indicate glomeruli and a T to indicate tubules. <span class="html-italic">n</span> = 5 per group. Two-tailed Student´s t-test. Scale bar = 100 µm. * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001 compared to the WT group.</p>
Full article ">
26 pages, 2228 KiB  
Article
Serum-Based Proteomics Profiling in Adult Patients with Cystic Fibrosis
by Hicham Benabdelkamel, Hanadi Alamri, Meshail Okla, Afshan Masood, Mai Abdel Jabar, Ibrahim O. Alanazi, Assim A. Alfadda, Imran Nizami, Majed Dasouki and Anas M. Abdel Rahman
Int. J. Mol. Sci. 2020, 21(19), 7415; https://doi.org/10.3390/ijms21197415 - 8 Oct 2020
Cited by 13 | Viewed by 3631
Abstract
Cystic fibrosis (CF), the most common lethal autosomal recessive disorder among Caucasians, is caused by mutations in the CF transmembrane conductance regulator (CFTR) chloride channel gene. Despite significant advances in the management of CF patients, novel disease-related biomarkers and therapies must be identified. [...] Read more.
Cystic fibrosis (CF), the most common lethal autosomal recessive disorder among Caucasians, is caused by mutations in the CF transmembrane conductance regulator (CFTR) chloride channel gene. Despite significant advances in the management of CF patients, novel disease-related biomarkers and therapies must be identified. We performed serum proteomics profiling in CF patients (n = 28) and healthy subjects (n = 10) using the 2D-DIGE MALDI-TOF proteomic approach. Out of a total of 198 proteins identified, 134 showed a statistically significant difference in abundance and a 1.5-fold change (ANOVA, p < 0.05), including 80 proteins with increased abundance and 54 proteins with decreased abundance in CF patients. A multiple reaction monitoring-mass spectrometry analysis of six differentially expressed proteins identified by a proteomic approach (DIGE-MALD-MS) showed a significant increase in C3 and CP proteins and a decrease in APOA1, Complement C1, Hp, and RBP4proteins compared with healthy controls. Fifteen proteins were identified as potential biomarkers for CF diagnosis. An ingenuity pathway analysis of the differentially regulated proteins indicates that the central nodes dysregulated in CF subjects involve pro-inflammatory cytokines, ERK1/2, and P38 MAPK, which are primarily involved in catalytic activities and metabolic processes. The involved canonical pathways include those related to FXR/RXR, LXR/RXR, acute phase response, IL12, nitric oxide, and reactive oxygen species in macrophages. Our data support the current efforts toward augmenting protease inhibitors in patients with CF. Perturbations in lipid and vitamin metabolism frequently observed in CF patients may be partly due to abnormalities in their transport mechanism. Full article
(This article belongs to the Special Issue Biomarkers in Rare Diseases)
Show Figures

Figure 1

Figure 1
<p>Representative fluorescent protein profiles of 2D-DIGE containing a control sample labeled with Cy3 (<b>A</b>), a CF sample labeled with Cy5, (<b>B</b>) a pooled internal control labeled with Cy2, (<b>C</b>), and a merged 2D-DIGE comparison Cy3/Cy5 (<b>D</b>).</p>
Full article ">Figure 2
<p>Multiple reaction monitoring (MRM)mass spectrometry for validating the study findings. The MRM method based on signature peptides was developed to validate the expression of six proteins found in the proteomics approach (DIGE-MALD-MS). The expression of these six proteins in CF patients was expressed in fold changes compared with the healthy controls (Ctrl). The statistical significance was evaluated using an unpaired t-test (<span class="html-italic">n</span> = 10), in which * represents <span class="html-italic">p</span> &lt; 0.05, and **** represents <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 3
<p>Statistical analysis of proteomics expression for the CF patients compared with the healthy control subjects. A volcano plot between the Ctrl and CF groups shows the significantly dysregulated proteins (42 down-regulated and 22 up-regulated in the CF group), with cutoffs of 2 and 0.05 for the fold change (x-axes) and t-test, respectively (<b>A</b>). An orthogonal PLS-DA score plot with eight components indicates a significant separation between the study groups (Q2: 0.805, and R2: 0.971) for 1000 permutations due to the proteomics dysregulations (<b>B</b>).</p>
Full article ">Figure 4
<p>Biomarker statistical evaluation. An ROC exploratory analysis was generated by the PLS-DA model with many latent variable 2 and with increased sensitivity (x-axis) and specificity (y-axis), in which the area under the curve is at least 0.959 with a minimum combination of five variables (<b>A</b>). The features are ranked based on the selected frequency using the PLS-DA model, in which the color change from green to red indicates their relative expression of low to high, respectively (<b>B</b>); represents isoforms of the same protein found in different spots of the gel. Representative ROC curves for apolipoprotein A-I, which is down-regulated in CF (<b>C</b>), and protein-glutamine gamma-glutamyltransferase 6, which is up-regulated in CF (<b>D</b>).</p>
Full article ">Figure 5
<p>Schematic representation of the most significant IPA networks involving the proteins that were differentially regulated between the CF and control states. The IPA analysis found that the functional interaction networks pathway with the highest score (33) was related to “metabolic diseases, neurological diseases, and disease organismal injury and abnormalities.” This pathway incorporated pro-inflammatory cytokines, ERK1/2, and P38 MAPK as central nodes that were deregulated in CF samples. The nodes in green and red correspond to down-regulated and up-regulated proteins, respectively. The colorless nodes were proposed by the IPA and suggest potential targets that are functionally coordinated with the differentially abundant proteins (<b>A</b>). The solid lines indicate direct molecular interactions, and the dashed lines represent indirect interactions. The diagram shows the top six canonical pathways ranked by the P-values obtained from the IPA (<b>B</b>).</p>
Full article ">
25 pages, 5362 KiB  
Article
Investigating the Transition of Pre-Symptomatic to Symptomatic Huntington’s Disease Status Based on Omics Data
by Christiana C. Christodoulou, Margarita Zachariou, Marios Tomazou, Evangelos Karatzas, Christiana A. Demetriou, Eleni Zamba-Papanicolaou and George M. Spyrou
Int. J. Mol. Sci. 2020, 21(19), 7414; https://doi.org/10.3390/ijms21197414 - 8 Oct 2020
Cited by 23 | Viewed by 4062
Abstract
Huntington’s disease is a rare neurodegenerative disease caused by a cytosine–adenine–guanine (CAG) trinucleotide expansion in the Huntingtin (HTT) gene. Although Huntington’s disease (HD) is well studied, the pathophysiological mechanisms, genes and metabolites involved in HD remain poorly understood. Systems bioinformatics can [...] Read more.
Huntington’s disease is a rare neurodegenerative disease caused by a cytosine–adenine–guanine (CAG) trinucleotide expansion in the Huntingtin (HTT) gene. Although Huntington’s disease (HD) is well studied, the pathophysiological mechanisms, genes and metabolites involved in HD remain poorly understood. Systems bioinformatics can reveal synergistic relationships among different omics levels and enables the integration of biological data. It allows for the overall understanding of biological mechanisms, pathways, genes and metabolites involved in HD. The purpose of this study was to identify the differentially expressed genes (DEGs), pathways and metabolites as well as observe how these biological terms differ between the pre-symptomatic and symptomatic HD stages. A publicly available dataset from the Gene Expression Omnibus (GEO) was analyzed to obtain the DEGs for each HD stage, and gene co-expression networks were obtained for each HD stage. Network rewiring, highlights the nodes that change most their connectivity with their neighbors and infers their possible implication in the transition between different states. The CACNA1I gene was the mostly highly rewired node among pre-symptomatic and symptomatic HD network. Furthermore, we identified AF198444 to be common between the rewired genes and DEGs of symptomatic HD. CNTN6, DEK, LTN1, MST4, ZFYVE16, CEP135, DCAKD, MAP4K3, NUPL1 and RBM15 between the DEGs of pre-symptomatic and DEGs of symptomatic HD and CACNA1I, DNAJB14, EPS8L3, HSDL2, SNRPD3, SOX12, ACLY, ATF2, BAG5, ERBB4, FOCAD, GRAMD1C, LIN7C, MIR22, MTHFR, NABP1, NRG2, OTC, PRAMEF12, SLC30A10, STAG2 and Y16709 between the rewired genes and DEGs of pre-symptomatic HD. The proteins encoded by these genes are involved in various biological pathways such as phosphatidylinositol-4,5-bisphosphate 3-kinase activity, cAMP response element-binding protein binding, protein tyrosine kinase activity, voltage-gated calcium channel activity, ubiquitin protein ligase activity, adenosine triphosphate (ATP) binding, and protein serine/threonine kinase. Additionally, prominent molecular pathways for each HD stage were then obtained, and metabolites related to each pathway for both disease stages were identified. The transforming growth factor beta (TGF-β) signaling (pre-symptomatic and symptomatic stages of the disease), calcium (Ca2+) signaling (pre-symptomatic), dopaminergic synapse pathway (symptomatic HD patients) and Hippo signaling (pre-symptomatic) pathways were identified. The in silico metabolites we identified include Ca2+, inositol 1,4,5-trisphosphate, sphingosine 1-phosphate, dopamine, homovanillate and L-tyrosine. The genes, pathways and metabolites identified for each HD stage can provide a better understanding of the mechanisms that become altered in each disease stage. Our results can guide the development of therapies that may target the altered genes and metabolites of the perturbed pathways, leading to an improvement in clinical symptoms and hopefully a delay in the age of onset. Full article
Show Figures

Figure 1

Figure 1
<p>Network topological analysis of the gene co-expression. (<b>a</b>) Gene co-expression networks for controls versus pre-symptomatic and controls versus symptomatic HD. Blue nodes represent: the genes involved in pre-symptomatic HD, orange nodes represent: the genes involved in the symptomatic HD stage and green nodes represent the genes which appear in both HD networks. Edge colour represents co-expression in the respective groups (either or both HD stages) while edge thickness represents co-occurrence score (<b>b</b>–<b>e</b>) Distribution of the calculated centralities for the pre-symptomatic and symptomatic HD networks, i.e., (<b>b</b>) Degree (<b>c</b>) Betweenness (<b>d</b>) Coreness and (<b>e</b>) Closeness.</p>
Full article ">Figure 2
<p>Central reference network of the pre-symptomatic and symptomatic HD network using the Cytoscape plug-in DyNet. Dark red nodes: Most highly re-wired nodes, Medium red: Highly re-wired nodes, Light red: Least most re-wired nodes and White nodes: No re-wiring. The square node indicates the <span class="html-italic">CACNA1I</span> gene, which was the most highly re-wired node based on the DyNet re-wiring score.</p>
Full article ">Figure 3
<p>Venn diagram of rewired genes and DEGs of pre-symptomatic and symptomatic HD. Venn diagram illustrates the number of common genes between the rewired genes and DEGs of the two HD stages.</p>
Full article ">Figure 4
<p>Cluster of connected pathways for pre-symptomatic and symptomatic HD using PathwayConnector (<b>a</b>) Clusters of pathways in the pre-symptomatic HD stage. There is a total of six clusters, each shaded in a different color. (<b>b</b>) Three clusters of pathways in the symptomatic HD stage.</p>
Full article ">Figure 4 Cont.
<p>Cluster of connected pathways for pre-symptomatic and symptomatic HD using PathwayConnector (<b>a</b>) Clusters of pathways in the pre-symptomatic HD stage. There is a total of six clusters, each shaded in a different color. (<b>b</b>) Three clusters of pathways in the symptomatic HD stage.</p>
Full article ">Figure 5
<p>PathWalks derived pathway to pathway networks and odds ratio analysis for (<b>a</b>) pre-symptomatic versus symptomatic HD, (<b>b</b>) pre-symptomatic HD, (<b>c</b>) symptomatic HD. In each network the node size represents the odds ratio (OR) score in 4 bins. The top 20 pathways w.r.t to OR are shown in colour, while the remaining nodes are shown in grey. Edges represent walker transitions between pathways. Colour shading shows the identified communities of highly connected pathways w.r.t to PathWalks scores.</p>
Full article ">Figure 5 Cont.
<p>PathWalks derived pathway to pathway networks and odds ratio analysis for (<b>a</b>) pre-symptomatic versus symptomatic HD, (<b>b</b>) pre-symptomatic HD, (<b>c</b>) symptomatic HD. In each network the node size represents the odds ratio (OR) score in 4 bins. The top 20 pathways w.r.t to OR are shown in colour, while the remaining nodes are shown in grey. Edges represent walker transitions between pathways. Colour shading shows the identified communities of highly connected pathways w.r.t to PathWalks scores.</p>
Full article ">Figure 6
<p>Pathway-metabolite network with pathways and the common and exclusive metabolites for pre-symptomatic and symptomatic HD (<b>a</b>) Pathways and common and exclusive metabolites for pre-symptomatic HD (<b>b</b>) Pathways and exclusive metabolites for symptomatic HD. The top pathways are shown in colour. Green nodes represent the number of exclusive metabolites for either the pre-symptomatic or symptomatic HD stage. Blue nodes represent the common metabolites in both HD stages. The node size represents the number of common metabolites and smaller nodes represent; the smaller the number of metabolites and larger nodes, the greater the number of metabolites. Edge width represents the number of common metabolites across the pathways.</p>
Full article ">Figure 6 Cont.
<p>Pathway-metabolite network with pathways and the common and exclusive metabolites for pre-symptomatic and symptomatic HD (<b>a</b>) Pathways and common and exclusive metabolites for pre-symptomatic HD (<b>b</b>) Pathways and exclusive metabolites for symptomatic HD. The top pathways are shown in colour. Green nodes represent the number of exclusive metabolites for either the pre-symptomatic or symptomatic HD stage. Blue nodes represent the common metabolites in both HD stages. The node size represents the number of common metabolites and smaller nodes represent; the smaller the number of metabolites and larger nodes, the greater the number of metabolites. Edge width represents the number of common metabolites across the pathways.</p>
Full article ">Figure 7
<p>Flowchart of methodology and results.</p>
Full article ">
18 pages, 3284 KiB  
Review
Pericyte-Endothelial Interactions in the Retinal Microvasculature
by Hu Huang
Int. J. Mol. Sci. 2020, 21(19), 7413; https://doi.org/10.3390/ijms21197413 - 8 Oct 2020
Cited by 116 | Viewed by 10190
Abstract
Retinal microvasculature is crucial for the visual function of the neural retina. Pericytes and endothelial cells (ECs) are the two main cellular constituents in the retinal microvessels. Formation, maturation, and stabilization of the micro-vasculatures require pericyte-endothelial interactions, which are perturbed in many retinal [...] Read more.
Retinal microvasculature is crucial for the visual function of the neural retina. Pericytes and endothelial cells (ECs) are the two main cellular constituents in the retinal microvessels. Formation, maturation, and stabilization of the micro-vasculatures require pericyte-endothelial interactions, which are perturbed in many retinal vascular disorders, such as retinopathy of prematurity, retinal vein occlusion, and diabetic retinopathy. Understanding the cellular and molecular mechanisms of pericyte-endothelial interaction and perturbation can facilitate the design of therapeutic intervention for the prevention and treatment of retinal vascular disorders. Pericyte-endothelial interactions are indispensable for the integrity and functionality of retinal neurovascular unit (NVU), including vascular cells, retinal neurons, and glial cells. The essential autocrine and paracrine signaling pathways, such as Vascular endothelial growth factor (VEGF), Platelet-derived growth factor subunit B (PDGFB), Notch, Angipointein, Norrin, and Transforming growth factor-beta (TGF-β), have been well characterized for the regulation of pericyte-endothelial interactions in the neo-vessel formation processes (vasculogenesis and angiogenesis) during embryonic development. They also play a vital role in stabilizing and remodeling mature vasculature under pathological conditions. Awry signals, aberrant metabolisms, and pathological conditions, such as oxidative stress and inflammation, can disrupt the communication between pericytes and endothelial cells, thereby resulting in the breakdown of the blood-retinal barrier (BRB) and other microangiopathies. The emerging evidence supports extracellular exosomes’ roles in the (mis)communications between the two cell types. This review summarizes the essential knowledge and updates about new advancements in pericyte-EC interaction and communication, emphasizing the retinal microvasculature. Full article
Show Figures

Figure 1

Figure 1
<p>Retinal microvasculature in the adult mouse retina. The retinas were dissected from the adult mice and processed with clarity method. The cleared retinas were performed with the immunofluorescence staining of anti-PECAM1/CD31 primary antibody and Alexa Fluor 594 secondary antibody (red). The three-dimensional (3D) architectures of retinal vasculatures are visualized with low (<b>A</b>) and high magnification (<b>B</b>). One points to superficial vascular plexus at the nerve fiber layer. Two points to intermediate vascular plexus at the inner plexiform form layer. Three points to deep vascular plexus at the outer plexiform layer. The outer inner nuclear layer and the inner nuclear layer were stained with DAPI. The DAPI signals in Panel B were shown with pseudocolor (grey).</p>
Full article ">Figure 2
<p>Pericytes-endothelial cell interactions and abnormalities in the retinal microvasculature. (<b>A</b>) Overview of an entire retinal microvasculature made from the adult mouse retina with the trypsin digestion method. The circles divided the vascular network into the periphery, middle, and central zones. (<b>B</b>) The high magnitude of retinal vasculature indicates the close interactions of pericytes and endothelial cells in the normal blood vessel walls. Red arrows point to the nuclei of endothelial cells. Green arrows point to the nuclei of pericytes. (<b>C</b>) The retinal vasculatures are made from the CXCR5 knockout mouse, which was subjected to ischemia-reperfusion injury, leading to a substantial loss of endothelial cells (ECs) and pericytes. Arrows point to the acellular capillary. This image is excerpted from a previous publication with the journal’s permission [<a href="#B22-ijms-21-07413" class="html-bibr">22</a>].</p>
Full article ">Figure 3
<p>The proposed regulation of P<span class="html-italic">l</span>GF and TGF-β in early diabetic retinopathy. Diabetes upregulates both P<span class="html-italic">l</span>GF and TGF-β in endothelial cells. P<span class="html-italic">l</span>GF can promote early diabetic retinopathy through the activation of VEGFR1 and Erk1/2 signaling, and the expression of downstream target genes, such as G6pdh (pentose phosphate pathway), Prdx3 and 6 (antioxidants), as well as the tight and adhesion junction genes (Cadh5, ZO1, and occludin). The transcription factor(s) that regulate downstream genes’ expressions are to be identified (question maker). Increased TGF-β by diabetes can protect the retina from diabetic injury in the early disease phase through the activation of TGFBR2/ALK5 signaling, which regulates the nuclear translocation of Smad2/3 and then activates the transcription of downstream target genes, such as Egf2, Edn2, and Pcam1. TGF-β can regulate P<span class="html-italic">l</span>GF through the unknown factor (question maker). Arrow lines: stimulation. Blunt line: inhibition.</p>
Full article ">Figure 4
<p>The features of vascular organoids derived from human-induced pluripotent stem cells. (<b>A</b>–<b>C</b>) The vascular organoids are differentiated from human induced pluripotent stem cells (iPSCs). The 10-micron cryosections are prepared and stained with the endothelial cell (ECs) marker CD31 (or PECAM1, <b>A</b>, red) and the pericyte marker PDGFRb (<b>B</b>, green). The nuclei were stained with DAPI (blue). (<b>C</b>) The merged image shows the co-localization of the two markers in the blood vessels. (<b>D</b>–<b>E</b>) the enlarged images of the boxed areas in panels <b>A</b>–<b>C</b>. Note that the integral vascular networks of organoids vasculatures with vessel lumens, ECs, and pericytes in the blood vessels. Human organoids can be an excellent model to study EC-pericyte interaction and other human vascular diseases’ pathophysiology.</p>
Full article ">Figure 5
<p>Co-localization analysis of pericytes and endothelial cells in the organoid vasculature. Part of the organoid vasculature (1/6) is used for analysis with the ImageJ software and EzColocalization plugin [<a href="#B96-ijms-21-07413" class="html-bibr">96</a>]. (<b>A</b>,<b>B</b>) The heat maps show the localization of PDGFRB and CD31 staining signals in the vessels. The scale bars indicate the signal intensity from low to high. (<b>C</b>) The scatterplot shows the relationship between the signal intensity for PDGFRB-Alex fluor 488 and CD31-Cy 5 channels. Note that (1) the numbers of data represent the vessel segments identified by the software, and (2) data with similar signal intensity for each channel indicate colocalization versus anti-colocalization by the differences in signal intensity. (<b>D</b>) The metric matrix for the threshold overlap score (TOS) linear median values. The X-axis and Y-axis values are the top percentile (F<sub>T</sub>) of threshold pixels for signal intensity. The black color is not informative. The red color indicates co-localization.</p>
Full article ">
19 pages, 2167 KiB  
Review
p38 MAPK Pathway in the Heart: New Insights in Health and Disease
by Rafael Romero-Becerra, Ayelén M. Santamans, Cintia Folgueira and Guadalupe Sabio
Int. J. Mol. Sci. 2020, 21(19), 7412; https://doi.org/10.3390/ijms21197412 - 8 Oct 2020
Cited by 89 | Viewed by 8014
Abstract
The p38 mitogen-activated kinase (MAPK) family controls cell adaptation to stress stimuli. p38 function has been studied in depth in relation to cardiac development and function. The first isoform demonstrated to play an important role in cardiac development was p38α; however, all p38 [...] Read more.
The p38 mitogen-activated kinase (MAPK) family controls cell adaptation to stress stimuli. p38 function has been studied in depth in relation to cardiac development and function. The first isoform demonstrated to play an important role in cardiac development was p38α; however, all p38 family members are now known to collaborate in different aspects of cardiomyocyte differentiation and growth. p38 family members have been proposed to have protective and deleterious actions in the stressed myocardium, with the outcome of their action in part dependent on the model system under study and the identity of the activated p38 family member. Most studies to date have been performed with inhibitors that are not isoform-specific, and, consequently, knowledge remains very limited about how the different p38s control cardiac physiology and respond to cardiac stress. In this review, we summarize the current understanding of the role of the p38 pathway in cardiac physiology and discuss recent advances in the field. Full article
(This article belongs to the Special Issue P38 Signaling Pathway)
Show Figures

Figure 1

Figure 1
<p>p38 in cardiovascular regeneration. <b>(A)</b> p38α blocks cardiovascular regeneration by inhibiting the expression of genes involved in cardiomyocyte proliferation and regeneration, such as <span class="html-italic">Ect2,</span> <span class="html-italic">Crp1, ki67, cdc2, cyclin A</span>, and p27, and reducing Rb phosphorylation to block cell-cycle progression. FGF1 stimulation has the opposite effect. <b>(B)</b> p38γ activates differentiation and myogenesis in satellite cells by phosphorylating MyoD and activating proliferation. p38α prevents p38γ activation in satellite cells, blocking regeneration. ↑ increase, ↓ decrease.</p>
Full article ">Figure 2
<p>Dual role of p38 activation during preconditioning and ischemia–reperfusion injury. Activation of p38β during preconditioning triggers pro-survival signaling pathways, whereas decreased p38α activation during the ischemic episode leads to cardioprotection. On the other hand, ROS-induced p38α activation during the ischemic insult triggers HIF1-α stabilization; increases (↑) fibrosis, arrhythmias, and inflammation; and disrupts mitochondrial homeostasis. SB203580 administration during preconditioning increases myocardial injury, whereas administration during ischemia–reperfusion improves cardiac outcome. Indirect p38 downregulators, such as gamboge, statins, and antioxidants, seem to have beneficial effects when administered during or after the ischemia. Further research is needed to determine the precise reciprocity of ROS–p38 regulation. ↑ increase.</p>
Full article ">Figure 3
<p>p38 in heart failure and cardiac arrhythmia. p38 activation participates in the development of heart failure and cardiac arrhythmia through three main mechanisms: <b>(A)</b> Increased cardiac fibrosis by induction of TNF-α and IL-6 in cardiomyocytes, differentiation of fibroblasts, and induction of TGF-β in cardiac myofibroblasts; <b>(B)</b> Reduced cardiac contractility due to dephosphorylation of a-tropomyosin and troponin I via PP2C-α/PP2C-β; <b>(C)</b> Promotion of cardiac arrhythmias due to reduced expression and activity of SERCA2 (<span class="html-italic">Atp2a2</span>), increased expression of NCX1 (<span class="html-italic">Ncx1</span>) and Cx43 (<span class="html-italic">Cnx43</span>), and altered Cx43 phosphorylation, inducing cardiac contractile dysfunction and altered action potential propagation. ↑ increase, ↓ decrease.</p>
Full article ">
37 pages, 4518 KiB  
Review
Amphiphilic Aminoglycosides as Medicinal Agents
by Clément Dezanet, Julie Kempf, Marie-Paule Mingeot-Leclercq and Jean-Luc Décout
Int. J. Mol. Sci. 2020, 21(19), 7411; https://doi.org/10.3390/ijms21197411 - 8 Oct 2020
Cited by 15 | Viewed by 4211
Abstract
The conjugation of hydrophobic group(s) to the polycationic hydrophilic core of the antibiotic drugs aminoglycosides (AGs), targeting ribosomal RNA, has led to the development of amphiphilic aminoglycosides (AAGs). These drugs exhibit numerous biological effects, including good antibacterial effects against susceptible and multidrug-resistant bacteria [...] Read more.
The conjugation of hydrophobic group(s) to the polycationic hydrophilic core of the antibiotic drugs aminoglycosides (AGs), targeting ribosomal RNA, has led to the development of amphiphilic aminoglycosides (AAGs). These drugs exhibit numerous biological effects, including good antibacterial effects against susceptible and multidrug-resistant bacteria due to the targeting of bacterial membranes. In the first part of this review, we summarize our work in identifying and developing broad-spectrum antibacterial AAGs that constitute a new class of antibiotic agents acting on bacterial membranes. The target-shift strongly improves antibiotic activity against bacterial strains that are resistant to the parent AG drugs and to antibiotic drugs of other classes, and renders the emergence of resistant Pseudomonas aeruginosa strains highly difficult. Structure–activity and structure–eukaryotic cytotoxicity relationships, specificity and barriers that need to be crossed in their development as antibacterial agents are delineated, with a focus on their targets in membranes, lipopolysaccharides (LPS) and cardiolipin (CL), and the corresponding mode of action against Gram-negative bacteria. At the end of the first part, we summarize the other recent advances in the field of antibacterial AAGs, mainly published since 2016, with an emphasis on the emerging AAGs which are made of an AG core conjugated to an adjuvant or an antibiotic drug of another class (antibiotic hybrids). In the second part, we briefly illustrate other biological and biochemical effects of AAGs, i.e., their antifungal activity, their use as delivery vehicles of nucleic acids, of short peptide (polyamide) nucleic acids (PNAs) and of drugs, as well as their ability to cleave DNA at abasic sites and to inhibit the functioning of connexin hemichannels. Finally, we discuss some aspects of structure–activity relationships in order to explain and improve the target selectivity of AAGs. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Structures of natural antibiotic aminoglycosides <b>1</b>–<b>5</b>, of some corresponding constitutive derivatives <b>6</b>–<b>8</b> and of synthetic intermediates used to prepare amphiphilic aminoglycosides (AAGs) <b>9</b>–<b>13</b> (Tr = trityl group = triphenylmethyl, PMB = <span class="html-italic">para</span>-methoxyphenyl group).</p>
Full article ">Figure 2
<p>Structure of polymyxin E (COL), showing the five amine functions protonated at physiological pH.</p>
Full article ">Figure 3
<p>Structures of the first identified broad-spectrum antibacterial amphiphilic neamine (NEA) (<b>6</b>) and paromamine (PARA) (<b>7</b>) derivatives [<a href="#B5-ijms-21-07411" class="html-bibr">5</a>,<a href="#B34-ijms-21-07411" class="html-bibr">34</a>].</p>
Full article ">Figure 4
<p>Structures of the identified broad-spectrum antibacterial amphiphilic dialkyl (<b>24</b> and <b>25</b>) and dialkylnaphthyl (<b>26</b>–<b>28</b>) NEA derivatives [<a href="#B34-ijms-21-07411" class="html-bibr">34</a>].</p>
Full article ">Figure 5
<p>Structures of the broad-spectrum antibacterial 3′,4′-dinonyl NEA derivative and of the corresponding antibacterial analogues synthesized in the 1-methyl neosamine series [<a href="#B36-ijms-21-07411" class="html-bibr">36</a>].</p>
Full article ">Figure 6
<p>Structures of antibacterial homodialkyl (<b>38</b>–<b>41</b>, <b>46</b>, <b>47</b>) and heterodialkyl (<b>42</b>–<b>45</b>) NEA derivatives synthesized [<a href="#B35-ijms-21-07411" class="html-bibr">35</a>].</p>
Full article ">Figure 7
<p>Values of 1/(MIC (mL/µg) as a function of clogP values for 3′,6-dinaphthylalkyl NEAs (di2NM <b>14</b>, di2NP <b>26</b>, di2NB <b>27</b> and di2-naphthylhexyl) and 3′,6-dialkyl NEAs (diC4, diC6, diC7 <b>38</b>, diC8 <b>39</b>, diC9 <b>24</b>, diC10 <b>40</b>, diC11 <b>41</b> and diC18). (<b>A</b>) Against MRSA; (<b>B</b>) against susceptible <span class="html-italic">P. aeruginosa</span> ATCC 27853. Naphthylalkyl derivatives: red squares; alkyl derivatives: green triangles [<a href="#B35-ijms-21-07411" class="html-bibr">35</a>].</p>
Full article ">Figure 8
<p>Comparison of the structures of lipid A and cardiolipin (CL) to the structure of the antibacterial 3′,6-dinonyl NEA derivative <b>24</b>.</p>
Full article ">Figure 9
<p>Structures of the tobramycin (TOB) conjugates to lysine <b>48</b> [<a href="#B105-ijms-21-07411" class="html-bibr">105</a>,<a href="#B106-ijms-21-07411" class="html-bibr">106</a>], and to the efflux pump inhibitors (EPIs), 1-(1′-naphthylmethyl)piperazine (NMP) (<b>49</b>), paroxetine (PAR) (<b>50</b>) and dibasic naphthyl peptide (DBP) (<b>51</b>) [<a href="#B107-ijms-21-07411" class="html-bibr">107</a>,<a href="#B108-ijms-21-07411" class="html-bibr">108</a>].</p>
Full article ">Figure 10
<p>Structures of the TOB (<b>52</b>–<b>54</b>) and nebramine (NEB) (<b>56</b>) conjugates to the fluoroquinolones moxifloxacin (MOX) and ciprofloxacin (CIP) [<a href="#B109-ijms-21-07411" class="html-bibr">109</a>,<a href="#B111-ijms-21-07411" class="html-bibr">111</a>], respectively <b>55</b> and <b>57</b>, and, of the NEB conjugates to the efflux pump inhibitor 1-(1′-naphthylmethyl)piperazine (NEB-NMP) <b>58</b> [<a href="#B112-ijms-21-07411" class="html-bibr">112</a>].</p>
Full article ">Figure 11
<p>Structures of the TOB-cyclam and NEB-cyclam hybrids <b>59</b> [<a href="#B115-ijms-21-07411" class="html-bibr">115</a>] and <b>60</b> [<a href="#B116-ijms-21-07411" class="html-bibr">116</a>].</p>
Full article ">Figure 12
<p>Structures of the synthesized broad-spectrum antibacterial 4′,5,6-tri- and 4′,5-di-alkylated NEB derivatives [<a href="#B37-ijms-21-07411" class="html-bibr">37</a>,<a href="#B38-ijms-21-07411" class="html-bibr">38</a>].</p>
Full article ">Figure 13
<p>Structure of the amphiphilic aminoglycoside K20, capable of inhibiting many fungal species such as <span class="html-italic">Fusarium graminearum</span>, the causal agent wheat <span class="html-italic">Fusarium</span> head blight (FHB) [<a href="#B125-ijms-21-07411" class="html-bibr">125</a>,<a href="#B126-ijms-21-07411" class="html-bibr">126</a>,<a href="#B127-ijms-21-07411" class="html-bibr">127</a>,<a href="#B128-ijms-21-07411" class="html-bibr">128</a>].</p>
Full article ">Figure 14
<p>Structure of one of the kanamycin (KANA)-cholesterol conjugates, <b>61</b>, developed for gene transfection [<a href="#B130-ijms-21-07411" class="html-bibr">130</a>].</p>
Full article ">Figure 15
<p>Structures of the most efficient NEA-based vectors for gene transfection [<a href="#B134-ijms-21-07411" class="html-bibr">134</a>].</p>
Full article ">Figure 16
<p>Structures of the AAGs developed for small interfering RNA (siRNA) delivery made of the TOB, KANA, PARO and NEO cores, respectively, linked to two dioleyl chains by a succinyl spacer [<a href="#B135-ijms-21-07411" class="html-bibr">135</a>].</p>
Full article ">Figure 17
<p>Structures of the anti-HIV (anti-TAR RNA) PNA conjugates to NEA [<a href="#B145-ijms-21-07411" class="html-bibr">145</a>,<a href="#B149-ijms-21-07411" class="html-bibr">149</a>] and to 6-amino-6-deoxy-1-methylglucosamine (1-methyl neosamine) [<a href="#B151-ijms-21-07411" class="html-bibr">151</a>].</p>
Full article ">Figure 18
<p>Structure of the most efficient DNA-cleaving AAG identified, <b>76,</b> at abasic sites [<a href="#B157-ijms-21-07411" class="html-bibr">157</a>], and of the amphiphilic azobenzene-NEO conjugate <b>77</b> forming nanostructures [<a href="#B158-ijms-21-07411" class="html-bibr">158</a>].</p>
Full article ">Figure 19
<p>Structures of AAGs that are non-bactericidal and non-toxic or moderately toxic to mammalian HeLa cells, which are connexin hemichannel (HC) inhibitors [<a href="#B160-ijms-21-07411" class="html-bibr">160</a>].</p>
Full article ">
16 pages, 11529 KiB  
Article
Self-Replication of Prion Protein Fragment 89-230 Amyloid Fibrils Accelerated by Prion Protein Fragment 107-143 Aggregates
by Tomas Sneideris, Mantas Ziaunys, Brett K.-Y. Chu, Rita P.-Y. Chen and Vytautas Smirnovas
Int. J. Mol. Sci. 2020, 21(19), 7410; https://doi.org/10.3390/ijms21197410 - 8 Oct 2020
Cited by 4 | Viewed by 3250
Abstract
Prion protein amyloid aggregates are associated with infectious neurodegenerative diseases, known as transmissible spongiform encephalopathies. Self-replication of amyloid structures by refolding of native protein molecules is the probable mechanism of disease transmission. Amyloid fibril formation and self-replication can be affected by many different [...] Read more.
Prion protein amyloid aggregates are associated with infectious neurodegenerative diseases, known as transmissible spongiform encephalopathies. Self-replication of amyloid structures by refolding of native protein molecules is the probable mechanism of disease transmission. Amyloid fibril formation and self-replication can be affected by many different factors, including other amyloid proteins and peptides. Mouse prion protein fragments 107-143 (PrP(107-143)) and 89-230 (PrP(89-230)) can form amyloid fibrils. β-sheet core in PrP(89-230) amyloid fibrils is limited to residues ∼160–220 with unstructured N-terminus. We employed chemical kinetics tools, atomic force microscopy and Fourier-transform infrared spectroscopy, to investigate the effects of mouse prion protein fragment 107-143 fibrils on the aggregation of PrP(89-230). The data suggest that amyloid aggregates of a short prion-derived peptide are not able to seed PrP(89-230) aggregation; however, they accelerate the self-replication of PrP(89-230) amyloid fibrils. We conclude that PrP(107-143) fibrils could facilitate the self-replication of PrP(89-230) amyloid fibrils in several possible ways, and that this process deserves more attention as it may play an important role in amyloid propagation. Full article
(This article belongs to the Special Issue Amyloid Hetero-Aggregation)
Show Figures

Figure 1

Figure 1
<p>Representative curves of PrP-seed-induced aggregation reaction kinetics performed in the absence (<b>a</b>) or presence (<b>b</b>) of PrP(107-143) fibrils. Logarithmic plot of <math display="inline"><semantics> <msub> <mi>t</mi> <mrow> <mi>l</mi> <mi>a</mi> <mi>g</mi> </mrow> </msub> </semantics></math> values of seed-induced aggregation reaction performed in the absence and presence of 10% of PrP(107-143) fibrils (<b>c</b>). Relative <math display="inline"><semantics> <msub> <mi>t</mi> <mrow> <mi>l</mi> <mi>a</mi> <mi>g</mi> </mrow> </msub> </semantics></math> values (ratio between <math display="inline"><semantics> <msub> <mi>t</mi> <mrow> <mi>l</mi> <mi>a</mi> <mi>g</mi> </mrow> </msub> </semantics></math> in the absence and presence of 10% peptide fibrils) (<b>d</b>). Dependence of PrP(89-230)-seed-induced aggregation reaction kinetics on the initial concentration of PrP(107-143) fibrils (<b>e</b>,<b>f</b>). Error bars are standard deviations (<span class="html-italic">n</span> = 9).</p>
Full article ">Figure 2
<p>AFM images of PrP(107-143) and PrP(89-230) aggregates. Images of the spontaneously formed PrP(107-143) (<b>a</b>) and PrP(89-230) (<b>b</b>) aggregates. Images of sonicated PrP(107-143) (<b>c</b>) and PrP(89-230) (<b>d</b>) aggregates. Images of aggregates formed during PrP(107-143)-fibril-induced (<b>e</b>) or PrP(89-230)-seed-induced (<b>f</b>) aggregation reaction. Images of aggregates formed in the presence of both PrP(107-143)-fibrils and PrP(89-230)-seed (<b>g</b>).</p>
Full article ">Figure 3
<p>Absorbance and second derivative (inset) FTIR spectra of PrP(89-230) and PrP(107-143) fibrils. 1.—Spontaneously formed PrP(89-230) fibrils. 2.—Fibrils formed in the presence of 1% of PrP(89-230) fibrils. 3.—Fibrils formed in the presence of 10% of PrP(107-143) fibrils and 1% of PrP(89-230) fibrils. 4.—Spontaneously formed PrP(107-143) fibrils.</p>
Full article ">Figure A1
<p>AFM images (<b>a</b>,<b>b</b>), FTIR spectra (<b>c</b>) and PK digestion SDS-PAGE (<b>d</b>) of PrP fibrils. (<b>a</b>) AFM image of PrP(89-230) fibrils. (<b>b</b>) AFM image of PrP(89-230) fibrils formed in the presence of 5% of PrP(107-143) aggregates. (<b>c</b>) FTIR spectra of PrP(89-230) fibrils formed in the absence or presence of PrP(107-143) fibrils, and PrP(107-143) fibrils alone. (<b>d</b>) MM—molecular weight marker; 1—PrP(89-230) fibrils; 2—PrP(89-230) fibrils formed in the presence of 5% of PrP(107-143) aggregates; time denotes PK digestion time. For PK digestion experiments, fibril samples (0.5 mg/mL) were centrifuged at 10,000× <span class="html-italic">g</span> for 30 min. Subsequently, fibrils were resuspended in 50 mM Tris buffer solution (pH 8) and centrifuged again. Then fibrils were resuspended in 200 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>L of 50 mM Tris buffer solution (pH 8) and sonicated for 30 s using Bandelin Sonopuls ultrasonic homogenizer equipped with a MS 72 tip (using 20% of the power, total energy applied to the sample ∼0.36 kJ). After sonication, the fibril solution was supplemented with 2.5 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>L of 2 mg/mL Proteinase K and incubated for 0 min, 5 min and 30 min at 37 <math display="inline"><semantics> <msup> <mrow/> <mo>∘</mo> </msup> </semantics></math>C with 600 RPM agitation in a Ditabis thermomixer MHR 23. Then, 17 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>L of each sample was collected, supplemented with 3 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>L of 10 mM PMSF, 96% EtOH and 20 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>L of 2 × SDS-Page sample buffer containing 6 M Urea. Samples were heated for 15 min at 98 <math display="inline"><semantics> <msup> <mrow/> <mo>∘</mo> </msup> </semantics></math>C and subsequently analyzed via Tricine–SDS-PAGE [<a href="#B40-ijms-21-07410" class="html-bibr">40</a>].</p>
Full article ">Figure A2
<p>Representative curves of PrP(107-143)-fibril-induced aggregation reaction kinetics monitored at 37 <math display="inline"><semantics> <msup> <mrow/> <mo>∘</mo> </msup> </semantics></math>C under quiescent conditions. For aggregation experiments, solutions of 0.5 mg/mL of monomeric PrP(89-230) protein in 1 M GuHCl, 3 M Urea, 1 × PBS (pH 7.4), containing 50 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>M ThT were supplemented with none, 0.025, 0.050, or 0.075 mg/mL of PrP(107-143) fibrils and incubated under quiescent conditions at 37 <math display="inline"><semantics> <msup> <mrow/> <mo>∘</mo> </msup> </semantics></math>C temperature. Aggregation kinetics were monitored by ThT fluorescence assay (excitation at 470 nm, emission at 510 nm) using Qiagen Rotor-Gene Q real-time analyzer.</p>
Full article ">Figure A3
<p>Average kinetic curves (<b>a</b>) and <span class="html-italic">t</span><math display="inline"><semantics> <msub> <mrow/> <mrow> <mi>l</mi> <mi>a</mi> <mi>g</mi> </mrow> </msub> </semantics></math> values (<b>b</b>) of PrP(89-230) self-replication reaction performed with introduction of PrP(107-143) seed at different time-points of the reaction. Then, 10 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>L of 0.5 mg/mL PrP(107-143) fibrils were added to 100 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>L of PrP(89-230) solution (0.5 mg/mL of monomeric PrP(89-230) and 0.001 mg/mL PrP(89-230) fibrils) at a different time-points. The reaction kinetics were monitored by measuring ThT fluorescence intensity (ex. 440 nm; em. 480 nm) at 60 <math display="inline"><semantics> <msup> <mrow/> <mo>∘</mo> </msup> </semantics></math>C through a bottom of non-biding 96-well plate (Corning No. 3881, New York, United States) using ClarioStar Plus (BMG Labtech, Ortenberg, Germany) microplate reader. The measurements were performed every three minutes. The error bars are standard deviations estimated from 3 repeats.</p>
Full article ">Figure A4
<p>Average kinetic curves (<b>a</b>) and <span class="html-italic">t</span><math display="inline"><semantics> <msub> <mrow/> <mrow> <mi>l</mi> <mi>a</mi> <mi>g</mi> </mrow> </msub> </semantics></math> values (<b>b</b>) of PrP(89-230) self-replication reaction performed with introduction of PrP(89-230) seed at different time-points of the reaction. Then, 10 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>L of 0.01 mg/mL PrP(107-143) fibrils were added to 100 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>L of PrP(89-230) solution (0.5 mg/mL of monomeric PrP(89-230) and 0.05 mg/mL PrP(107-143) fibrils) at a different time-points. The reaction kinetics were monitored by measuring ThT fluorescence intensity (ex. 440 nm; em. 480 nm) at 60 <math display="inline"><semantics> <msup> <mrow/> <mo>∘</mo> </msup> </semantics></math>C through a bottom of non-biding 96-well plate (Corning No. 3881, New York, United States) using ClarioStar Plus (BMG Labtech, Ortenberg, Germany) microplate reader. The measurements were performed every ten minutes (the measurement frequency was reduced to avoid agitation-induced formation of PrP(89-230) aggregates, which in the presence of PrP(107-143) fibrils alone could lead to the false-positive results). The error bars are standard deviations estimated from 4 repeats.</p>
Full article ">Figure A5
<p>AFM images of PrP(107-143) and PrP(89-230) aggregates. Images of the spontaneously formed PrP(107-143) (<b>a</b>) and PrP(89-230) (<b>b</b>) aggregates. Images of sonicated PrP(107-143) (<b>c</b>) and PrP(89-230) (<b>d</b>) aggregates. Images of aggregates formed during PrP(107-143)-fibril-induced (<b>e</b>) or PrP(89-230)-seed-induced (<b>f</b>) aggregation reaction. Image of aggregates formed in the presence of both PrP(107-143)-fibrils and PrP(89-230)-seed (<b>g</b>).</p>
Full article ">Figure A6
<p>Comparison of <math display="inline"><semantics> <msub> <mi>t</mi> <mrow> <mi>l</mi> <mi>a</mi> <mi>g</mi> </mrow> </msub> </semantics></math> values of seed-induced aggregation reaction performed in the absence and presence of 10% of PrP(107-143) fibrils.</p>
Full article ">Figure A7
<p>Representative curves of PrP(89-230) aggregation kinetics in the presence of monomeric PrP(107-143). For aggregation experiments, solutions of 0.5 mg/mL of monomeric PrP(89-230) in 0.5 M GuHCl, 50 mM phosphate buffer (pH 6), containing 50 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>M ThT were supplemented with either 0.001 mg/mL of PrP(89-230) seed, 0.05 mg/mL of PrP(107-143) monomers, or both 0.001 mg/mL of PrP(89-230) seed and 0.05 mg/mL of PrP(107-143) monomers. Aggregation reaction was performed under quiescent conditions at a 60 <math display="inline"><semantics> <msup> <mrow/> <mo>∘</mo> </msup> </semantics></math>C temperature. Aggregation kinetics were monitored by ThT fluorescence assay (excitation at 470 nm, emission at 510 nm) using Qiagen Rotor-Gene Q real-time analyzer.</p>
Full article ">Figure A8
<p>Measurements of PrP(89-230) concentration bound to the fibrils formed by PrP(107-143). For the measurements of monomeric PrP(89-230) binding to PrP(107-143) fibrils, 1 mL solutions of 0.5 mg/mL of monomeric PrP(89-230) in 0.5 M GuHCl, 50 mM phosphate buffer (pH 6) supplemented with none or 0.05 mg/mL of PrP(107-143) fibrils were centrifuged at 12,000× <span class="html-italic">g</span> for 30 min. Subsequently, concentration of monomeric PrP(89-230) remaining in the supernatant was determined by measuring UV-absorption at 280 nm using NanoDrop 2000 (Thermo Fisher Scientific). Error bars are standard deviations estimated from three repeats.</p>
Full article ">Figure A9
<p>Kinetics of PrP(89-230)-seed-induced aggregation reaction performed in the absence (<b>a</b>) or presence (<b>b</b>) of PrP(107-143) fibrils. Continuous lines denote the fit obtained by fitting “Saturated Elongation and Secondary Nucleation” model to the experimental data using the AmyloFit software [<a href="#B13-ijms-21-07410" class="html-bibr">13</a>].</p>
Full article ">
14 pages, 627 KiB  
Review
Genetic Aspects of Inflammation and Immune Response in Stroke
by Dejan Nikolic, Milena Jankovic, Bojana Petrovic and Ivana Novakovic
Int. J. Mol. Sci. 2020, 21(19), 7409; https://doi.org/10.3390/ijms21197409 - 8 Oct 2020
Cited by 22 | Viewed by 4979
Abstract
Genetic determinants play important role in the complex processes of inflammation and immune response in stroke and could be studied in different ways. Inflammation and immunomodulation are associated with repair processes in ischemic stroke, and together with the concept of preconditioning are promising [...] Read more.
Genetic determinants play important role in the complex processes of inflammation and immune response in stroke and could be studied in different ways. Inflammation and immunomodulation are associated with repair processes in ischemic stroke, and together with the concept of preconditioning are promising modes of stroke treatment. One of the important aspects to be considered in the recovery of patients after the stroke is a genetic predisposition, which has been studied extensively. Polymorphisms in a number of candidate genes, such as IL-6, BDNF, COX2, CYPC19, and GPIIIa could be associated with stroke outcome and recovery. Recent GWAS studies pointed to the variant in genesPATJ and LOC as new genetic markers of long term outcome. Epigenetic regulation of immune response in stroke is also important, with mechanisms of histone modifications, DNA methylation, and activity of non-coding RNAs. These complex processes are changing from acute phase over the repair to establishing homeostasis or to provoke exaggerated reaction and death. Pharmacogenetics and pharmacogenomics of stroke cures might also be evaluated in the context of immuno-inflammation and brain plasticity. Potential novel genetic treatment modalities are challenged but still in the early phase of the investigation. Full article
(This article belongs to the Special Issue Immunoinflammatory Background of Neuronal Damage in Stroke)
Show Figures

Figure 1

Figure 1
<p>Spatio-temporal pattern of epigenetic regulation in immune response in stroke. The scheme is highlighting multidimensional relations of the immune system and epigenetic mechanisms in different stroke zones over time. Stroke is unequally affecting brain tissues, with irreversible neuronal damage in stroke core and metabolic changes with the possibility counteract tissue injury in the penumbra area. Additionally, stroke is provoking immune response leading to inflammation and starting an immune cascade with consequences not only in the brain but also in the whole organism. Alteration of immune function is conducted through complex epigenetic regulation which is sensitive to temporal changes in the tissue microenvironment.</p>
Full article ">
18 pages, 3478 KiB  
Article
Rose Bengal Crosslinking to Stabilize Collagen Sheets and Generate Modulated Collagen Laminates
by Stefanie Eckes, Joy Braun, Julia S. Wack, Ulrike Ritz, Daniela Nickel and Katja Schmitz
Int. J. Mol. Sci. 2020, 21(19), 7408; https://doi.org/10.3390/ijms21197408 - 8 Oct 2020
Cited by 6 | Viewed by 3651
Abstract
For medical application, easily accessible biomaterials with tailored properties are desirable. Collagen type I represents a biomaterial of choice for regenerative medicine and tissue engineering. Here, we present a simple method to modify the properties of collagen and to generate collagen laminates. We [...] Read more.
For medical application, easily accessible biomaterials with tailored properties are desirable. Collagen type I represents a biomaterial of choice for regenerative medicine and tissue engineering. Here, we present a simple method to modify the properties of collagen and to generate collagen laminates. We selected three commercially available collagen sheets with different thicknesses and densities and examined the effect of rose bengal and green light collagen crosslinking (RGX) on properties such as microstructure, swelling degree, mechanical stability, cell compatibility and drug release. The highest impact of RGX was measured for Atelocollagen, for which the swelling degree was reduced from 630% (w/w) to 520% (w/w) and thickness measured under force application increased from 0.014 mm to 0.455 mm, indicating a significant increase in mechanical stability. Microstructural analysis revealed that the sponge-like structure was replaced by a fibrous structure. While the initial burst effect during vancomycin release was not influenced by crosslinking, RGX increased cell proliferation on sheets of Atelocollagen and on Collagen Solutions. We furthermore demonstrate that RGX can be used to covalently attach different sheets to create materials with combined properties, making the modification and combination of readily available sheets with RGX an attractive approach for clinical application. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Overview of methods used for collagen analysis.</p>
Full article ">Figure 2
<p>Photographs of collagen sheets and their thickness as analyzed via height gauge. Sample size: 1 × 1 cm<sup>2</sup>. (<b>C</b>) Collagen Solutions, (<b>V</b>) Viscofan and (<b>A</b>) Atelocollagen.</p>
Full article ">Figure 3
<p>Swelling behavior and microstructure of collagen sheets. (<b>A</b>) Swelling degree of collagen sheets “as received” at 37 °C and RT after 2 h, 4 h and 24 h in relation to their dry weight (100%). (<b>B</b>) transmitted light microscopy (TLM) and SEM images of collagen sheets. Samples analyzed via TLM were conditioned in phosphate-buffered saline (PBS). TLM magnification: 200×. SEM magnification: V and A: 1000×, C: 200×. C: Collagen Solutions, V: Viscofan and A: Atelocollagen.</p>
Full article ">Figure 4
<p>Impact of rose bengal and green light collagen crosslinking (RGX) on swelling degree and microstructure. (<b>A</b>) Swelling degree of unmodified and modified (RGX: 10 min, 0.01% rose bengal (RB)) collagen sheets after 2 h at 37 °C in relation to their dry weight (100%). (<b>B</b>–<b>D</b>) TLM and SEM images of collagen sheets before (unmodified) and after RGX (10 min, 0.01% RB). Samples analyzed via TLM were conditioned in PBS. (<b>B</b>) Collagen Solutions. TLM magnification: 50×. SEM magnification: 200× and 1000× (inset). (<b>C</b>) Viscofan. TLM magnification: 200×. SEM magnification: 1000×. (<b>D</b>) Atelocollagen. TLM magnification: 200×. SEM magnification: 1000×; C: Collagen Solutions, V: Viscofan and A: Atelocollagen.</p>
Full article ">Figure 5
<p>Thickness of unmodified and modified (RGX: 10 min, 0.01% RB) collagen sheets at 37 °C and RT. Samples were conditioned in PBS for 24 h before measurement. (<b>A</b>) Collagen Solutions, (<b>B</b>) Viscofan, (<b>C</b>) Atelocollagen.</p>
Full article ">Figure 6
<p>Proliferation of (<b>A</b>) human osteoblasts and (<b>B</b>) primary human muscle cells on (1) unmodified and (2) modified (RGX: 10 min, 0.01% RB) collagen sheets for 1, 7 and 10 days. The results are presented in percentage of the control without collagen sheet (100%). C: Collagen Solutions, V: Viscofan and A: Atelocollagen.</p>
Full article ">Figure 7
<p>Release of vancomycin from unmodified and modified (RGX: 10 min, 0.01% RB) collagen samples over the course of 24 h. Collagen sheets were loaded with 1000 µg vancomycin (100%). (<b>A</b>) Collagen Solutions, (<b>B</b>) Atelocollagen.</p>
Full article ">Figure 8
<p>Thickness of collagen laminates consisting of Collagen Solutions collagen (C) and Atelocollagen (A) prepared by RGX (10 min, 0.1% RB) compared to their theoretical thickness (sum of thicknesses of collagen single sheets) of unmodified and modified samples (RGX) at 37 °C and RT. Samples were conditioned in PBS for 24 h before measurement.</p>
Full article ">
15 pages, 8399 KiB  
Article
LINC00885 a Novel Oncogenic Long Non-Coding RNA Associated with Early Stage Breast Cancer Progression
by Martin C. Abba, Romina Canzoneri, Agustina Gurruchaga, Jaeho Lee, Pradeep Tatineni, Hyunsuk Kil, Ezequiel Lacunza and C. Marcelo Aldaz
Int. J. Mol. Sci. 2020, 21(19), 7407; https://doi.org/10.3390/ijms21197407 - 8 Oct 2020
Cited by 15 | Viewed by 3121
Abstract
Long intergenic non-protein coding RNA 885 (LINC00885) was identified as significantly upregulated in breast ductal carcinoma in situ (DCIS). The aim of this study was to characterize the phenotypic effects and signaling pathways modulated by LINC00885 in non-invasive and invasive breast [...] Read more.
Long intergenic non-protein coding RNA 885 (LINC00885) was identified as significantly upregulated in breast ductal carcinoma in situ (DCIS). The aim of this study was to characterize the phenotypic effects and signaling pathways modulated by LINC00885 in non-invasive and invasive breast cancer models. We determined that LINC00885 induces premalignant phenotypic changes by increasing cell proliferation, motility, migration and altering 3D growth in normal and DCIS breast cell lines. Transcriptomic studies (RNA-seq) identified the main signaling pathways modulated by LINC00885, which include bioprocesses related to TP53 signaling pathway and proliferative signatures such as activation of EREG, EGFR and FOXM1 pathways. LINC00885 silencing in breast cancer lines overexpressing this lncRNA leads to downregulation of proliferation related transcripts such as EREG, CMYC, CCND1 and to significant decrease in cell migration and motility. TCGA-BRCA data analyses show an association between high LINC00885 expression and worse overall survival in patients with primary invasive breast carcinomas (p = 0.024), suggesting that the pro-tumorigenic effects of LINC00885 overexpression persist post-invasion. We conclude that LINC00885 behaves as a positive regulator of cell growth both in normal and DCIS breast cells possibly operating as a ceRNA and representing a novel oncogenic lncRNA associated with early stage breast cancer progression. Full article
(This article belongs to the Special Issue Epithelial Cells and Cancer: Victims, Villains or Heroes?)
Show Figures

Figure 1

Figure 1
<p><span class="html-italic">LINC00885</span> expression and subcellular localization in breast cell lines. (<b>a</b>) Comparative expression profile of <span class="html-italic">LINC00885</span> among luminal-like and basal-like breast cancer cell lines based on RNA-seq data obtained from the Broad Institute Cancer Cell Line Encyclopedia (CCLE) available at UCSC Xena resource (<a href="https://xenabrowser.net/" target="_blank">https://xenabrowser.net/</a>). RPKM: reads per kilobase of transcript, per million map reads. Insert shows box and whisker plot comparing median and variability of luminal-like cell lines (dark gray) vs. basal-like cell lines (light gray), statistical significance of comparison (<span class="html-italic">p =</span> 0.022) indicated. (<b>b</b>) Fold changes in <span class="html-italic">LINC00885</span> expression levels as determined by RT-qPCR in DCIS and breast cancer cell lines compared to MCF10A normal cell line. All assays were performed in triplicate and normalized to housekeeping gene <span class="html-italic">GAPDH</span>. (<b>c</b>) Subcellular localization of <span class="html-italic">LINC00885</span> in T47D cells. <span class="html-italic">MALAT1</span> and <span class="html-italic">MT-RNR1</span> were analyzed as nuclear and cytoplasmic markers respectively. Cellular homogenates were separated into nuclear and cytoplasmic fractions and relative <span class="html-italic">LINC00885</span> expression was evaluated by RT-qPCR. (<b>d</b>) Violin plot showing the <span class="html-italic">LINC00885</span> expression distribution (medians as white dots and interquartile as black vertical lines) among normal breast tissue (<span class="html-italic">n =</span> 179) and primary invasive carcinomas (<span class="html-italic">n =</span> 1092) obtained from TCGA-TARGER-GTEx dataset available at UCSC Xena resource. Mann–Whitney–Wilcoxon test was used to compare the <span class="html-italic">LINC00885</span> expression distribution of both groups. (<b>e</b>) Kaplan–Meier survival analysis on data of 725 patients obtained from the TCGA-TARGER-GTEx dataset. Breast cancer patients with high <span class="html-italic">LINC00885</span> expression (<span class="html-italic">n =</span> 360, red line) showed reduced overall survival compared to patients with low expression (<span class="html-italic">n =</span> 365, blue line) (log-rank <span class="html-italic">p =</span> 0.024).</p>
Full article ">Figure 2
<p>Stable overexpression of <span class="html-italic">LINC00885</span> induces increased cell proliferation and colony growth in normal and DCIS breast cell lines. (<b>a</b>) Levels of <span class="html-italic">LINC00885</span> expression in stably transduced normal breast cell lines (MCF10A and 184A1) and DCIS.COM cell line. Levels of expression were determined by means of RT-qPCR in pLOC-<span class="html-italic">LINC00885</span> or pLOC-empty transfected cells (+). All assays were performed in triplicate and normalized to housekeeping gene <span class="html-italic">GAPDH</span>. (<b>b</b>) Stable overexpression of <span class="html-italic">LINC00885</span> increases cell proliferation in normal breast and DCIS cells (* <span class="html-italic">p</span> &lt; 0.05). Cells were plated at 1000 cells per well on 96 well plates in triplicate and cell proliferation was determined by means of the colorimetric MTT assay and measuring optical density (OD). (<b>c</b>) Cells stably transduced with lentivirus expressing <span class="html-italic">LINC00885</span> or vector control were plated at clonal density in 6-well plates. Cells were allowed to grow for 9 days, fixed and stained with crystal violet. (<b>d</b>) Box and whisker plots display increased colony size (left) and area occupied by colonies (right) for <span class="html-italic">LINC00885</span> stably transduced cells (MCF10 on the top and DCIS.COM on the bottom) compared with vector control. Statistical significance was determined using Mann–Whitney–Wilcoxon test (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>Increase in 3D growth and branching of normal breast cells upon stably expressing <span class="html-italic">LINC00885</span>. MCF10A after plating at clonal density increase growth effects were evident when compared with vector control at 7 and 10 days of growth in Matrigel as indicated. Scale bar = 400 µm</p>
Full article ">Figure 4
<p><span class="html-italic">LINC00885</span> overexpression increases motility and invasion in normal breast and DCIS cells. (<b>a</b>) MCF10A and DCIS.COM stably transduced cells with either vector control or lentivirus expressing <span class="html-italic">LINC00885</span> were compared using the in vitro wound-healing assay. As can be observed in representative images, 24 h. after the original scratch the area covered by migrating cells from the edges was compared in both conditions and cell lines. Scale bar = 20 µm. (<b>b</b>) Transwell migration assay of DCIS.COM cells stably transduced with <span class="html-italic">LINC00885</span>. On left comparative pictures of cells that migrated through the membrane, on the right box and whisker plot of numbers of cells per membrane (<span class="html-italic">p</span> &lt; 0.01). Statistical significance was determined using a Mann–Whitney–Wilcoxon test.</p>
Full article ">Figure 5
<p>Silencing of <span class="html-italic">LINC00885</span> expression significantly decreases cell motility in breast cancer lines. (<b>a</b>) Relative <span class="html-italic">LINC00885</span> RNA levels in MCF7 and T47D cells that were transiently silenced for <span class="html-italic">LINC00885</span> by means of siRNA (siRNA1-3) and compared with scrambled non-specific siRNA control (siNC). (<b>b</b>) A significant decrease in cell motility upon <span class="html-italic">LINC00885</span> silencing was observed in both breast cancer lines using the wound-healing assay. Forty-eight hours after the original scratch the area covered by migrating cells from the edges was compared. Scale bar = 400 µm (<b>c</b>) Graphs showing the comparative effects of negative control siRNA (blue line) and specific <span class="html-italic">LINC00885</span> siRNAs (siRNA1 in red and siRNA2 in yellow) on migration distance in wound-healing assays.</p>
Full article ">Figure 6
<p>Transcriptomic analysis of <span class="html-italic">LINC00885</span> overexpressing cells. (<b>a</b>) Hierarchical clustering of MCF10A and DCIS.COM stably transduced cells with either vector control (pLOC-empty) or lentivirus expressing <span class="html-italic">LINC00885</span> (pLOC-<span class="html-italic">LINC00885</span>) based on RNA-seq profiles. (<b>b</b>) Heat map representation of the differentially expressed genes (DEG) obtained by RNA-seq analysis (<span class="html-italic">p</span> &lt; 0.05; FC &gt; 2). Red represents upregulated genes and green down-modulated genes. (<b>c</b>) Functional enrichment of bioprocesses identified as affected by expression of <span class="html-italic">LINC00885</span> in MCF10A and DCIS.COM cells. On the left bioprocesses enriched due <span class="html-italic">LINC00885</span> overexpression in both cell lines. The red dotted line represent the basic significance level (<span class="html-italic">p</span> &lt; 0.05). At right specific genes identified by RNA-seq analysis and upregulated in MCF10A <span class="html-italic">LINC00885</span> expressing cells relative to vector control as Log2 FC. (<b>d</b>) RT-qPCR validation of cell proliferation related transcripts (<span class="html-italic">EREG</span>, <span class="html-italic">CMYC</span> and <span class="html-italic">CCND1</span>) in MCF10A <span class="html-italic">LINC00885</span> expressing cells compared to vector control (<span class="html-italic">p</span> &lt; 0.05). (<b>e</b>) Western blot analysis of EREG protein expression in MCF10A cells overexpressing <span class="html-italic">LINC00885</span> and vector control. (<b>f</b>) RT-qPCR validation of cell proliferation related transcripts in T47D (<span class="html-italic">EREG</span> and <span class="html-italic">CMYC</span>) and MCF7 cells (<span class="html-italic">CCND1</span> and <span class="html-italic">AREG</span>) transiently silenced for <span class="html-italic">LINC00885</span> and compared with negative control siRNA (<span class="html-italic">p</span> &lt; 0.01). All assays were performed in triplicate and normalized to housekeeping gene <span class="html-italic">RNA18S</span>. Statistical significance was determined using Mann–Whitney–Wilcoxon test.</p>
Full article ">
37 pages, 974 KiB  
Review
Therapeutic Potential of Endothelial Colony-Forming Cells in Ischemic Disease: Strategies to Improve their Regenerative Efficacy
by Pawan Faris, Sharon Negri, Angelica Perna, Vittorio Rosti, Germano Guerra and Francesco Moccia
Int. J. Mol. Sci. 2020, 21(19), 7406; https://doi.org/10.3390/ijms21197406 - 7 Oct 2020
Cited by 33 | Viewed by 4605
Abstract
Cardiovascular disease (CVD) comprises a range of major clinical cardiac and circulatory diseases, which produce immense health and economic burdens worldwide. Currently, vascular regenerative surgery represents the most employed therapeutic option to treat ischemic disorders, even though not all the patients are amenable [...] Read more.
Cardiovascular disease (CVD) comprises a range of major clinical cardiac and circulatory diseases, which produce immense health and economic burdens worldwide. Currently, vascular regenerative surgery represents the most employed therapeutic option to treat ischemic disorders, even though not all the patients are amenable to surgical revascularization. Therefore, more efficient therapeutic approaches are urgently required to promote neovascularization. Therapeutic angiogenesis represents an emerging strategy that aims at reconstructing the damaged vascular network by stimulating local angiogenesis and/or promoting de novo blood vessel formation according to a process known as vasculogenesis. In turn, circulating endothelial colony-forming cells (ECFCs) represent truly endothelial precursors, which display high clonogenic potential and have the documented ability to originate de novo blood vessels in vivo. Therefore, ECFCs are regarded as the most promising cellular candidate to promote therapeutic angiogenesis in patients suffering from CVD. The current briefly summarizes the available information about the origin and characterization of ECFCs and then widely illustrates the preclinical studies that assessed their regenerative efficacy in a variety of ischemic disorders, including acute myocardial infarction, peripheral artery disease, ischemic brain disease, and retinopathy. Then, we describe the most common pharmacological, genetic, and epigenetic strategies employed to enhance the vasoreparative potential of autologous ECFCs by manipulating crucial pro-angiogenic signaling pathways, e.g., extracellular-signal regulated kinase/Akt, phosphoinositide 3-kinase, and Ca2+ signaling. We conclude by discussing the possibility of targeting circulating ECFCs to rescue their dysfunctional phenotype and promote neovascularization in the presence of CVD. Full article
(This article belongs to the Special Issue Endothelial Progenitor Cells in Health and Disease)
Show Figures

Figure 1

Figure 1
<p>Manipulating dysfunctional endothelial colony-forming cells (ECFCs) to improve their regenerative potential for therapeutic angiogenesis. ECFCs isolated from the peripheral blood of individuals suffering from cardiovascular disease (CVD) present a reduced therapeutic efficacy. A number of strategies were designed to improve the regenerative potential of these ECFCs in the view of autologous cell-based therapy. These treatments include pharmacological pre-conditioning (e.g., with bioactive cues), genetic manipulation, and epigenetic activation, to improve their pro-angiogenic potential. It has been shown that ECFC manipulation remarkably improves neovascularizaiton and restores local blood flow in animal models of acute myocardial infarction (AMI), ischemic retinopathy, peripheral artery disease (PAD), and stroke.</p>
Full article ">
13 pages, 1963 KiB  
Article
Effect of Vesicle Size on the Cytolysis of Cell-Penetrating Peptides (CPPs)
by Kazutami Sakamoto, Takeshi Kitano, Haruka Kuwahara, Megumi Tedani, Kenichi Aburai, Shiroh Futaki, Masahiko Abe, Hideki Sakai, Hiroyasu Ohtaka and Yuji Yamashita
Int. J. Mol. Sci. 2020, 21(19), 7405; https://doi.org/10.3390/ijms21197405 - 7 Oct 2020
Cited by 8 | Viewed by 3309
Abstract
A specific series of peptides, called a cell-penetrating peptide (CPP), is known to be free to directly permeate through cell membranes into the cytosol (cytolysis); hence, this CPP would be a potent carrier for a drug delivery system (DDS). Previously, we proposed the [...] Read more.
A specific series of peptides, called a cell-penetrating peptide (CPP), is known to be free to directly permeate through cell membranes into the cytosol (cytolysis); hence, this CPP would be a potent carrier for a drug delivery system (DDS). Previously, we proposed the mechanism of cytolysis as a temporal and local phase transfer of membrane lipid caused by positive membrane curvature generation. Moreover, we showed how to control the CPP cytolysis. Here, we investigate the phospholipid vesicle’s size effect on CPP cytolysis because this is the most straightforward way to control membrane curvature. Contrary to our expectation, we found that the smaller the vesicle diameter (meaning a higher membrane curvature), the more cytolysis was suppressed. Such controversial findings led us to seek the reason for the unexpected results, and we ended up finding out that the mobility of membrane lipids as a liquid crystal is the key to cytolysis. As a result, we could explain the cause of cytolysis suppression by reducing the vesicle size (because of the restriction of lipid mobility); osmotic pressure reduction to enhance positive curvature generation works as long as the membrane is mobile enough to modulate the local structure. Taking all the revealed vital factors and their effects as a tool, we will further explore how to control CPP cytolysis for developing a DDS system combined with appropriate cargo selection to be tagged with CPPs. Full article
(This article belongs to the Special Issue Assembly Superstructures in Chemistry)
Show Figures

Figure 1

Figure 1
<p>Confirmation of the vesicle formation for 1-stearoyl-2-oleoyl-phosphatidylcholine (SOPC). (<b>a</b>) Freeze-fractured transmission electron microscope (FF-TEM) image of small unilamellar vesicle (SUV); (<b>b</b>) FF-TEM image of large unilamellar vesicle (LUV); (<b>c</b>) optical microscope image of giant unilamellar vesicle (GUV).</p>
Full article ">Figure 2
<p>Size distribution analysis for egg yolk phosphatidylcholine (EPC)-vesicles by dynamic light scattering (DLS).</p>
Full article ">Figure 3
<p>Effect of vesicle size with osmotic pressure change for FITC–R8 cytolysis to EPC vesicles. L/P = 1000, 37 °C, 10 min, <span class="html-italic">n</span> = 3; cytolysis amount of FITC–R8/outer membrane lipids (μmol/mol).</p>
Full article ">Figure 4
<p>Effect of osmotic pressure for FITC–R8 cytolysis vs. EPC vesicle size. L/P = 1000, 37 °C, 10 min, <span class="html-italic">n</span> = 3; cytolysis amount of FITC–R8 (CPP)/vesicle (number of CPP molecule/each vesicle).</p>
Full article ">Figure 5
<p>Influence of temperature on cytolysis of FITC–R8 to SOPC vesicle (SOPC; L/P = 1000, 10 min, <span class="html-italic">n</span> = 3).</p>
Full article ">Figure 6
<p>DSC profiles of SOPC vesicles.</p>
Full article ">Figure 7
<p>Gel to liquid crystal (LC) phase transition enthalpy (Δ<span class="html-italic">H</span>) for SOPC vesicles.</p>
Full article ">Figure 8
<p>Effect of osmotic pressure on the DSC peak profiles of SOPC vesicles. DSC peaks are recorded for a repeated freeze-and-thaw process (5–8 times). Osmotic pressures (mOsm) are 112 for hypertonic, 56 for isotonic, and 14 for hypotonic. Arrow (a) indicates the diminution of the trace of the ripple phase, and arrow (b) shows the incremental height of the phase transition peak.</p>
Full article ">
25 pages, 11077 KiB  
Article
Conserved and Opposite Transcriptome Patterns during Germination in Hordeum vulgare and Arabidopsis thaliana
by Yanqiao Zhu, Oliver Berkowitz, Jennifer Selinski, Andreas Hartmann, Reena Narsai, Yan Wang, Peisheng Mao and James Whelan
Int. J. Mol. Sci. 2020, 21(19), 7404; https://doi.org/10.3390/ijms21197404 - 7 Oct 2020
Cited by 7 | Viewed by 3260
Abstract
Seed germination is a critical process for completion of the plant life cycle and for global food production. Comparing the germination transcriptomes of barley (Hordeum vulgare) to Arabidopsis thaliana revealed the overall pattern was conserved in terms of functional gene ontology; [...] Read more.
Seed germination is a critical process for completion of the plant life cycle and for global food production. Comparing the germination transcriptomes of barley (Hordeum vulgare) to Arabidopsis thaliana revealed the overall pattern was conserved in terms of functional gene ontology; however, many oppositely responsive orthologous genes were identified. Conserved processes included a set of approximately 6000 genes that peaked early in germination and were enriched in processes associated with RNA metabolism, e.g., pentatricopeptide repeat (PPR)-containing proteins. Comparison of orthologous genes revealed more than 3000 orthogroups containing almost 4000 genes that displayed similar expression patterns including functions associated with mitochondrial tricarboxylic acid (TCA) cycle, carbohydrate and RNA/DNA metabolism, autophagy, protein modifications, and organellar function. Biochemical and proteomic analyses indicated mitochondrial biogenesis occurred early in germination, but detailed analyses revealed the timing involved in mitochondrial biogenesis may vary between species. More than 1800 orthogroups representing 2000 genes displayed opposite patterns in transcript abundance, representing functions of energy (carbohydrate) metabolism, photosynthesis, protein synthesis and degradation, and gene regulation. Differences in expression of basic-leucine zippers (bZIPs) and Apetala 2 (AP2)/ethylene-responsive element binding proteins (EREBPs) point to differences in regulatory processes at a high level, which provide opportunities to modify processes in order to enhance grain quality, germination, and storage as needed for different uses. Full article
(This article belongs to the Special Issue Plant Respiration)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Barley grain germination. (<b>A</b>) Developmental stage of barley embryos that were analyzed. Barley embryos were manually dissected from dry (0 h) or imbibed (3 to 48 h) seeds to determine dry and wet weight, to isolate mitochondria and RNA, and for use in oxygen consumption measurements. (<b>B</b>) To determine the water content, we dissected embryos from barley grain 0 to 48 h after imbibition. Wet and dry weights were measured for barley embryos and the water content was calculated from these data. For each time point, the mean standard error is shown (<span class="html-italic">n</span> = 50). Both wet and dry weight increased steadily over time up until 36 h, when the radicle had emerged and thus germination had occurred. A larger increase in both wet and dry weight was observed after 36 h. Water content initially increased up until 6 h, plateaued, and continued steadily after 12 h.</p>
Full article ">Figure 2
<p>Transcriptome of barley grain germination. Differentially expressed genes (DEGs; |log<sub>2</sub>(fold-change)|&gt;1, false discovery rate (FDR) &lt; 0.05) during the course of germination were identified by comparing to the imbibed seed (0 h) using RNAseq. (<b>A</b>) The number of expressed genes at 0 h is shown, and those that overlapped with the Arabidopsis dry seed are indicated in bold. (<b>B</b>) The number of DEGs are shown for barely and Arabidopsis as well as the heatmaps following average linkage hierarchical clustering of the fold changes. Genes that peaked at 12 h or earlier during germination are indicated in the green boxes. (<b>C</b>) The top 15 Gene Ontology (GO) biological processes that were enriched in the (i) downregulated genes, (ii) genes that peaked at 12 h or earlier, and (iii) upregulated genes are shown. Categories are in order of fold enrichment. The functional categories that overlapped with the outputs following the same analysis in Arabidopsis are indicated in bold. All GO enrichment was carried out using Fisher’s test and Bonferroni FDR correction.</p>
Full article ">Figure 3
<p>Comparison of transcript abundance profiles during germination in barley and Arabidopsis. For cross-species comparison of transcript abundance changes during germination, we compared changes in transcript abundance in barley to published data in Arabidopsis [<a href="#B10-ijms-21-07404" class="html-bibr">10</a>]. Genes in both species were classified into orthologous groups of genes (“orthogroups”) using OrthoFinder 2 [<a href="#B35-ijms-21-07404" class="html-bibr">35</a>] and clustered using the Clust algorithm [<a href="#B36-ijms-21-07404" class="html-bibr">36</a>]. A summary of the enriched GO terms for genes in each cluster is given below each panel, and the full list available in <a href="#app1-ijms-21-07404" class="html-app">Table S6</a>. The number in each panel gives the number of orthogroups in each cluster and the percentage of all orthogroups per species.</p>
Full article ">Figure 4
<p>Expression of genes in orthogroups. (<b>A</b>) The number of DEGs in barley and Arabidopsis that were orthologous and showed conserved or opposite responses. Conserved was defined as orthogroups containing genes that were only up/downregulated over the time course at one or more time points in both species, while opposite orthogroups were identified as those in which opposite contrasting responses were observed between both species. Both groups exclude any orthogroups in which genes were both up- and downregulated over the time course and in which orthologues within the same orthogroup showed contrasting opposite responses for each species. (<b>B</b>) Heatmaps showing the conserved and oppositely responsive genes in barley and Arabidopsis, (i) genes in orthogroups that are conserved with an increase in transcript abundance in Arabidopsis and barley, and (ii) conserved with a decrease in transcript abundance in Arabidopsis and barley, (iii) genes in orthogroups that displayed a decrease in transcript abundance in Arabidopsis but an increase in barley, and (iv) genes in orthogroups that displayed an increase in transcript abundance in Arabidopsis but a decrease in barley.</p>
Full article ">Figure 5
<p>Analysis of genes showing the distinct responses during seed germination in barley and Arabidopsis. (<b>A</b>) The oppositely responsive gene sets in barley and Arabidopsis were analyzed for significantly over-represented functional categories (Fisher’s test, <span class="html-italic">p</span> &lt; 0.05) using Pageman [<a href="#B37-ijms-21-07404" class="html-bibr">37</a>], and the outputs for the genes that were (i) upregulated in barley and orthologous to the (ii) downregulated genes in Arabidopsis are shown. (<b>B</b>) The significantly over-represented functional categories for the genes (i) downregulated in barley that were orthologous to the (ii) upregulated genes in Arabidopsis. (<b>C</b>) Orthologous genes encoding starch synthesis functions that showed opposing responses between barley and Arabidopsis.</p>
Full article ">Figure 6
<p>Analysis of distinct expression responses for energy- and protein-related functions during germination in barley and Arabidopsis. The over-represented functional categories in which more genes showed the opposite response to genes in the same category in Arabidopsis were identified, and the total number of DEGs that were upregulated, downregulated, and both up- and downregulated during germination are indicated for each functional category. (<b>A</b>) The energy-related functional categories in which more genes showed opposite responses. (<b>B</b>) The protein-related functional categories in which more genes showed opposite responses in barley and Arabidopsis. (<b>C</b>) Heatmap showing these responses for genes annotated as encoding cytosolic ribosome components.</p>
Full article ">Figure 7
<p>Expression of genes encoding DNA/RNA binding functions during germination. (<b>A</b>) The over-represented DNA/RNA binding-related functional categories in which more genes showed the opposite response to genes in the same category in Arabidopsis were identified, with the total number of DEGs that were upregulated, downregulated, and up- and downregulated during germination being indicated for each. (<b>B</b>) Heatmaps showing the expression of all genes encoding (i) chromatin organization components and (ii) basic leucine zipper (bZIP) TFs in Arabidopsis and barley. (<b>C</b>) Of the 1020 TFs identified, significantly enriched TF families in the up/downregulated genes as well as the genes that were both up- and downregulated during germination are shown (* <span class="html-italic">p</span> &lt; 0.05). (<b>D</b>) The 22 genes encoding Apetala2 (AP2)/ethylene-responsive element binding proteins (EREBP) that were both up- and downregulated compared to 0 h during germination. Of these, 12 genes (indicated in bold) were identified from barely gene annotation only and not orthology.</p>
Full article ">Figure 8
<p>Mitochondrial dynamics during barley grain germination. (<b>A</b>) Oxygen consumption in whole embryos and isolated mitochondria. Oxygen consumption was examined in whole barley embryos and isolated mitochondria using a Clark-type oxygen electrode. Each data point represents an average of three to four independent measurements. (<b>B</b>) Changes in abundance of mitochondrial proteins. A representative image is shown from probing three independent isolated mitochondrial proteins using both 10 and 20 µg of isolated protein. The abundance of the protein is normalized to a value of 1 at 48 h and all other values are expressed in a relative manner. For NDUFS4 “a” and “b”, all values are relative to 12 h and 0 h, respectively. For B14.7 “a”, “b”, and “c”, all values are relative to 0 h, 0 h, and 24 h, respectively. The apparent molecular mass is indicated in kDa. NDUFS4 = 17 kDa subunit complex I, B14.7 = complex I subunit, SDH = subunit I of succinate dehydrogenase/complex II, RISP = Rieske FeS protein of complex III, COXII = cytochrome oxidase II of complex IV, ATPβ = β subunit of ATP synthase, Cyt c = cytochrome c, AOX = alternative oxidase, Tom20-3 = translocase of the outer membrane 20-3, Tom40 = translocase of the outer membrane 40, Porin = voltage-dependent anion channel, UCP = uncoupling protein, LETM1 = leucine zipper-EF-hand-containing transmembrane protein 1, Tim44 = translocase of the outer membrane 44. (<b>C</b>) The heatmap represents the abundance of 173 mitochondrial proteins at the indicated time points after seed inhibition. Quantification was performed by proteomic analysis (label-free quantification) of purified mitochondrial proteins, and abundances are represented by z-scored intensity-based absolute quantification (iBAQ) values. The position of several protein isoforms is indicated in the heatmap (see <a href="#app1-ijms-21-07404" class="html-app">Table S10</a> for details): CYC B/C, CYTOCHROM b/c subunit; ND, NADH DEHYDROGENASE; SDH, SUCCINATE DEHYDROGENASE; NDA/NDB, ALTERNATIVE NAD(P)H DEHYDROGENASE; COX, CYTOCHROME OXIDASE; LETM1, LEUCINE ZIPPER-EF-HAND-CONTAINING TRANSMEMBRANE PROTEIN 1.</p>
Full article ">Figure 9
<p>Expression of genes encoding mitochondrial proteins during germination. (<b>A</b>) Hierarchical clustering of the fold changes (logFC) compared to 0 h for the 1208 genes encoding mitochondrial proteins in barley and 1226 genes encoding mitochondrial proteins in Arabidopsis [<a href="#B10-ijms-21-07404" class="html-bibr">10</a>]. (<b>B</b>) GO over-representation analysis was carried out for the three clusters of responses in barley, and the top 15 significantly over-represented categories are shown, (i) for genes that decrease in transcript abundance during germination, (ii) for genes that display a transient increase in transcript abundance, and (iii) for genes that increase in transcript abundance during germination (Bonferronni; <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 10
<p>Analysis of genes encoding orthologous mitochondrial proteins that are responsive during germination in barley and Arabidopsis. (<b>A</b>) Heatmap showing the orthologous genes that showed conserved responses to germination in barley and Arabidopsis, with (i) conserved upregulated genes and (ii) conserved downregulated gene sets shown. The top 15 gene ontology-enriched categories are shown next to these (<span class="html-italic">p</span> &lt; 0.05, Bonferroni). (<b>B</b>) Heatmaps showing orthologous genes that showed distinct responses to germination in barley and Arabidopsis, with (i) upregulated genes in Arabidopsis orthologous to downregulated genes in barley shown, and (ii) downregulated genes in Arabidopsis orthologous to upregulated genes in barley shown. The top 15 gene ontology (GO)-enriched categories are shown next to these (<span class="html-italic">p</span> &lt; 0.05, Bonferroni).</p>
Full article ">
12 pages, 4065 KiB  
Article
N-Butylidenephthalide Inhibits the Phenotypic Switch of VSMCs through Activation of AMPK and Prevents Stenosis in an Arteriovenous Fistula Rat Model
by Hsin-Han Yang, Yue-Xuan Xu, Jie-Yi Chen, Horng-Jyh Harn and Tzyy-Wen Chiou
Int. J. Mol. Sci. 2020, 21(19), 7403; https://doi.org/10.3390/ijms21197403 - 7 Oct 2020
Cited by 6 | Viewed by 2588
Abstract
The phenotypic switch of vascular smooth muscle cells (VSMCs) plays a pivotal role in the development of vascular disorders, such as atherosclerosis, stenosis and restenosis, after vascular intervention. In our previous study, n-butylidenephthalide (BP) was reported to have anti-proliferating and apoptotic effects on [...] Read more.
The phenotypic switch of vascular smooth muscle cells (VSMCs) plays a pivotal role in the development of vascular disorders, such as atherosclerosis, stenosis and restenosis, after vascular intervention. In our previous study, n-butylidenephthalide (BP) was reported to have anti-proliferating and apoptotic effects on VSMCs. The purpose of the current study is to further investigate its role in platelet-derived growth factor (PDGF)-induced VSMC phenotypic modulation in an arteriovenous fistula model. In vitro, we observed that BP inhibited the PDGF-induced cytoskeleton reorganization of the VSMCs. The enhanced expression of vimentin and collagen, as well as the migration ability induced by PDGF, were also inhibited by BP. By cell cycle analysis, we found that BP inhibited the PDGF-induced VSMCs proliferation and arrested the VSMCs in the G0/G1 phase. In an arteriovenous fistula rat model, the formation of stenosis, which was coupled with a thrombus, and the expression of vimentin and collagen in VSMCs, were also inhibited by administration of BP, indicating that BP inhibited the PDGF-induced phenotypic switch and the migration of VSMCs. Besides, the inhibitory effects of BP on the phenotypic switch were found to accompany the activated 5’ AMP-activated protein kinase (AMPK) as well as the inhibited phosphorylation of mTOR. Knockdown of AMPK by gene silencing conflicted the effects of BP and further exacerbated the PDGF-induced VSMCs phenotypic switch, confirming the modulating effect that BP exerted on the VSMCs by this pathway. These findings suggest that BP may contribute to the vasculoprotective potential in vasculature. Full article
(This article belongs to the Section Biochemistry)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>n-Butylidenephthalide (BP) inhibited PDGF-induced phenotypic switch. Measurements of morphology and protein expression in A7r5 treated with BP: (<b>A</b>) Representative images of PDGF-induced A7r5 in the absence or presence of BP (25 and 75 μg/mL) for 24 h. Scale bar = 50 μm; (<b>B</b>,<b>C</b>) Western blotting analysis of phenotypic marker expression. All data are presented as the means ± standard deviation of three independent experiments. * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 versus the PDGF group.</p>
Full article ">Figure 2
<p>BP inhibits migration and arrests the PDGF-induced A7r5 in the G0/G1 phase in a dose-dependent manner. (<b>A</b>) A wound-healing assay (<b>upper</b> panel, scale bar = 500 μm) and Oris system migrating assay (<b>lower</b> panel, scale bar = 500 μm) were performed since A7r5 was treated with PDGF (10 ng/mL) and BP (0, 25, 50, 75 and 100 µg/mL) for 24 h; (<b>B</b>) Cell cycle analysis through PI staining and following flow cytometry for the A7r5 after 24 h of PDGF (10 ng/mL) and BP treatment (0, 25 and 75 ug/mL); (<b>C</b>,<b>D</b>) Quantification of the Oris system migrating assay and cell cycle analysis. Data are expressed as the mean ± SEM. * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 3
<p>Histological staining of a venous limb of arteriovenous fistula (AVF) 30 days after establishment, and the comparison of lumen size among rats treated with different doses of BP. The thrombus and neointimal hyperplasia were indicated by a red arrow. (<b>A</b>) WT; (<b>B</b>) Sham; (<b>C</b>) 20 mg/kg-QD BP; (<b>D</b>) 50 mg/kg-QD BP; and (<b>E</b>) 50 mg/kg-BID BP. A significant reduction in thrombus and neointimal hyperplasia was observed in rats treated with BP. (<b>F</b>) Measurement of lumen size as reflected by the perimeter of the venous limb. Scale bar = 100 μm. Data are expressed as the mean ± SEM. <span class="html-italic">n</span> = 4 per group. * <span class="html-italic">p</span> &lt; 0.05 versus sham.</p>
Full article ">Figure 4
<p>Immunohistochemical detection (brown color) of a venous limb of AVF from rats treated with different doses of BP for 30 days. The expression of a phenotypic-specific marker (SMA, vimentin and collagen I) in an adjacent section of each group is indicated by a red arrow. Significant inhibition of phenotypic switch was observed in rats treated with BP compared to the control (sham) group. Scale bar = 100 μm.</p>
Full article ">Figure 5
<p>BP inhibited the PDGF-induced phenotypic switch by activating AMPK. (<b>A</b>) Protein expression of AMPK and mTOR after treatment of PDGF and BP for 24 h; (<b>B</b>) Protein expression of AMPK, mTOR and phenotypic markers in PDGF-induced A7r5 treated without or with BP in the absence or presence of siAMPK; (<b>C</b>) Representative images of PDGF-induced A7r5 treated without or with BP in the absence or presence of siAMPK. Scale bar = 50 μm</p>
Full article ">
17 pages, 1946 KiB  
Article
Antioxidant Amelioration of Riboflavin Transporter Deficiency in Motoneurons Derived from Patient-Specific Induced Pluripotent Stem Cells
by Chiara Marioli, Valentina Magliocca, Stefania Petrini, Alessia Niceforo, Rossella Borghi, Sara Petrillo, Piergiorgio La Rosa, Fiorella Colasuonno, Tiziana Persichini, Fiorella Piemonte, Keith Massey, Marco Tartaglia, Sandra Moreno, Enrico Bertini and Claudia Compagnucci
Int. J. Mol. Sci. 2020, 21(19), 7402; https://doi.org/10.3390/ijms21197402 - 7 Oct 2020
Cited by 13 | Viewed by 4771
Abstract
Mitochondrial dysfunction is a key element in the pathogenesis of neurodegenerative disorders, such as riboflavin transporter deficiency (RTD). This is a rare, childhood-onset disease characterized by motoneuron degeneration and caused by mutations in SLC52A2 and SLC52A3, encoding riboflavin (RF) transporters (RFVT2 and [...] Read more.
Mitochondrial dysfunction is a key element in the pathogenesis of neurodegenerative disorders, such as riboflavin transporter deficiency (RTD). This is a rare, childhood-onset disease characterized by motoneuron degeneration and caused by mutations in SLC52A2 and SLC52A3, encoding riboflavin (RF) transporters (RFVT2 and RFVT3, respectively), resulting in muscle weakness, ponto-bulbar paralysis and sensorineural deafness. Based on previous findings, which document the contribution of oxidative stress in RTD pathogenesis, we tested possible beneficial effects of several antioxidants (Vitamin C, Idebenone, Coenzyme Q10 and EPI-743, either alone or in combination with RF) on the morphology and function of neurons derived from induced pluripotent stem cells (iPSCs) from two RTD patients. To identify possible improvement of the neuronal morphotype, neurite length was measured by confocal microscopy after β-III tubulin immunofluorescent staining. Neuronal function was evaluated by determining superoxide anion generation by MitoSOX assay and intracellular calcium (Ca2+) levels, using the Fluo-4 probe. Among the antioxidants tested, EPI-743 restored the redox status, improved neurite length and ameliorated intracellular calcium influx into RTD motoneurons. In conclusion, we suggest that antioxidant supplementation may have a role in RTD treatment. Full article
(This article belongs to the Special Issue hiPSC-Derived Cells as Models for Drug Discovery)
Show Figures

Figure 1

Figure 1
<p>Quantification of superoxide anion in riboflavin transporter deficiency (RTD) induced pluripotent stem cells (iPSCs) following treatment with antioxidants ascorbic acid (AA), CoQ<sub>10</sub>, Idebenone, and EPI-743, showing their effect on RTD iPSCs superoxide anion production. For each RTD cell line, P1 (<b>A</b>) and P2 (<b>B</b>), the arrows indicate the antioxidant concentration most effective in lowering the levels of superoxide anion. Experiments were conducted in triplicate and values expressed as mean ± standard error of the mean (SEM). According to Kruskal-Wallis tests * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001, compared with controls’ group (Ctrl); # <span class="html-italic">p</span> &lt; 0.05, ## <span class="html-italic">p</span> &lt; 0.01 respect to untreated patients.</p>
Full article ">Figure 2
<p>RTD iPSCs show increased lipid peroxidation. (<b>A</b>) Fluorescence micrographs of iPSCs labeled with BODIPY using the green (on the left, indicating oxidation of the butadienyl portion of the dye) and the red (on the middle) filter and then overlay of the red (nonoxidized) and green (oxidized) images (in the right column). Colocalization of oxidized and reduced BODIPY fluorescence appears in yellow. Bar = 100 μm. Control and RTD iPSCs were incubated for 45 min with BODIPY 581/591 C11. Treatment with EPI-743, but not other antioxidants, results in significantly reduced levels of oxidized lipids as shown by the shift of green to red fluorescent signal. (<b>B</b>) Bar graph reporting the quantitative analyses of the BODIPY experiments performed on control and RTD iPSCs. Values are expressed as mean ± SEM. According to Kruskal-Wallis test * <span class="html-italic">p</span> &lt; 0.05 compared with controls iPSCs; # <span class="html-italic">p</span> &lt; 0.05, ## <span class="html-italic">p</span> &lt; 0.01, respect to untreated patients.</p>
Full article ">Figure 3
<p>Analysis of neurite length following antioxidant treatment of MNs derived from RTD patient-derived iPSCs. Immunofluorescence images of β III tubulin (in red) show shorter neurites in RTD MNs compared to control cells. In both P1 and P2 RTD MNs, treatment with AA, CoQ10 and IDEB fails to cause significant changes. RF and EPI-743 ± RF causes improvement in neurite length for RTD MNs. Nuclei are counterstained with Hoechst (in blue). Scale bars = 50 μm. Data derived from four independent experiments, and values are expressed as mean ± SEM. According to Kruskal-Wallis test ** <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, compared with control group (Ctrl); # <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 with respect to untreated RTD patient MNs.</p>
Full article ">Figure 4
<p>Intracellular Ca<sup>2+</sup> flux in RTD motoneurons following EPI-743 treatment. (<b>A</b>) Confocal images showing changes in intracellular calcium flux in RTD and control MNs before (basal level) and after stimulation with 5 μM ionomycin. (<b>B</b>) Graphical representation of the mean fluorescence intensity over time of Control and RTD MNs following RF or EPI-743 treatment, showing an increase in intracellular Ca<sup>2+</sup> following ionomycin supplementation (indicated by the black arrow) and decrease following addition of EGTA to the medium (30 s following ionomycin in all samples, as indicated by the grey arrow). Experiments were conducted in triplicate.</p>
Full article ">Figure 5
<p>Schemae showing the morphological changes of the RTD MNs before and following treatment with EPI-743. The drawing depicts neurons with short neurites in RTD P1 and P2, but following EPI-743 treatment they extend longer neurites that, for RTD P2, are very similar to Ctrl MNs, while, for RTD P1 neurite length is improved but still not comparable to that of Ctrl MNs. N = Nucleus. n = neurite.</p>
Full article ">
Previous Issue
Back to TopTop