Nothing Special   »   [go: up one dir, main page]

Next Issue
Volume 8, October
Previous Issue
Volume 8, August
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 

Cells, Volume 8, Issue 9 (September 2019) – 170 articles

Cover Story (view full-size image): Intermediate filaments (also known as nanofilaments) of astrocytes are dynamic structures involved in cell signaling and migration and determining cell responses in health, disease, and regeneration. Upregulation of the intermediate filament proteins GFAP, vimentin (VIM), and nestin is a hallmark of reactive gliosis, which is protective in ischemic stroke or neurotrauma, but inhibits some regenerative responses. Immature astrocytes from mice with VIM mutations of serine sites phosphorylated during mitosis (VIMSA/SA) exhibit cytokinetic failure and contain VIM accumulations that colocalize with mitochondria (green: VIM-containing intermediate filaments and VIM accumulations; red: mitochondria). This phenotype is transient and disappears with VIMSA/SA astrocyte maturation and can be alleviated by the inhibition of cell proliferation. 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:
21 pages, 3551 KiB  
Article
Proteomic Analysis of miR-195 and miR-497 Replacement Reveals Potential Candidates that Increase Sensitivity to Oxaliplatin in MSI/P53wt Colorectal Cancer Cells
by Dennis Poel, Lenka N.C. Boyd, Robin Beekhof, Tim Schelfhorst, Thang V. Pham, Sander R. Piersma, Jaco C. Knol, Connie R. Jimenez, Henk M.W. Verheul and Tineke E. Buffart
Cells 2019, 8(9), 1111; https://doi.org/10.3390/cells8091111 - 19 Sep 2019
Cited by 25 | Viewed by 4392
Abstract
Most patients with advanced colorectal cancer (CRC) eventually develop resistance to systemic combination therapy. miR-195-5p and miR-497-5p are downregulated in CRC tissues and associated with drug resistance. Sensitization to 5-FU, oxaliplatin, and irinotecan by transfection with miR-195-5p and miR-497-5p mimics was studied using [...] Read more.
Most patients with advanced colorectal cancer (CRC) eventually develop resistance to systemic combination therapy. miR-195-5p and miR-497-5p are downregulated in CRC tissues and associated with drug resistance. Sensitization to 5-FU, oxaliplatin, and irinotecan by transfection with miR-195-5p and miR-497-5p mimics was studied using cell viability and clonogenic assays in cell lines HCT116, RKO, DLD-1, and SW480. In addition, proteomic analysis of transfected cells was implemented to identify potential targets. Significantly altered proteins were subjected to STRING (protein-protein interaction networks) database analysis to study the potential mechanisms of drug resistance. Cell viability analysis of transfected cells revealed increased sensitivity to oxaliplatin in microsatellite instable (MSI)/P53 wild-type HCT116 and RKO cells. HCT116 transfected cells formed significantly fewer colonies when treated with oxaliplatin. In sensitized cells, proteomic analysis showed 158 and 202 proteins with significantly altered expression after transfection with miR-195-5p and miR-497-5p mimics respectively, of which CHUK and LUZP1 proved to be coinciding downregulated proteins. Resistance mechanisms of these proteins may be associated with nuclear factor kappa-B signaling and G1 cell-cycle arrest. In conclusion, miR-195-5p and miR-497-5p replacement enhanced sensitivity to oxaliplatin in treatment naïve MSI/P53 wild-type CRC cells. Proteomic analysis revealed potential miRNA targets associated with the cell-cycle which possibly bare a relation with chemotherapy sensitivity. Full article
Show Figures

Figure 1

Figure 1
<p>miRNA expression and transfection efficiency in CRC cells. Successful transfection of cel-miR-39-3p in the negative control transfected cells for all four different CRC cell lines (<b>A</b>). <span class="html-italic">y</span>-axis represents raw Cq value of cel-miR-39-3p. Successful transfection of miR-195-5p and miR-497-5p in the four different cell lines HCT116 (<b>B</b>), RKO (<b>C</b>), DLD1 (<b>D</b>) and SW480 (<b>E</b>). <span class="html-italic">y</span>-axis represents log2 expression of miR-195-5p and miR-497-5p relative to miR-16-5p (control). Data is presented as the mean of three independent transfection experiments (measured in duplicate) ± the standard error of the mean. *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001, wt: wild-type, cel: cel-miR-39-3p negative control transfection, mimic: expression level 48 h after transfection.</p>
Full article ">Figure 2
<p>HCT116 and RKO cells after transfection with miR-195-5p or miR-497-5p mimic. MTT sensitivity assays for HCT116 cells (<b>A</b>) and RKO cells (<b>D</b>) were performed in triplicate in three independent experiments and are presented as average % proliferation compared to the proliferation of a non-treated control triplicate ± the standard error of the mean (SEM). Presentation of a single clonogenic assay experiment of HCT116 (<b>B</b>) and RKO (<b>E</b>). Bar graphs of the percentage of formed colonies compared to the negative control transfection (cel-miR-39-3p) of HCT116 cells (<b>C</b>) and RKO cells (<b>F</b>) presented as averages of duplicate colony counts from three independent experiments ± SEM. 5-FU; 5-fluorouracil, OHP; oxaliplatin, SN-38; irinotecan, nt; no treatment, ** <span class="html-italic">p</span> &lt; 0.005, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 3
<p>miR-195-5p and miR-497-5p target selection and quantification. (<b>A</b>) Venn diagram of the targets selected with miRTarBase, DIANA LAB, and miRDB with the coinciding target mRNAs of the three databases presented in the center. (<b>B</b>) Mature miRNA sequence of miR-195-5p and miR-497-5p matched to the 3′UTR target region of the selected targets with the seed sequence of each miRNA shown in red. Expression levels, presented as log<sub>2</sub> relative to GAPDH of the specific mRNA targets CCNE1, E2F3, and WEE1, measured with RT-qPCR in HCT116 cells (<b>C</b>), RKO cells (<b>D</b>) and DLD-1 cells (<b>E</b>). Each target for each transfection is quantified in duplicate in three independent experiments, presented as average ± the SEM. wt: wild-type non-transfected control, neg: negative control transfection (cel-miR-39-3p). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 4
<p>Number of significantly differential expressed proteins after transfection with miR-195-5p and miR-497-5p miRNA mimics. Venn diagram of the downregulated proteins in the four cell lines after transfection with mimics of miR-195-5p (<b>A</b>) and miR-497-5p (<b>B</b>). Venn diagram of the upregulated proteins in the four cell lines after transfection with mimics of miR-195-5p (<b>C</b>) and miR-497-5p (<b>D</b>). Venn diagram of overlapping downregulated proteins in MSI CRC cells after transfection with mimics of miR-195-5p (<b>E</b>) and miR-497-5p (<b>F</b>). Coinciding significantly differential expressed proteins in MSI/P53 wt CRC cells transfected with miR-195-5p mimic and miR-497-5p mimic are listed in (<b>E</b>) and (<b>F</b>) respectively. In red two proteins downregulated in all four conditions.</p>
Full article ">Figure 5
<p>miRNA target evidence of differential expressed proteins from the proteomics screen. Venn diagram of mRNA target evidence found in the different databases for miR-195-5p (<b>A</b>) and miR-497-5p (<b>C</b>). mRNA target sites of potential targets of miR-195-5p (<b>B</b>) and miR-497-5p (<b>D</b>) based on the seed sequence (given in red).</p>
Full article ">
28 pages, 1031 KiB  
Review
FOXO3a from the Nucleus to the Mitochondria: A Round Trip in Cellular Stress Response
by Candida Fasano, Vittoria Disciglio, Stefania Bertora, Martina Lepore Signorile and Cristiano Simone
Cells 2019, 8(9), 1110; https://doi.org/10.3390/cells8091110 - 19 Sep 2019
Cited by 137 | Viewed by 16163
Abstract
Cellular stress response is a universal mechanism that ensures the survival or negative selection of cells in challenging conditions. The transcription factor Forkhead box protein O3 (FOXO3a) is a core regulator of cellular homeostasis, stress response, and longevity since it can modulate a [...] Read more.
Cellular stress response is a universal mechanism that ensures the survival or negative selection of cells in challenging conditions. The transcription factor Forkhead box protein O3 (FOXO3a) is a core regulator of cellular homeostasis, stress response, and longevity since it can modulate a variety of stress responses upon nutrient shortage, oxidative stress, hypoxia, heat shock, and DNA damage. FOXO3a activity is regulated by post-translational modifications that drive its shuttling between different cellular compartments, thereby determining its inactivation (cytoplasm) or activation (nucleus and mitochondria). Depending on the stress stimulus and subcellular context, activated FOXO3a can induce specific sets of nuclear genes, including cell cycle inhibitors, pro-apoptotic genes, reactive oxygen species (ROS) scavengers, autophagy effectors, gluconeogenic enzymes, and others. On the other hand, upon glucose restriction, 5′-AMP-activated protein kinase (AMPK) and mitogen activated protein kinase kinase (MEK)/extracellular signal-regulated kinase (ERK) -dependent FOXO3a mitochondrial translocation allows the transcription of oxidative phosphorylation (OXPHOS) genes, restoring cellular ATP levels, while in cancer cells, mitochondrial FOXO3a mediates survival upon genotoxic stress induced by chemotherapy. Interestingly, these target genes and their related pathways are diverse and sometimes antagonistic, suggesting that FOXO3a is an adaptable player in the dynamic homeostasis of normal and stressed cells. In this review, we describe the multiple roles of FOXO3a in cellular stress response, with a focus on both its nuclear and mitochondrial functions. Full article
(This article belongs to the Special Issue Molecular and Cellular Mechanisms of Stress Responses)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Post-translational regulation of FOXO3a. (<b>A</b>) Schematic representation of FOXO3a domains (MPP-MIP motifs, consensus motifs for mitochondrial processing peptidase (MPP) and mitochondrial intermediate peptidase (MIP); forkhead domain, FH; nuclear localization signal, NLS, Kinase-inducible domain interacting domain binding domain, KIX; transactivation domain TAD). (<b>B</b>) Summary of FOXO3a post-translational modifications (PTMs). Depicted are the most important upstream signals (AKT8 virus oncogene cellular homolog, AKT; serum and glucocorticoid-induced kinase, SGK;5′-AMP-activated protein kinase AMPK; c-Jun N-terminal kinase, JNK; extracellular signal-regulated kinase, ERK, SET domain protein 9, SET9) regulating FOXO3a subcellular localization and activity through reversible PTMs, which include phosphorylation (blue circle) and methylation (green circle) on specific amino acid residues (T, threonine; S, serine; K, methionine).</p>
Full article ">Figure 2
<p>Schematic representation of Forkhead box O3 (FOXO3a)-mediated stress response. Perturbations of cellular homeostasis, such as nutrient shortage, high concentration of intracellular ROS, or genotoxic stress, activate FOXO3a upstream stress sensors (purple boxes), which in turn modulate FOXO3a subcellular localization and/or activity through various post-translational modifications (PTMs) (green arrows represent activation signals, red bar-headed lines represent inhibitory effects). In the cytoplasm, FOXO3a is inactive and is targeted for poly-ubiquitination, which leads to its further proteasomal degradation. Upon phosphorylation by c-Jun N-terminal kinase (JNK), FOXO3a is shuttled into the nucleus, where its transcriptional activity is further regulated by an activator (e.g., 5′-AMP-activated protein kinase, AMPK; sirtuin 1, SIRT1) or repressor (e.g., AKT8 virus oncogene cellular homolog AKT, CREB binding protein and p300 (CBP/p300) signals. Depending on the PTM pattern, FOXO3a orchestrates different transcriptional programs involved in several cellular processes, including apoptosis, cell cycle progression, DNA repair, reactive oxygen species (ROS) detoxification, and cellular metabolism. Recent evidence showed that metabolic stress or chemotherapy treatment can also promote AMPK- and extracellular signal-regulated kinase (ERK) dependent mitochondrial accumulation of a FOXO3a cleaved form, which activates the expression of mitochondrial oxidative phosphorylation (OXPHOS) genes involved in cell survival. The crosstalk between FOXO3a nuclear and mitochondrial functions is crucial for the restoration and maintenance of cellular homeostasis.</p>
Full article ">
13 pages, 6923 KiB  
Article
Dynamic Interplay between Pericytes and Endothelial Cells during Sprouting Angiogenesis
by Giulia Chiaverina, Laura di Blasio, Valentina Monica, Massimo Accardo, Miriam Palmiero, Barbara Peracino, Marianela Vara-Messler, Alberto Puliafito and Luca Primo
Cells 2019, 8(9), 1109; https://doi.org/10.3390/cells8091109 - 19 Sep 2019
Cited by 47 | Viewed by 5789
Abstract
Vascular physiology relies on the concerted dynamics of several cell types, including pericytes, endothelial, and vascular smooth muscle cells. The interactions between such cell types are inherently dynamic and are not easily described with static, fixed, experimental approaches. Pericytes are mural cells that [...] Read more.
Vascular physiology relies on the concerted dynamics of several cell types, including pericytes, endothelial, and vascular smooth muscle cells. The interactions between such cell types are inherently dynamic and are not easily described with static, fixed, experimental approaches. Pericytes are mural cells that support vascular development, remodeling, and homeostasis, and are involved in a number of pathological situations including cancer. The dynamic interplay between pericytes and endothelial cells is at the basis of vascular physiology and few experimental tools exist to properly describe and study it. Here we employ a previously developed ex vivo murine aortic explant to study the formation of new blood capillary-like structures close to physiological situation. We develop several mouse models to culture, identify, characterize, and follow simultaneously single endothelial cells and pericytes during angiogenesis. We employ microscopy and image analysis to dissect the interactions between cell types and the process of cellular recruitment on the newly forming vessel. We find that pericytes are recruited on the developing sprout by proliferation, migrate independently from endothelial cells, and can proliferate on the growing capillary. Our results help elucidating several relevant mechanisms of interactions between endothelial cells and pericytes. Full article
(This article belongs to the Special Issue Angiogenesis in Cancer)
Show Figures

Figure 1

Figure 1
<p>Characterization of NG2-dsRed mAR. (<b>A</b>) Representative bright-field image of mouse aortic ring cultured for one week, showing microvessel outgrowth (scale bar represents 200 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m). (<b>B</b>) Whole-mount immunofluorescence staining of a mAR fixed after 6 days of culture. The mAR was stained for the endothelial marker VE-cadherin (green) and the pericyte marker NG2 (magenta) (Scale bar: 50 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m). (<b>C</b>) mARs obtained from NG2-dsRed mouse model were cultured for 6 days then fixed to perform a whole-mount immunofluorescence staining. mARs were stained for classical pericyte markers, like PDGFR<math display="inline"><semantics> <mi>β</mi> </semantics></math> (magenta, top row) and <math display="inline"><semantics> <mi>α</mi> </semantics></math>SMA (magenta, middle row), to verify the colocalization with dsRed signal (red). mARs were also stained for laminin (magenta, bottom row), one of the main components of vBM. Phalloidin staining was performed to identify microvessel outgrowths (green) (Scale bar: 50 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m). (<b>D</b>) Snapshots of time-lapse microscopy of angiogenic outgrowths from NG2-dsRed mAR.</p>
Full article ">Figure 2
<p>LifeAct-EGFP or H2B-EGFP NG2-dsRed mAR models for pericyte–EC interactions. (<b>A</b>) Time-lapse microscopy was performed on LifeAct-EGFP NG2-dsRed mARs cultured for 5 days then imaged for 72 h. This mouse model allows to clearly identify pericytes thanks to dsRed signal (red) running over the endothelial layer labeled by LifeAct-EGFP (green) (scale bar: 50 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m). (<b>B</b>) Time-lapse microscopy of H2B-EGFP NG2-dsRed mARs cultured for 5 days then imaged for 72 h. Thanks to this model we could identify pericytes (red) as well as track all individual cells thanks to histone H2B-EGFP (green) (scale bar: 50 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m).</p>
Full article ">Figure 3
<p>Pericytes recruited on newly formed sprouting capillaries originate from the aortic ring but can proliferate. (<b>A</b>) Time-lapse microscopy analysis of H2B-EGFP NG2-dsRed mARs cultured for 5 days then imaged for 72 h. The observed pericyte originates from the edge of the mAR and moves along the ECs during sprouting process (scale bar: 50 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m). (<b>B</b>) Cell trajectory analysis shows that endothelial cells and pericytes move along the sprout. This set of trajectories is used for the following two panels. (<b>C</b>) The distance of each cell from the ring plotted as a function of time shows that endothelial cells (plotted in different shades of green) move at a rather constant speed and, in some cases, suddenly change direction, as witnessed by the change in slope of some trajectories. (<b>D</b>) The velocity of each cell shows that pericytes and endothelial cells do not proceed coherently, but conversely they move independently of each other. In particular cells proceed with bursts (accelerations and decelerations) (<b>E</b>) Whole-mount immunofluorescence staining of a NG2-dsRed mA-sheet embedding in a collagen gel and fixed. mA-sheet was stained for the endothelial marker CD31 (green) and DAPI (blue). The rightmost panel shows an enlarged detail. (scale bar: 50 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m). (<b>F</b>) mARs were cultured in standard medium supplemented with EdU for 1 week after explant in order to detect whether pericytes and/or ECs coming from the ring originate from a proliferative event or not. Images show a representative experiment. Arrows indicate pericytes Edu-positive (white) or -negative (blue). (Scale bar: 50 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m.)</p>
Full article ">Figure 4
<p>H2B-EGFP NG2-dsRed mAR mouse models to study pericytes dynamics during the sprouting process. (<b>A</b>) Time-lapse widefield microscopy analysis of H2B-EGFP NG2-dsRed mARs. White arrow indicate a pericyte that undergoes cell division after 5 h from the beginning of the experiment, generating the two daughter cells marked with blue arrows. (Scale bar: 50 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m.) (<b>B</b>) For each cell in a sprout, we verified if it appeared on the sprout coming out of the ring (recruited) or if it divided on the sprout (divided). Each dot represents a single sprout, with green dots for endothelial cells and red dots for pericytes. (<b>C</b>) To measure the recruitment rate of cells on the sprouts, we measured the number of cells at the end of each sprouting assay (each dot represents a sprout) and measured the number of cells during the time of the experiment. (<b>D</b>) Each line represents a division event of a pericyte. Red and blue trajectories are projections of the trajectories of the daughter cells on the direction of the sprout, where the origin has been defined as the coordinate of the mother cells at the frame preceding mitosis. (<b>E</b>) Plot of the distance between daughter cells as a function of time.</p>
Full article ">
22 pages, 1839 KiB  
Review
Environmentally-Induced Transgenerational Epigenetic Inheritance: Implication of PIWI Interacting RNAs
by Karine Casier, Antoine Boivin, Clément Carré and Laure Teysset
Cells 2019, 8(9), 1108; https://doi.org/10.3390/cells8091108 - 19 Sep 2019
Cited by 20 | Viewed by 7353
Abstract
Environmentally-induced transgenerational epigenetic inheritance is an emerging field. The understanding of associated epigenetic mechanisms is currently in progress with open questions still remaining. In this review, we present an overview of the knowledge of environmentally-induced transgenerational inheritance and associated epigenetic mechanisms, mainly in [...] Read more.
Environmentally-induced transgenerational epigenetic inheritance is an emerging field. The understanding of associated epigenetic mechanisms is currently in progress with open questions still remaining. In this review, we present an overview of the knowledge of environmentally-induced transgenerational inheritance and associated epigenetic mechanisms, mainly in animals. The second part focuses on the role of PIWI-interacting RNAs (piRNAs), a class of small RNAs involved in the maintenance of the germline genome, in epigenetic memory to put into perspective cases of environmentally-induced transgenerational inheritance involving piRNA production. Finally, the last part addresses how genomes are facing production of new piRNAs, and from a broader perspective, how this process might have consequences on evolution and on sporadic disease development. Full article
(This article belongs to the Special Issue Evolution of Epigenetic Mechanisms and Signatures)
Show Figures

Figure 1

Figure 1
<p>Germline PIWI-interacting RNAs (piRNA) biogenesis. In <span class="html-italic">Drosophila melanogaster</span> germ cells, active piRNA clusters are enriched in H3K9me3 and HP1-like protein Rhino (R), which interact with Deadlock (Del), Cutoff (Cuff). Moonshiner (Moon) interacts with Del and recruits transcription initiation factors, allowing transcription of piRNA clusters in the sense and antisense direction. Bootlegger, UAP56, the THO complex, Nxt1, Nxf3 are recruited to the piRNA cluster transcription site. The transcripts are then exported to the cytoplasmic piRNAs biogenesis site via the CRM1 exportin. In the cytoplasm, piRNA precursors are cleaved by Zucchini (Zuc) into primary piRNA and charged by Piwi or Aubergine (Aub) proteins, forming Piwi-piRISC and Aub-piRISC complexes. Piwi-piRISC is translocated into the nucleus and recognizes nascent TE transcripts due to the interaction with Panoramix, the Nxf2, and Nxt1. The association of these factors to nascent TE transcripts causes recruitment of the histone methyltransferase Eggless that transcriptionally silences TEs by local heterochromatinization. In the cytoplasm, Aub-piRISC cleaves TE transcripts into secondary piRNAs, which are loaded onto Ago3, forming Ago3-piRISC complexes. Ago3-piRISC recognizes and cleaves piRNA precursors into new secondary piRNAs, loaded onto Aub. This mechanism, known as the ‘Ping-Pong’ cycle, allows the amplification of piRISC complexes and post-transcriptional regulation of TEs.</p>
Full article ">Figure 2
<p>The paramutation process in <span class="html-italic">Drosophila melanogaster</span> relies on maternal piRNA inheritance. At G0, females (grey chromosomes) carrying the <span class="html-italic">BX2</span> clusters made of repeated <span class="html-italic">P-lacZ-white</span> transgenes (blue arrowheads) producing piRNAs are crossed to males (white chromosomes) carrying the same cluster not producing piRNA. At G1, maternal piRNA deposition in the embryo leads to activation of piRNA production from the paternal cluster allele. The newly activated cluster can activate the paternally-inherited inactive cluster in G2 due to maternal piRNA inheritance (MpI). Then, this process of activation leads to stable piRNAs production across generation.</p>
Full article ">Figure 3
<p>Stable, heritable stress-induced de novo piRNAs. In flies, the <span class="html-italic">BX2</span> cluster, a sequence made of repeated <span class="html-italic">P-lacZ-white</span> transgenes (dark blue arrowheads), is capable of producing piRNA after one generation at 29 °C, then these piRNAs can functionally repress a homologous <span class="html-italic">P-lacZ</span> sequence (light blue arrowheads) in <span class="html-italic">trans</span>. This silencing capacity is maintained at the second generation after return to normal temperature (25 °C) and then up to 50 generations. Hence, when stress is removed, piRNAs can be self-maintained by maternal piRNA inheritance (MpI) at each generation.</p>
Full article ">
19 pages, 917 KiB  
Review
Functions and Regulatory Mechanisms of lncRNAs in Skeletal Myogenesis, Muscle Disease and Meat Production
by Shanshan Wang, Jianjun Jin, Zaiyan Xu and Bo Zuo
Cells 2019, 8(9), 1107; https://doi.org/10.3390/cells8091107 - 19 Sep 2019
Cited by 67 | Viewed by 6201
Abstract
Myogenesis is a complex biological process, and understanding the regulatory network of skeletal myogenesis will contribute to the treatment of human muscle related diseases and improvement of agricultural animal meat production. Long noncoding RNAs (lncRNAs) serve as regulators in gene expression networks, and [...] Read more.
Myogenesis is a complex biological process, and understanding the regulatory network of skeletal myogenesis will contribute to the treatment of human muscle related diseases and improvement of agricultural animal meat production. Long noncoding RNAs (lncRNAs) serve as regulators in gene expression networks, and participate in various biological processes. Recent studies have identified functional lncRNAs involved in skeletal muscle development and disease. These lncRNAs regulate the proliferation, differentiation, and fusion of myoblasts through multiple mechanisms, such as chromatin modification, transcription regulation, and microRNA sponge activity. In this review, we presented the latest advances regarding the functions and regulatory activities of lncRNAs involved in muscle development, muscle disease, and meat production. Moreover, challenges and future perspectives related to the identification of functional lncRNAs were also discussed. Full article
Show Figures

Figure 1

Figure 1
<p>Functional lncRNAs in skeletal muscle development.</p>
Full article ">Figure 2
<p>Functional lncRNAs involved in skeletal muscle disease.</p>
Full article ">
18 pages, 2389 KiB  
Article
Regulation of Ketogenic Enzyme HMGCS2 by Wnt/β-catenin/PPARγ Pathway in Intestinal Cells
by Ji Tae Kim, Chang Li, Heidi L. Weiss, Yuning Zhou, Chunming Liu, Qingding Wang and B. Mark Evers
Cells 2019, 8(9), 1106; https://doi.org/10.3390/cells8091106 - 19 Sep 2019
Cited by 48 | Viewed by 7866
Abstract
The Wnt/β-catenin pathway plays a crucial role in development and renewal of the intestinal epithelium. Mitochondrial 3-hydroxy-3-methylglutaryl-CoA synthase 2 (HMGCS2), a rate-limiting ketogenic enzyme in the synthesis of ketone body β-hydroxybutyrate (βHB), contributes to the regulation of intestinal cell differentiation. Here, we have [...] Read more.
The Wnt/β-catenin pathway plays a crucial role in development and renewal of the intestinal epithelium. Mitochondrial 3-hydroxy-3-methylglutaryl-CoA synthase 2 (HMGCS2), a rate-limiting ketogenic enzyme in the synthesis of ketone body β-hydroxybutyrate (βHB), contributes to the regulation of intestinal cell differentiation. Here, we have shown that HMGCS2 is a novel target of Wnt/β-catenin/PPARγ signaling in intestinal epithelial cancer cell lines and normal intestinal organoids. Inhibition of the Wnt/β-catenin pathway resulted in increased protein and mRNA expression of HMGCS2 and βHB production in human colon cancer cell lines LS174T and Caco2. In addition, Wnt inhibition increased expression of PPARγ and its target genes, FABP2 and PLIN2, in these cells. Conversely, activation of Wnt/β-catenin signaling decreased protein and mRNA levels of HMGCS2, βHB production, and expression of PPARγ and its target genes in LS174T and Caco2 cells and mouse intestinal organoids. Moreover, inhibition of PPARγ reduced HMGCS2 expression and βHB production, while activation of PPARγ increased HMGCS2 expression and βHB synthesis. Furthermore, PPARγ bound the promoter of HMGCS2 and this binding was enhanced by β-catenin knockdown. Finally, we showed that HMGCS2 inhibited, while Wnt/β-catenin stimulated, glycolysis, which contributed to regulation of intestinal cell differentiation. Our results identified HMGCS2 as a downstream target of Wnt/β-catenin/PPARγ signaling in intestinal epithelial cells. Moreover, our findings suggest that Wnt/β-catenin/PPARγ signaling regulates intestinal cell differentiation, at least in part, through regulation of ketogenesis. Full article
(This article belongs to the Section Cell Signaling)
Show Figures

Figure 1

Figure 1
<p>Inhibition of Wnt/β-catenin signaling increased 3-hydroxy-3-methylglutaryl-CoA synthase 2 (HMGCS2) expression in human intestinal cancer cells. (<b>A</b>,<b>B</b>) LS174T or Caco2 cells transfected with non-target control (NTC) siRNA or β-catenin (β-cat) siRNA were incubated for 48 h. (<b>A</b>) Western blot analysis was performed using the antibodies as indicated. HMGCS2 expression from three separate western blots was quantitated densitometrically and is expressed as fold change with respect to β-actin (<span class="html-italic">n</span> = 3, data represent mean ± SD; * <span class="html-italic">p</span> &lt; 0.05 vs. NTC). (<b>B</b>) The level of <span class="html-italic">HMGCS2</span> mRNA was assessed by real-time RT-PCR (<span class="html-italic">n</span> = 3, data represent mean ± SD; * <span class="html-italic">p</span> &lt; 0.05 vs. NTC). (<b>C</b>,<b>D</b>) Inhibition of Wnt/β-catenin signaling increased the expression of HMGCS2 in LS174T and Caco2 cells. LS174T or Caco2 cells were treated with iCRT3 for 24 h (LS174T) or 48 h (Caco2). (<b>C</b>) Western blot analysis was performed using the antibodies as indicated. (<b>D</b>) <span class="html-italic">HMGCS2</span> mRNA expression was assessed by real time RT-PCR (<span class="html-italic">n</span> = 3, data represent mean ± SD; * <span class="html-italic">p</span> &lt; 0.05 vs. 0 μM iCRT3).</p>
Full article ">Figure 2
<p>Activation of Wnt/β-catenin pathway suppressed HMGCS2 expression. (<b>A</b>,<b>B</b>) LS174T and Caco2 cells were treated with 200 ng/mL Wnt3a for 24 h. (<b>A</b>) Western blot analysis was performed using the antibodies as indicated. Densitometric quantification from three separate experiments was performed and is represented as fold change with respect to β-actin (<span class="html-italic">n</span> = 3, data represent mean ± SD; * <span class="html-italic">p</span> &lt; 0.05 vs. 0 ng/mL Wnt3a). (<b>B</b>) The level of <span class="html-italic">HMGCS2</span> mRNA was determined by real-time RT-PCR. (<span class="html-italic">n</span> = 3, data represent mean ± SD; * <span class="html-italic">p</span> &lt; 0.05 vs. 0 ng/mL Wnt3a). (<b>C</b>,<b>D</b>) LS174T and Caco2 cells were treated with various dosages of LiCl or 40 mM of NaCl as control for 24 h. (<b>C</b>) Western blot analysis was performed using the antibodies as indicated. (<b>D</b>) The level of <span class="html-italic">HMGCS2</span> mRNA was determined by real-time RT-PCR. (<span class="html-italic">n</span> = 3, data represent mean ± SD; * <span class="html-italic">p</span> &lt; 0.05 vs. 40 mM NaCl). (<b>E</b>). Mouse small intestinal organoids were treated with 100 ng/mL Wnt3a for 3 days. HMGCS2 protein expression was determined by western blotting. Densitometric analysis from three independent experiments was performed and is represented as fold change with respect to β-actin (<span class="html-italic">n</span> = 3, data represent mean ± SD; * <span class="html-italic">p</span> &lt; 0.05 vs. 0 ng/mL Wnt3a).</p>
Full article ">Figure 3
<p>Regulation of β-hydroxybutyrate (βHB) production by Wnt/β-catenin signaling in intestinal cells. (<b>A</b>) LS174T and Caco2 cells, transfected with NTC or β-catenin siRNA (β-cat) and incubated for 48 h or treated with iCRT3 for 24 h (LS174T) or 48 h (Caco2), were lysed and βHB content was measured using a βHB assay kit (<span class="html-italic">n</span> = 3, data represents mean ± SD; * <span class="html-italic">p</span> &lt; 0.05 vs. NTC or 0 μM iCRT3). (<b>B</b>) LS174T and Caco2 cells were treated with 200 ng/mL Wnt3a or 40 mM LiCl (40 mM NaCl as control) for 24 h. Cell lysates were used for measurement of βHB content using a βHB assay kit (<span class="html-italic">n</span> = 3, data represents mean ± SD; * <span class="html-italic">p</span> &lt; 0.05 vs. 0 ng/mL Wnt3a or 40 mM NaCl). (<b>C</b>) Mouse small intestinal organoids were treated with 100 ng/mL Wnt3a for 3 days. βHB content in cell lysates and conditioned media was determined using a βHB assay kit (<span class="html-italic">n</span> = 3, data represents mean ± SD; * <span class="html-italic">p</span> &lt; 0.05 vs. 0 ng/mL Wnt3a).</p>
Full article ">Figure 4
<p>Effect of c-Myc on HMGCS2 regulation by Wnt/β-catenin pathway in LS174T cells. (<b>A</b>) LS174T cells transfected with NTC or c-Myc siRNA were treated with iCRT3 for 24 h. (<b>B</b>) LS174T cells infected with Ad-GFP control or Ad-c-Myc were treated with iCRT3 (25 μM) for 24 h. (<b>C</b>) LS174T cells transfected with NTC or β-catenin were infected with Ad-GFP control or Ad-c-Myc. Western blot analysis was performed using the antibodies as indicated. HMGCS2 expression from three separate western blots was quantitated densitometrically and is expressed as fold change with respect to β-actin (<span class="html-italic">n</span> = 3, data represent mean ± SD; * <span class="html-italic">p</span> &lt; 0.05 vs. NTC or GFP control).</p>
Full article ">Figure 5
<p>The expression and activity of PPARγ was regulated by the Wnt/β-catenin pathway in intestinal cancer cell lines. (<b>A</b>,<b>B</b>) LS174T or Caco2 cells were transfected with NTC or β-catenin siRNA and incubated for 48 h. (<b>A</b>) Western blot analysis was performed using the antibodies as indicated. PPARγ expression from three independent experiments was quantitated densitometrically and is expressed as fold change with respect to β-actin (<span class="html-italic">n</span> = 3, data represent mean ± SD; * <span class="html-italic">p</span> &lt; 0.05 vs. NTC). (<b>B</b>) <span class="html-italic">PPAR</span>γ mRNA was assessed by real-time RT-PCR (<span class="html-italic">n</span> = 3, data represent mean ± SD; * <span class="html-italic">p</span> &lt; 0.05 vs. NTC). (<b>C</b>,<b>D</b>) LS174T or Caco2 cells were treated with NaCl (40 mM) as control or LiCl (40 mM) for 24 h. (<b>C</b>) Western blotting was performed using the antibodies as indicated. Densitometric analysis from three separate experiments was performed and PPARγ expression is represented as fold change with respect to β-actin (<span class="html-italic">n</span> = 3, data represent mean ± SD; * <span class="html-italic">p</span> &lt; 0.05 vs. 40 mM/mL NaCl). (<b>D</b>) <span class="html-italic">PPAR</span>γ mRNA was assessed by real-time RT-PCR (<span class="html-italic">n</span> = 3, data represents mean ± SD; * <span class="html-italic">p</span> &lt; 0.05 vs. 40 mM/mL NaCl control). (<b>E</b>) LS174T cells were transfected with NTC or β-catenin siRNA for 48 h or treated with 40 mM NaCl or 40 mM LiCl for 24 h and PPARγ DNA-binding activity was determined as described under Materials and Methods (<span class="html-italic">n</span> = 3, data represents mean ± SD; * <span class="html-italic">p</span> &lt; 0.05 vs. NTC or 40 mM NaCl). (<b>F</b>) Expression of <span class="html-italic">FABP2</span> and <span class="html-italic">PLIN2</span> mRNA was assessed by real-time RT-PCR. (<span class="html-italic">n</span> = 3, data represents mean ± SD; * <span class="html-italic">p</span> &lt; 0.05 vs. NTC or 40 mM/mL NaCl).</p>
Full article ">Figure 6
<p>PPARγ regulated HMGCS2 expression in intestinal cancer cell lines. (<b>A</b>). LS174T cells were transfected with NTC or PPARγ siRNA. Expression of PPARγ and HMGCS2 was determined by western blot analysis. HMGCS2 expression from three separate western blots was quantitated densitometrically and is expressed as fold change with respect to β-actin (<span class="html-italic">n</span> = 3, data represent mean ± SD; * <span class="html-italic">p</span> &lt; 0.05 vs. NTC). (<b>B</b>,<b>C</b>). (<b>B</b>) LS174T cells were treated with rosiglitazone (RGZ), an agonist of PPARγ for 24 h; (<b>C</b>) LS174T cells were treated with iCRT3 in the presence or absence of T 0070907 (T007), an antagonist of PPARγ for 24 h. Protein expression of HMGCS2 was measured by western blot analysis. Densitometric quantification from three independent experiments was performed and is represented as fold change with respect to β-actin (<span class="html-italic">n</span> = 3, data represent mean ± SD; * <span class="html-italic">p</span> &lt; 0.05 vs. 40 mM NaCl or vehicle control). The level of <span class="html-italic">HMGCS2</span> mRNA was assessed by real-time RT-PCR (<span class="html-italic">n</span> = 3, data represents mean ± SD; * <span class="html-italic">p</span> &lt; 0.05 vs. control). (<b>D</b>). LS174T cells were transfected with NTC or β-catenin siRNA. DNA was extracted and ChIP-qPCR was performed using PPARγ or IgG antibody. The binding efficiency of PPARγ to the promoter regions of <span class="html-italic">HMGCS2</span> or <span class="html-italic">IFN-λ1</span> (as non-regulated control (NRC)) was analyzed and is represented as a percent of input (<span class="html-italic">n</span> = 3, data represents mean ± SD; * <span class="html-italic">p</span> &lt; 0.05 vs. NTC). (<b>E</b>). LS174T cells were treated with RGZ for 24 h, or transfected with NTC or β-catenin siRNA; Caco2 cells were treated with RGZ or T007 for 24 h. Cells were lysed and βHB content was determined using a βHB assay kit. (<span class="html-italic">n</span> = 3, data represents mean ± SD; * <span class="html-italic">p</span> &lt; 0.05 vs. control or NTC).</p>
Full article ">Figure 7
<p>Regulation of glycolysis by Wnt/β-catenin/HMGCS2 pathway. (<b>A</b>) LS174T cells transfected with NTC siRNA or siRNA targeting HMGCS2 or β-catenin were subjected to Seahorse Extracellular Flux analysis (left) (<span class="html-italic">n</span> = 9, data represent mean ± SD; * <span class="html-italic">p</span> &lt; 0.05 vs. NTC). Knockdown of HMGCS2 and β-catenin was confirmed by western blot analysis (right). (<b>B</b>,<b>C</b>) LS174T cells were treated with 2-DG, a glycolysis inhibitor for 24 h. Caco2 cells were treated with 2-DG for 24 h (RNA) or 48 h (IAP activity assay). (<b>B</b>) IAP activity was measured (<span class="html-italic">n</span> = 3, data represents mean ± SD; * <span class="html-italic">p</span> &lt; 0.05 vs. 0 mM 2-DG). (<b>C</b>) Expression of <span class="html-italic">SI</span>, <span class="html-italic">p21<sup>Waf1</sup></span> and <span class="html-italic">CDX2</span> mRNA was assessed by real-time RT-PCR. (<span class="html-italic">n</span> = 3, data represents mean ± SD; * <span class="html-italic">p</span> &lt; 0.05 vs. 0 mM 2-DG). (<b>D</b>) Inhibition of the Wnt/β-catenin pathway resulted in the increased expression and activation of PPARγ, and thus increased ketogenesis by induction of HMGCS2. Ketogenesis contributes to intestinal cell differentiation via the inhibition of glycolysis.</p>
Full article ">
18 pages, 2784 KiB  
Review
Cellular Stress Responses in Radiotherapy
by Wanyeon Kim, Sungmin Lee, Danbi Seo, Dain Kim, Kyeongmin Kim, EunGi Kim, JiHoon Kang, Ki Moon Seong, HyeSook Youn and BuHyun Youn
Cells 2019, 8(9), 1105; https://doi.org/10.3390/cells8091105 - 18 Sep 2019
Cited by 212 | Viewed by 16888
Abstract
Radiotherapy is one of the major cancer treatment strategies. Exposure to penetrating radiation causes cellular stress, directly or indirectly, due to the generation of reactive oxygen species, DNA damage, and subcellular organelle damage and autophagy. These radiation-induced damage responses cooperatively contribute to cancer [...] Read more.
Radiotherapy is one of the major cancer treatment strategies. Exposure to penetrating radiation causes cellular stress, directly or indirectly, due to the generation of reactive oxygen species, DNA damage, and subcellular organelle damage and autophagy. These radiation-induced damage responses cooperatively contribute to cancer cell death, but paradoxically, radiotherapy also causes the activation of damage-repair and survival signaling to alleviate radiation-induced cytotoxic effects in a small percentage of cancer cells, and these activations are responsible for tumor radio-resistance. The present study describes the molecular mechanisms responsible for radiation-induced cellular stress response and radioresistance, and the therapeutic approaches used to overcome radioresistance. Full article
(This article belongs to the Special Issue Molecular and Cellular Mechanisms of Stress Responses)
Show Figures

Figure 1

Figure 1
<p>Radiation-induced reactive oxygen species (ROS) response associated with p53 signaling. Irradiation increases intracellular ROS levels facilitated by radiation-mediated mitochondrial damage. In the presence of elevated ROS levels, p53 may importantly ameliorate radiation-induced oxidative stress.</p>
Full article ">Figure 2
<p>Radiation-induced double-strand breaks (DSB) response. When DSBs are induced by irradiation, DNA damage-sensing and repair proteins such as ATM, ATR, DNA-PK, H2AX, MDC1, Chk1, and Chk2 are activated. Subsequently, p53 is activated and induces cell cycle arrest for the homologous recombination (HR) or non-homologous end joining (NHEJ) pathways or induces apoptosis by upregulating proapoptotic genes.</p>
Full article ">Figure 3
<p>Radiation-induced lipid peroxidation and ceramide signaling. Exposure of the plasma membrane to penetrating radiation leads to the production of homologous recombination (HNE), arachidonic acid-derived lipid metabolites, and ceramide. HNE is associated with the stimulation of unfolded protein response (UPR), and arachidonic acid metabolites promote cell proliferation, inflammation, and protect cells from apoptosis, and thus, contribute to tumor radioresistance. On the other hand, ceramide triggers apoptosis by activating Fas and Bak/Bax signaling and inhibiting PI3K/Akt signaling.</p>
Full article ">Figure 4
<p>Radiation-induced mitochondrial response. Radiation induces mitochondrial damage largely via ROS generation. Excessive ROS levels and radiation-induced p53-dependent upregulations of PUMA and Bak/Bax result in mitochondrial membrane permeabilization and subsequent release of cytochrome <span class="html-italic">c</span> into cytosol, and thus, promote intrinsic apoptotic signaling.</p>
Full article ">Figure 5
<p>Radiation-induced ER stress response. Radiation can induce ER stress directly or indirectly by generating ROS. Under radiation-induced ER stress, specific signaling by PERK, ATF6, and IRE1 may be activated, and augment the upregulations of UPR-related genes to improve chaperone activity and induce autophagy to recover and recycle misfolded proteins.</p>
Full article ">
18 pages, 4489 KiB  
Article
Interleukin 21 Receptor/Ligand Interaction Is Linked to Disease Progression in Pancreatic Cancer
by Alica Linnebacher, Philipp Mayer, Nicole Marnet, Frank Bergmann, Esther Herpel, Steffie Revia, Libo Yin, Li Liu, Thilo Hackert, Thomas Giese, Ingrid Herr and Matthias M. Gaida
Cells 2019, 8(9), 1104; https://doi.org/10.3390/cells8091104 - 18 Sep 2019
Cited by 14 | Viewed by 4785
Abstract
Pancreatic ductal adenocarcinoma (PDAC) displays a marked fibro-inflammatory microenvironment in which infiltrated immune cells fail to eliminate the tumor cells and often—rather paradoxically—promote tumor progression. Of special interest are tumor-promoting T cells that assume a Th17-like phenotype because their presence in PDAC tissue [...] Read more.
Pancreatic ductal adenocarcinoma (PDAC) displays a marked fibro-inflammatory microenvironment in which infiltrated immune cells fail to eliminate the tumor cells and often—rather paradoxically—promote tumor progression. Of special interest are tumor-promoting T cells that assume a Th17-like phenotype because their presence in PDAC tissue is associated with a poor prognosis. In that context, the role of IL-21, a major cytokine released by Th17-like cells, was assessed. In all tissue samples (n = 264) IL-21+ immune cells were detected by immunohistochemistry and high density of those cells was associated with poor prognosis. In the majority of patients (221/264), tumor cells expressed the receptor for IL-21 (IL-21R) and also a downstream target of IL-21, Blimp-1 (199/264). Blimp-1 expression closely correlated with IL-21R expression and multivariate analysis revealed that expression of both IL-21R and Blimp-1 was associated with shorter survival time of the patients. In vitro data using pancreatic tumor cells lines provided a possible explanation: IL-21 activated ERK and STAT3 pathways and upregulated Blimp-1. Moreover, IL-21 increased invasion of tumor cell lines in a Blimp-1-dependent manner. As an in vivo correlate, an avian xenograft model was used. Here again Blimp-1 expression was significantly upregulated in IL-21 stimulated tumor cells. In summary, our data showed an association of IL-21+ immune cell infiltration and IL-21 receptor expression in PDAC with poor survival, most likely due to an IL-21-mediated promotion of tumor cell invasion and enhanced colony formation, supporting the notion of the tumor-promoting abilities of the tumor microenvironment. Full article
(This article belongs to the Special Issue Molecular and Cellular Mechanisms of Cancers: Pancreatic Cancer)
Show Figures

Figure 1

Figure 1
<p>IL-21R, IL-21, and Blimp-1 expression in human PDAC and healthy pancreas tissue samples. (<b>A</b>) Infiltration of IL-21<sup>+</sup> immune cells (red) in PDAC. Black arrow: IL-21<sup>+</sup> cells. (<b>B</b>) Infiltration of IL-21<sup>+</sup> Th17 cells in PDAC: IL-21<sup>+</sup> (red) and RORC<sup>+</sup> (brown) cells. Black arrowhead: PDAC tumor cells, black arrow: IL-21<sup>+</sup>/RORC<sup>+</sup> immune cells. (<b>C</b>) Examples of negative, weak and high expression of IL-21R (brown) and Blimp-1 (red) in human PDAC tissue. Black arrows: tumor cells. (<b>D</b>) Examples of IL-21R and Blimp-1 expression in healthy pancreas tissue. (<b>E</b>) IL-21R expression in human PDAC varied among the patients. (<b>F</b>) Association of IL-21R and Blimp-1 expression in human PDAC. * <span class="html-italic">P</span> &lt; 0.05 and **** <span class="html-italic">P</span> &lt; 0.0001. Unpaired <span class="html-italic">t</span>-test. (<b>G</b>) Co-expression of IL-21R (brown, membrane) and Blimp-1 (red, nucleus). Black arrowhead: IL-21R<sup>+</sup>/Blimp-1<sup>+</sup>, white arrow head: IL-21R<sup>+</sup>/Blimp-1<sup>−</sup>. Black bar: 20 µm.</p>
Full article ">Figure 1 Cont.
<p>IL-21R, IL-21, and Blimp-1 expression in human PDAC and healthy pancreas tissue samples. (<b>A</b>) Infiltration of IL-21<sup>+</sup> immune cells (red) in PDAC. Black arrow: IL-21<sup>+</sup> cells. (<b>B</b>) Infiltration of IL-21<sup>+</sup> Th17 cells in PDAC: IL-21<sup>+</sup> (red) and RORC<sup>+</sup> (brown) cells. Black arrowhead: PDAC tumor cells, black arrow: IL-21<sup>+</sup>/RORC<sup>+</sup> immune cells. (<b>C</b>) Examples of negative, weak and high expression of IL-21R (brown) and Blimp-1 (red) in human PDAC tissue. Black arrows: tumor cells. (<b>D</b>) Examples of IL-21R and Blimp-1 expression in healthy pancreas tissue. (<b>E</b>) IL-21R expression in human PDAC varied among the patients. (<b>F</b>) Association of IL-21R and Blimp-1 expression in human PDAC. * <span class="html-italic">P</span> &lt; 0.05 and **** <span class="html-italic">P</span> &lt; 0.0001. Unpaired <span class="html-italic">t</span>-test. (<b>G</b>) Co-expression of IL-21R (brown, membrane) and Blimp-1 (red, nucleus). Black arrowhead: IL-21R<sup>+</sup>/Blimp-1<sup>+</sup>, white arrow head: IL-21R<sup>+</sup>/Blimp-1<sup>−</sup>. Black bar: 20 µm.</p>
Full article ">Figure 2
<p>Correlation of IL-21 receptor with Blimp-1 expression and effect on overall survival in human PDAC biopsies. Survival analysis of IL-21R and Blimp-1 expression. There was a non-significant trend towards shorter OS in PDAC patients with high IL-21R expression compared to patients with low IL-21R expression. PDAC patients expressing a high level of Blimp-1 had a significantly shorter survival compared to patients expressing a low level of Blimp-1. Survival was significantly shorter in patients with high IL-21 infiltrate compared to patients with low IL-21 infiltrate.</p>
Full article ">Figure 3
<p>IL-21R expression in human PDAC cell lines. Untreated human PDAC cell lines AsPC-1, BxPC-3, and Panc-1 were tested for IL-21R expression in (<b>A</b>) immunofluorescence staining (left: IL-21R<sup>+</sup> cells, right: Alexa488 goat anti-rabbit IgG only; nuclei: DAPI (blue), IL-21R: Alexa488 (green)) and (<b>B</b>) western blot. White bar: 20 µm.</p>
Full article ">Figure 4
<p>IL-21R activates ERK and STAT3 pathway; but not p38. PDAC cells BxPC-3 and Panc-1 were incubated with IL-21 (10 ng/mL) for the time indicated. Phosphorylation of (<b>A</b>) ERK, (<b>B</b>) STAT3, and (<b>C</b>) p38 was evaluated. β-tubulin was used as a loading control. After 10–20 minutes, phosphorylation of ERK and STAT3 could be detected. One of three independent experiments is shown.</p>
Full article ">Figure 5
<p>IL-21 upregulates Blimp-1 in an ERK-dependent manner. BxPC-3 and Panc-1 were incubated with IL-21 (10 and 25 ng/mL) for 24 h. (<b>A</b>) Treatment with IL-21 upregulates Blimp-1 expression. One of six independent experiments is shown. (<b>B</b>) Quantification of IL-21-dependent Blimp-1 upregulation (IL-21: 10 ng/mL). * <span class="html-italic">P</span> &lt; 0.05. Wilcoxon rank-sum test. (<b>C</b>) ERK inhibitor prevents IL-21-dependent Blimp-1 upregulation. One of five independent experiments is shown.</p>
Full article ">Figure 5 Cont.
<p>IL-21 upregulates Blimp-1 in an ERK-dependent manner. BxPC-3 and Panc-1 were incubated with IL-21 (10 and 25 ng/mL) for 24 h. (<b>A</b>) Treatment with IL-21 upregulates Blimp-1 expression. One of six independent experiments is shown. (<b>B</b>) Quantification of IL-21-dependent Blimp-1 upregulation (IL-21: 10 ng/mL). * <span class="html-italic">P</span> &lt; 0.05. Wilcoxon rank-sum test. (<b>C</b>) ERK inhibitor prevents IL-21-dependent Blimp-1 upregulation. One of five independent experiments is shown.</p>
Full article ">Figure 6
<p>Treatment of avian xenografts with IL-21 upregulates Blimp-1 expression. (<b>A</b>) The three-dimensional growth of BxPC-3 cells transplanted on the CAM of fertilized chicken eggs and resected tumors (three replicates) from vehicle control and IL-21 were photographed at day 17 of embryonic development. Upper row for vehicle control; lower row for IL-21-treated (150 ng/mL) tumors. (<b>B</b>) HE and IL-21R (brown) IHC staining of vehicle control and IL-21-treated (150 ng/mL) resected xenograft tumors. (<b>C</b>) Comparison and quantification of Blimp-1 (nuclei: red) expression of vehicle control and IL-21-treated (150 ng/mL) xenograft tumors. **** <span class="html-italic">P</span> &lt; 0.0001. Unpaired <span class="html-italic">t</span>-test.</p>
Full article ">Figure 7
<p>IL-21 promotes cell invasion in human PDAC cell lines in a Blimp-1-dependent manner. (<b>A</b>) Invasion of either untreated (RPMI) or IL-21-treated (10 ng/mL) BxPC-3 and Panc-1 cells was determined after 48 h. (<b>B</b>) Invasion of BxPC-3 and Panc-1 cells transfected with either a scrambled siRNA or a <span class="html-italic">PRDM1</span>-targeting siRNA (siRNA C) for Blimp-1 protein knockdown after 48 h of IL-21 (10 ng/mL) treatment. (<b>C</b>) Invasion of BxPC-3 and Panc-1 transfected with pLenti EV or pLenti Blimp-1 construct for CRISPR/Cas9-mediated Blimp-1 knockout after 48 h of IL-21 (10 ng/mL) treatment. (<b>D</b>) Invasion of BxPC-3 and Panc-1 transfected with a dox-inducible pTRIPZ EV or pTRIPZ Blimp-1 construct for Blimp-1 overexpression. Relative fold change of invaded cells treated with (dox<sup>+</sup>) and without dox (dox<sup>−</sup>) (1 µg/mL). Left panel: invasion of untreated (RPMI) cells, right panel: invasion of IL-21-treated (10 ng/mL) cells * <span class="html-italic">P</span> &lt; 0.05 and ** <span class="html-italic">P</span> &lt; 0.01. Unpaired <span class="html-italic">t</span>-test.</p>
Full article ">Figure 8
<p>IL-21 treatment has no profound effect on tumor cell proliferation. (<b>A</b>) Proliferation of BxPC-3 and Panc-1 was measured by a BrdU incorporation method at the indicated time points. Data from four independent experiments are shown. (<b>B</b>) Quantification of tumor volume and the ratio of tumor volume to chick weight. Unpaired <span class="html-italic">t</span>-test. (<b>C</b>) Comparison and quantification of Ki67 (nuclei: brown) and FVIII (tubular structure: brown) expression of vehicle control and IL-21-treated (150 ng/mL) xenograft tumors. Arrows indicate positive cells. Black bar: 100 µm. Unpaired <span class="html-italic">t</span>-test.</p>
Full article ">
18 pages, 3800 KiB  
Article
Understanding the Modus Operandi of MicroRNA Regulatory Clusters
by Arthur C. Oliveira, Luiz A. Bovolenta, Lucas Alves, Lucas Figueiredo, Amanda O. Ribeiro, Vinicius F. Campos, Ney Lemke and Danillo Pinhal
Cells 2019, 8(9), 1103; https://doi.org/10.3390/cells8091103 - 18 Sep 2019
Cited by 12 | Viewed by 4422
Abstract
MicroRNAs (miRNAs) are non-coding RNAs that regulate a wide range of biological pathways by post-transcriptionally modulating gene expression levels. Given that even a single miRNA may simultaneously control several genes enrolled in multiple biological functions, one would expect that these tiny RNAs have [...] Read more.
MicroRNAs (miRNAs) are non-coding RNAs that regulate a wide range of biological pathways by post-transcriptionally modulating gene expression levels. Given that even a single miRNA may simultaneously control several genes enrolled in multiple biological functions, one would expect that these tiny RNAs have the ability to properly sort among distinctive cellular processes to drive protein production. To test this hypothesis, we scrutinized previously published microarray datasets and clustered protein-coding gene expression profiles according to the intensity of fold-change levels caused by the exogenous transfection of 10 miRNAs (miR-1, miR-7, miR-9, miR-124, miR-128a, miR-132, miR-133a, miR-142, miR-148b, miR-181a) in a human cell line. Through an in silico functional enrichment analysis, we discovered non-randomic regulatory patterns, proper of each cluster identified. We demonstrated that miRNAs are capable of equivalently modulate the expression signatures of target genes in regulatory clusters according to the biological function they are assigned to. Moreover, target prediction analysis applied to ten vertebrate species, suggest that such miRNA regulatory modus operandi is evolutionarily conserved within vertebrates. Overall, we discovered a complex regulatory cluster-module strategy driven by miRNAs, which relies on the controlled intensity of the repression over distinct targets under specific biological contexts. Our discovery helps to clarify the mechanisms underlying the functional activity of miRNAs and makes it easier to take the fastest and most accurate path in the search for the functions of miRNAs in any distinct biological process of interest. Full article
(This article belongs to the Collection Regulatory Functions of microRNAs)
Show Figures

Figure 1

Figure 1
<p>Correlation between the number of enriched terms and the number of miRNA-responsive genes in each cluster. Clusters are ordered according to the number of gene members embraced.</p>
Full article ">Figure 2
<p>Functional enrichment of messenger RNA (mRNA) clusters after miR-133a transfection. Purple terms belongs to the cluster (−2.156:−0.542), red terms to the cluster (−0.539:−0.119), green terms to the cluster (−0.118:0.119), blue terms to the cluster (0.120:0.666) and yellow terms to the cluster (0.674:4.128). The numbers within brackets represent the minimum and maximum fold change values of each cluster. Terms with different colors inside a cluster indicate that the same term is also present in the cluster of that color. Grey edges connect highly similar terms, and the edge width indicates the similarity degree. Results for other miRNAs can be visualized at <a href="#app1-cells-08-01103" class="html-app">Supplementary Figure S4</a> and a detailed description of the terms found in each cluster can be visualized at <a href="#app1-cells-08-01103" class="html-app">Supplementary Table S3</a>.</p>
Full article ">Figure 3
<p>Combined action of distinct regulatory clusters. Each line represents a regulatory cluster connecting miRNAs to the enriched biological term associated with the cluster. The numbers within brackets represent the minimum and maximum fold change values of that cluster. Red lines represent up-regulated gene clusters, green lines represent down-regulated gene clusters, and gray dashed lines represent clusters of genes apparently not regulated by the miRNA.</p>
Full article ">Figure 4
<p>Proportion of genes predicted and do not predicted as targeted by miRNAs belonging to each cluster of miRNA-responsive genes for miR-7, miR-128, and miR-181a. Additional data of other miRNAs can be found in the <a href="#app1-cells-08-01103" class="html-app">Supplementary Figure S5</a>.</p>
Full article ">Figure 5
<p>miRNA-responsive biological pathways and the presence or absence of interactions between its genes and the transfected miRNA. (<b>A</b>) predicted interactions among miR-9 (yellow ellipses) and the mRNAs of a biological pathway perturbed after miR-9 transfection; (<b>B</b>) miR-1 responsive biological pathway with no predicted interaction between miR-1 and its genes; (<b>C</b>) predicted interactions among miR-142 (yellow ellipses) and genes either up- or down-regulated after miR-142 transfection. Green rectangles represent down-regulated mRNAs. Red rectangles represent up-regulated mRNAs. Gray rectangles represent mRNAs with little-to-no fold change after miRNA transfection.</p>
Full article ">Figure 6
<p>Conservation of target responsiveness over miRNA interaction. (<b>A</b>). Heatmaps of Context++Scores for miR-1 and miR-9 predicted targets among 10 vertebrate species. Negative Context++Score values are represented in green, positive values are represented in red, and near zero values are represented in black. Lower Context++Score values represent stronger miRNA regulation levels. Heatmaps for other miRNAs can be found in the <a href="#app1-cells-08-01103" class="html-app">Supplementary Figure S7</a>. (<b>B</b>). Example of evolutionary conservation of the 3’UTR of two genes targeted by miR-1 (PICALM) or miR-9 (GDNF) in the genome of the ten vertebrate species. Yellow marks represent the miRNA binding site at the target mRNA 3’UTR.</p>
Full article ">Figure 7
<p>MiRNA-mediated mRNA up-regulation scheme. (<b>A</b>) Interaction network between the miRNA, the up-regulated mRNA, and its suppressor. The miRNA down-regulates mRNA suppressor while fine-tunes the expression of the up-regulated mRNA. The continuous line represents direct interaction and dashed line represents either direct or indirect regulation; (<b>B</b>) Representative expression variation of the network members. Green line represents miRNA expression, the blue line represents up-regulated mRNA expression and the red line represents mRNA suppressor expression.</p>
Full article ">
35 pages, 1887 KiB  
Review
Trends and Challenges in Tumor Anti-Angiogenic Therapies
by József Jászai and Mirko H.H. Schmidt
Cells 2019, 8(9), 1102; https://doi.org/10.3390/cells8091102 - 18 Sep 2019
Cited by 160 | Viewed by 15277
Abstract
Excessive abnormal angiogenesis plays a pivotal role in tumor progression and is a hallmark of solid tumors. This process is driven by an imbalance between pro- and anti-angiogenic factors dominated by the tissue hypoxia-triggered overproduction of vascular endothelial growth factor (VEGF). VEGF-mediated signaling [...] Read more.
Excessive abnormal angiogenesis plays a pivotal role in tumor progression and is a hallmark of solid tumors. This process is driven by an imbalance between pro- and anti-angiogenic factors dominated by the tissue hypoxia-triggered overproduction of vascular endothelial growth factor (VEGF). VEGF-mediated signaling has quickly become one of the most promising anti-angiogenic therapeutic targets in oncology. Nevertheless, the clinical efficacy of this approach is severely limited in certain tumor types or shows only transient efficacy in patients. Acquired or intrinsic therapy resistance associated with anti-VEGF monotherapeutic approaches indicates the necessity of a paradigm change when targeting neoangiogenesis in solid tumors. In this context, the elaboration of the conceptual framework of “vessel normalization” might be a promising approach to increase the efficacy of anti-angiogenic therapies and the survival rates of patients. Indeed, the promotion of vessel maturation instead of regressing tumors by vaso-obliteration could result in reduced tumor hypoxia and improved drug delivery. The implementation of such anti-angiogenic strategies, however, faces several pitfalls due to the potential involvement of multiple pro-angiogenic factors and modulatory effects of the innate and adaptive immune system. Thus, effective treatments bypassing relapses associated with anti-VEGF monotherapies or breaking the intrinsic therapy resistance of solid tumors might use combination therapies or agents with a multimodal mode of action. This review enumerates some of the current approaches and possible future directions of treating solid tumors by targeting neovascularization. Full article
(This article belongs to the Special Issue Angiogenesis in Cancer)
Show Figures

Figure 1

Figure 1
<p>Steps of tissue hypoxia-triggered neoangiogenesis and progression of neoplastic lesions. VEGF: vascular endothelial growth factor. The initially avascular tumor cell mass, upon reaching a critical size, cannot ensure its further growth without the trophic support of its own circulation. Hypoxic cells of the tumor mass, via angiogenic factors, initiate the “angiogenic switch” by stimulating nearby endothelial cells of the microvasculature of the host/parental tissue in which they arise. Tumor angiogenesis initiates with a detachment of perivascular cells, the degradation of vessel basal lamina and angiogenic sprouting by the formation of filopodia bearing leading-edge endothelial tip cells from the parental vessels. Newly formed sprouts build anastomoses followed by lumen formation and the recruitment of the pericytic cells necessary for perivascular investment. Angiogenesis will continue as the expansive growth of the tumor mass requires further supply. This process forms a vicious circle that is fueled by the tumor vessel leakiness and the suboptimal tumor perfusion that cannot efficiently relief tissue hypoxia. Recruited inflammatory cells contribute significantly to a hostile microenvironment that boosts further uncontrolled neoangiogenesis and play a role in the evasive resistance of solid tumors. Abbreviations: Treg, regulatory T cell; TAM, tumor-associated macrophage; TEM, Tie-2- expressing monocyte.</p>
Full article ">Figure 2
<p>The VEGF/VEGFR signaling axis, its contribution to neoangiogenesis and treatment modalities interfering with its activity. Binding of VEGF ligands to their cognate receptors leads to receptor dimerization and autophosphorylation triggering a down-stream intracellular phosphorylation cascade. In principle, tumor anti-angiogenesis can be achieved (1) by prohibiting ligand binding to their cognate TK receptors (VEGFR1-3) receptors/non-tyrosine kinase (NRPs) co-receptors either by the withdrawal of pro-angiogenic ligands of the VEGF family (Aflibercept, Bevacizumab) or by blocking the accessibility of the binding pocket for ligands on a particular receptor (Ramucirumab); (2) by interfering with the kinase activity of VEGFRs (small molecule multikinase inhibitors: receptor tyrosine kinase (RTK)-inhibitors). Anti-VEGF-targeted therapies result in the inhibition of down-stream signaling mechanisms, governing a range of steps involved in neovessel formation and/or inflammatory infiltration. Abbreviations: EC, endothelial cell; MΦ, macrophage; PlGF, placental growth factor; RTK, receptor tyrosine kinase; TK-domain, tyrosine kinase-domain; VEGF, vascular endothelial growth factor; VEGFR, vascular endothelial growth factor receptor.</p>
Full article ">Figure 3
<p>Alternative angiogenic pathways playing a role in tumor angiogenesis, refractoriness, evasive resistance or relapses in response to anti-VEGF monotherapies and tools available for targeting their effects. Resistance to anti-angiogenic therapy might be mediated by angiogenic factors that trigger a switch from an initially VEGF-dependent state to a VEGF-independent angiogenic process. In order to interfere with these alternative pathways, besides the small molecule multikinase inhibitors (RTK-inhibitors) blocking down-stream signal cascade activation even in the presence of receptor occupancy, several recombinant biological tools have been developed. The range of these tools includes decoys (“ligand traps”) aimed at withdrawing alternative pro-angiogenic factors either alone (fibroblast growth factor (FGF)-Trap) or simultaneously with VEGF-A (VF-Trap), engineered multivalent monoclonals (CrossMabs) or antibody tools raised against RTK receptors blocking the access of ligands to their cognate receptor (onartuzumab). Abbreviations: EC, endothelial cell; PC, pericyte; TC, tumor cell; Ang-2, angiopoietin-2; FGF-2, fibroblast growth factor-2; FGFR, fibroblast growth factor receptor; HGF, hepatocyte growth factor; MET, mesenchymal-epithelial transition proto-oncogene; PDGF, platelet-derived growth factor; PDGFR, platelet-derived growth factor receptor; TIE2, Tyrosine kinase with immunoglobulin-like and EGF-like domains 2; RTK, receptor tyrosine kinase.</p>
Full article ">
14 pages, 1904 KiB  
Article
Increased Glycolysis and Higher Lactate Production in Hyperglycemic Myotubes
by Jenny Lund, D. Margriet Ouwens, Marianne Wettergreen, Siril S. Bakke, G. Hege Thoresen and Vigdis Aas
Cells 2019, 8(9), 1101; https://doi.org/10.3390/cells8091101 - 18 Sep 2019
Cited by 35 | Viewed by 7104
Abstract
Previous studies have shown that chronic hyperglycemia impairs glucose and fatty acid oxidation in cultured human myotubes. To further study the hyperglycemia-induced suppression of oxidation, lactate oxidation, mitochondrial function and glycolytic rate were evaluated. Further, we examined the intracellular content of reactive oxygen [...] Read more.
Previous studies have shown that chronic hyperglycemia impairs glucose and fatty acid oxidation in cultured human myotubes. To further study the hyperglycemia-induced suppression of oxidation, lactate oxidation, mitochondrial function and glycolytic rate were evaluated. Further, we examined the intracellular content of reactive oxygen species (ROS), production of lactate and conducted pathway-ANOVA analysis on microarray data. In addition, the roles of the pentose phosphate pathway (PPP) and the hexosamine pathway were evaluated. Lactic acid oxidation was suppressed in hyperglycemic versus normoglycaemic myotubes. No changes in mitochondrial function or ROS concentration were observed. Pathway-ANOVA analysis indicated several upregulated pathways in hyperglycemic cells, including glycolysis and PPP. Functional studies showed that glycolysis and lactate production were higher in hyperglycemic than normoglycaemic cells. However, there were no indications of involvement of PPP or the hexosamine pathway. In conclusion, hyperglycemia reduced substrate oxidation while increasing glycolysis and lactate production in cultured human myotubes. Full article
Show Figures

Figure 1

Figure 1
<p>Effects of hyperglycemia on lactic acid and glucose oxidation. Myotubes were exposed to 20 mM glucose (HG) or standard differentiation medium (NG, 5.5 mM glucose) the last 4 days of the differentiation period, and then incubated with either [<sup>14</sup>C(U)]lactic acid (1 µCi/mL, 100 µM) or <span class="html-small-caps">d</span>-[<sup>14</sup>C(U)]glucose (0.5 µCi/mL, 200 µM) for 4 h. Oxidation was measured as CO<sub>2</sub> trapped in a filter and counted by liquid scintillation. (<b>a</b>) Lactic acid oxidation after chronic HG. Results are presented as means ± SEM in nmol/mg protein from five individual experiments (n = 5). (<b>b</b>) Glucose oxidation after chronic HG. Results are presented as means ± SEM in nmol/mg protein from 15 individual experiments (n = 15). * Statistically significant vs. NG (<span class="html-italic">p</span> &lt; 0.05, paired Student’s <span class="html-italic">t</span>-test).</p>
Full article ">Figure 2
<p>Effects of chronic hyperglycemia on mitochondrial function. Human skeletal muscle cells were grown in 24-well Seahorse tissue culture plates, exposed to 20 mM glucose (HG) for the last four days of the differentiation period, or standard differentiation medium (NG, 5.5 mM glucose), before measurement of the oxygen consumption rate (OCR) with the Seahorse XF24e analyzer. OCR was recorded three times at 6 min intervals at baseline, and following injections with 5 µM oligomycin (Oli), 3 µM FCCP and 4 µM rotenone/antimycin A (Rot/AA), respectively (XF Cell Mito Stress Test Kit). Determinants of mitochondrial function were calculated as described in <a href="#sec2dot5-cells-08-01101" class="html-sec">Section 2.5</a>. (<b>a</b>) One representative experiment. (<b>b</b>–<b>g</b>) Mean ± SEM from five individual experiments (n = 5). (<b>b</b>): basal respiration calculated from OCR in pmol/min, (<b>c</b>): maximal respiration calculated from OCR in pmol/min, (<b>d</b>): proton leak calculated from OCR in pmol/min, (<b>e</b>): percentage coupling efficiency, (<b>f</b>): mitochondrial (Mito) ATP production rate in pmol/min, and (<b>g</b>): glycolytic (Glyco) ATP production rate in pmol/min. * Statistically significant vs. NG (<span class="html-italic">p</span> &lt; 0.05, unpaired Student’s <span class="html-italic">t</span>-test).</p>
Full article ">Figure 3
<p>Involvement of reactive oxygen species (ROS). (<b>a</b>) Myotubes were exposed to 20 mM glucose (hyperglycemia, HG) the last four days of the differentiation period and ROS production in HG versus normoglycaemic (NG) myotubes were calculated from a standard curve (OxiSelect intracellular ROS assay kit). Results are presented as means ± SEM from five individual experiments (n = 5). (<b>b</b>) Effect of the ROS-scavenger MitoTEMPO on glucose oxidation. Results are presented as means ± SEM in % of NG basal from four individual experiments (n = 4).</p>
Full article ">Figure 4
<p>Pathways involved in metabolic processes regulated by hyperglycemia. Myotubes were exposed to 20 mM glucose (HG) or normal glucose concentration (NG) for the last four days of the differentiation period. Gene expression was measured using the Illumina HumanWG-6 v3.0 expression BeadChip (microarray). Pathway-ANOVA was performed to identify pathways regulated by HG compared to NG in myotubes from three donors (n = 3). Selected significantly (FDR <span class="html-italic">p</span> &lt; 0.05) regulated pathways with relation to carbohydrate metabolism are presented.</p>
Full article ">Figure 5
<p>Effects of chronic hyperglycemia on the glycolytic rate. Human skeletal muscle cells were grown in 24-well Seahorse tissue culture plates, exposed to 20 mM glucose (HG) for the last four days of the differentiation period, or the standard differentiation medium (NG, 5.5 mM glucose), before measurement of the glycolytic rate with the Seahorse XF24e analyzer. Extracellular acidification rates and oxygen consumption rates were recorded six times at 6 min intervals at baseline, and following injections with 4 µM rotenone/antimycin A (Rot/AA) and 50 mM 2-deoxyglucose (2-DG), respectively. Proton efflux rate (PER), glycolytic proton efflux rate (glycoPER), basal glycolysis, basal PER, compensatory glycolysis, and post 2-DOG acidification were calculated as described in <a href="#sec2dot6-cells-08-01101" class="html-sec">Section 2.6</a>. (<b>a</b>) One representative experiment. (<b>b</b>–<b>e</b>) Mean ± SEM from five individual experiments (n = 5). (<b>b</b>): basal glycolysis glycoPER in pmol/min, (<b>c</b>): basal PER in pmol/min, (<b>d</b>): percentage PER from glycolysis, and (<b>e</b>): compensatory glycolysis glycoPER in pmol/min. * Statistically significant vs. NG (<span class="html-italic">p</span> &lt; 0.05, unpaired Student’s <span class="html-italic">t</span>-test).</p>
Full article ">Figure 6
<p>Concentration of lactate in cell media. Myotubes were treated with 20 mM glucose (HG) or 5 mM glucose (NG) for four days. Thereafter, the media was removed and lactate concentration measured using Accutrend Plus. Results are presented as means ± SEM in mM lactate from five individual experiments (n = 5). * Statistically significant vs. NG (<span class="html-italic">p</span> &lt; 0.05, paired Student’s <span class="html-italic">t</span>-test).</p>
Full article ">Figure 7
<p>Involvement of the pentose phosphate pathway. Myotubes were exposed to 5.5 mM (NG) or 20 mM glucose (HG) with or without 50 µM 6-aminonicotinamide (6-AN) the last four days of the differentiation period and then incubated with <span class="html-small-caps">d</span>-[<sup>14</sup>C(U)]glucose (0.5 µCi/mL, 200 µM) for 4 h. Oxidation was measured as CO<sub>2</sub> trapped in a filter and counted by liquid scintillation. Data are presented as means ± SEM relative to NG basal from four individual experiments (n = 4). Data representing NG basal (100%): 26.6 ± 10.3 nmol/mg protein.</p>
Full article ">Figure 8
<p>Involvement of the hexosamine pathway. (<b>a</b>) Myotubes were treated with 20 mM glucose (HG) or 5 mM glucose (NG) for four days in the presence of 4 mM (basal) or 10 mM <span class="html-small-caps">l</span>-glutamine before they were incubated with <span class="html-small-caps">d</span>-[<sup>14</sup>C(U)]glucose (0.5 µCi/mL, 200 µM) for 4 h. Glucose oxidation was measured as CO<sub>2</sub> trapped in a filter and counted by liquid scintillation. (<b>b</b>) Effects of the deglycosylation inhibitor PUGNAc (100 µM) and the glycosylation inhibitor azaserine (20 µM) in protein lysates from HG and NG cells. The inhibitors were added together with HG the last four days of the differentiation period. Aliquots of 15 µg cell protein from total cell lysates, including a positive control (cell lysate A549), were electrophoretically separated on 10% polyacrylamide gels, followed by immunoblotting with specific antibody for <span class="html-italic">O</span>-GlcNAc. One representative Western blot from three individual experiments (n = 3) is shown. (<b>c</b>) Effect of the deglycosylation inhibitor PUGNAc (100 µM) on glucose oxidation. (<b>d</b>) Effect of the glycosylation inhibitor azaserine (20 µM) on glucose oxidation. Results are presented as means ± SEM in nmol/mg protein from three individual experiments (n = 3).</p>
Full article ">
15 pages, 1545 KiB  
Review
New Insights into the Liver–Visceral Adipose Axis During Hepatic Resection and Liver Transplantation
by María Eugenia Cornide-Petronio, Mónica B. Jiménez-Castro, Jordi Gracia-Sancho and Carmen Peralta
Cells 2019, 8(9), 1100; https://doi.org/10.3390/cells8091100 - 18 Sep 2019
Cited by 8 | Viewed by 4029
Abstract
In the last decade, adipose tissue has emerged as an endocrine organ with a key role in energy homeostasis. In addition, there is close crosstalk between the adipose tissue and the liver, since pro- and anti-inflammatory substances produced at the visceral adipose tissue [...] Read more.
In the last decade, adipose tissue has emerged as an endocrine organ with a key role in energy homeostasis. In addition, there is close crosstalk between the adipose tissue and the liver, since pro- and anti-inflammatory substances produced at the visceral adipose tissue level directly target the liver through the portal vein. During surgical procedures, including hepatic resection and liver transplantation, ischemia–reperfusion injury induces damage and regenerative failure. It has been suggested that adipose tissue is associated with both pathological or, on the contrary, with protective effects on damage and regenerative response after liver surgery. The present review aims to summarize the current knowledge on the crosstalk between the adipose tissue and the liver during liver surgery. Therapeutic strategies as well as the clinical and scientific controversies in this field are discussed. The different experimental models, such as lipectomy, to evaluate the role of adipose tissue in both steatotic and nonsteatotic livers undergoing surgery, are described. Such information may be useful for the establishment of protective strategies aimed at regulating the liver–visceral adipose tissue axis and improving the postoperative outcomes in clinical liver surgery. Full article
(This article belongs to the Special Issue Adipose Tissue Inflammation 2022)
Show Figures

Figure 1

Figure 1
<p>Schematic illustration of functional differences between lean and obese adipose tissue. Abbreviations: IL, interleukin; NO, nitric oxide; TNFα, tumor necrosis factor α.</p>
Full article ">Figure 2
<p>Strategies aimed at regulating hepatic damage and regenerative failure considering the adipose tissue–liver crosstalk during partial hepatectomy (PH) with or without I/R [<a href="#B54-cells-08-01100" class="html-bibr">54</a>,<a href="#B55-cells-08-01100" class="html-bibr">55</a>,<a href="#B56-cells-08-01100" class="html-bibr">56</a>,<a href="#B57-cells-08-01100" class="html-bibr">57</a>,<a href="#B58-cells-08-01100" class="html-bibr">58</a>,<a href="#B59-cells-08-01100" class="html-bibr">59</a>,<a href="#B60-cells-08-01100" class="html-bibr">60</a>]. Abbreviations: I/R, ischemia–reperfusion; PH, partial hepatectomy; sFlt1, soluble form of the VEGF receptor 1; SIP1, smad interacting protein 1; VEGFA, vascular endothelial growth factor type A.</p>
Full article ">Figure 3
<p>Strategies aimed at evaluating the role of adipokines on hepatic damage and regenerative failure in LT [<a href="#B61-cells-08-01100" class="html-bibr">61</a>,<a href="#B62-cells-08-01100" class="html-bibr">62</a>,<a href="#B63-cells-08-01100" class="html-bibr">63</a>]. The adipose tissue–liver crosstalk during LT is still unknown. Abbreviations: AICAR, Cell-permeable adenosine analog that is a selective activator of AMPK; LT, liver transplantation; PC, preconditioning.</p>
Full article ">Figure 4
<p>Schematic illustration of lipectomy effects in liver surgery. Abbreviations: ATP, adenosine triphosphate; I/R, ischemia–reperfusion; LT, liver transplantation; PH, partial hepatectomy; sFlt1, soluble form of the VEGF receptor 1; VEGFA, vascular endothelial growth factor type A.</p>
Full article ">
15 pages, 3748 KiB  
Article
CX3CR1 Mediates the Development of Monocyte-Derived Dendritic Cells during Hepatic Inflammation
by Salvatore Sutti, Stefania Bruzzì, Felix Heymann, Anke Liepelt, Oliver Krenkel, Alberto Toscani, Naresh Naik Ramavath, Diego Cotella, Emanuele Albano and Frank Tacke
Cells 2019, 8(9), 1099; https://doi.org/10.3390/cells8091099 - 18 Sep 2019
Cited by 28 | Viewed by 5752
Abstract
Recent evidence suggests that hepatic dendritic cells (HDCs) contribute to the evolution of chronic liver diseases. However, the HDC subsets involved and the mechanisms driving these responses are still poorly understood. In this study, we have investigated the role of the fractalkine receptor [...] Read more.
Recent evidence suggests that hepatic dendritic cells (HDCs) contribute to the evolution of chronic liver diseases. However, the HDC subsets involved and the mechanisms driving these responses are still poorly understood. In this study, we have investigated the role of the fractalkine receptor CX3CR1 in modulating monocyte-derived dendritic cell (moDC) differentiation during liver inflammation. The phenotype of HDC and functional relevance of CX3CR1 was assessed in mice following necro-inflammatory liver injury induced by the hepatotoxic agent carbon tetrachloride (CCl4) and in steatohepatitis caused by a methionine/choline-deficient (MCD) diet. In both the experimental models, hepatic inflammation was associated with a massive expansion of CD11c+/MHCIIhigh/CD11b+ myeloid HDCs. These cells also expressed the monocyte markers Ly6C, chemokine (C-C Motif) receptor 2 (CCR2), F4/80 and CD88, along with CX3CR1, allowing their tentative identification as moDCs. Mice defective in CX3CR1 showed a reduction in liver-moDC recruitment following CCl4 poisoning in parallel with a defective maturation of monocytes into moDCs. The lack of CX3CR1 also affected moDC differentiation from bone marrow myeloid cells induced by granulocyte-macrophage colony stimulating factor (GM-CSF) and interleukin-4 (IL-4) in vitro. In wild-type mice, treatment with the CX3CR1 antagonist CX3-AT (150 µg, i.p.) 24 h after CCl4 administration reduced liver moDCS and significantly ameliorated hepatic injury and inflammation. Altogether, these results highlight the possible involvement of moDCs in promoting hepatic inflammation following liver injury and indicated a novel role of CX3CL1/CX3CR1 dyad in driving the differentiation of hepatic moDCs. Full article
(This article belongs to the Special Issue Recent Advances in Liver Repair Strategies)
Show Figures

Figure 1

Figure 1
<p>Hepatic inflammation induced by the acute administration of CCl<sub>4</sub> associates with the expansion and maturation of hepatic dendritic cells (HDCs). Parenchymal damage and lobular inflammation were analyzed in wild-type mice either naïve (Cont) or 36 h after receiving an acute dose of CCl<sub>4</sub> (CCl<sub>4</sub>). (<b>A</b>) Hematoxylin/eosin staining of formalin-fixed liver sections (magnification 10×). (<b>B</b>) Circulating levels of alanine aminotransferase (ALT) and aspartate aminotransferase (AST). (<b>C</b>) RT-PCR analysis of hepatic expression of the pro-inflammatory cyto/chemokines TNF-α, CCL2, CXCL1 and CX<sub>3</sub>CL1. The values are expressed as fold increase over control levels and are means ± SD of 6–8 animals in each experimental group. (<b>D</b>) The changes in the liver distribution of CD11c<sup>+</sup>/MHCII<sup>high</sup>/CD11b<sup>+</sup>/CD103<sup>−</sup> HDCs were analyzed by flow cytometry in mice either untreated or receiving CCl<sub>4</sub>. (<b>E</b>) The plasma membrane expression of maturation marker CD80 was evaluated in HDCs gated for CD11b. The values are expressed as means ± SD of three different cell preparations.</p>
Full article ">Figure 2
<p>Characterization of hepatic dendritic cells (HDCs) expanding in response to acute liver injury. (<b>A</b>) CD11c<sup>+</sup>/MHCII<sup>high</sup>/CD11b<sup>+</sup> HDCs from either naïve (Cont) or CCl<sub>4</sub>-treated mice (CCl<sub>4</sub>) were analyzed by flow cytometry for the expression of CX<sub>3</sub>CR1 and the monocyte markers Ly6C, F4-80 and CCR2. The values are expressed as means ± SD of three different cell preparations. (<b>B</b>) Relative distribution of the dendritic cell and macrophage markers CD26, CD35 and CD88 among CD11c<sup>+</sup>/MHCII<sup>high</sup>/CD11b<sup>+</sup>/Ly6C<sup>+</sup>/CX<sub>3</sub>CR1<sup>+</sup> HDCs.</p>
Full article ">Figure 3
<p>Monocyte-derived dendritic cells (moDCs) account for HDC expansion associated with the onset of non-alcoholic steatohepatitis (NASH). Steatohepatitis was induced by feeding wild-type mice with a choline/methionine-deficient (MCD) diet for one week. (<b>A</b>) Hematoxylin/eosin staining of formalin-fixed liver sections (magnification 10×). (<b>B</b>) Circulating levels of alanine aminotransferase (ALT) and liver content of triglycerides. (<b>C</b>) RT-PCR analysis of the hepatic expression of the pro-inflammatory cyto/chemokines TNF-α, CCL2. Values are expressed as means ± SD of 5–6 animals in each experimental group. (<b>D</b>) Flow cytometry analysis of the changes in the hepatic distribution of CD11c<sup>+</sup>/MHCII<sup>high</sup>/CD11b<sup>high</sup>/CD88<sup>+</sup>/Ly6c<sup>+</sup> moDCs analyzed in mice either untreated (Cont) or receiving the an MCD diet. Values are expressed as means ± SD of three different cell preparations.</p>
Full article ">Figure 4
<p>The lack of CX<sub>3</sub>CR1 reduced the differentiation of monocyte-derived dendritic cells (moDCs) in response to liver inflammation. The liver distribution of CD11b<sup>+</sup> myeloid dendritic cells, moDCs and monocytes were analyzed by flow cytometry in the livers of CX<sub>3</sub>CR1<sup>gfp/+</sup> and CX<sub>3</sub>CR1<sup>gfp/gfp</sup> mice 36 h after receiving an acute dose of CCl<sub>4</sub>. (<b>A</b>) The prevalence of CD11c<sup>+</sup>/MHCII<sup>high</sup>/CD11b<sup>+</sup> myeloid dendritic cells and CD11c<sup>+</sup>/MHCII<sup>high</sup>/CD11b<sup>+</sup>/CD88<sup>+</sup> moDCs in CX<sub>3</sub>CR1<sup>gfp/+</sup> and CX<sub>3</sub>CR1<sup>gfp/gfp</sup> mice receiving CCl<sub>4</sub>. (<b>B</b>) Hepatic distribution of Ly6G<sup>−</sup>/CD11b<sup>high</sup>/Ly6C<sup>high</sup> monocytes in control and CCl<sub>4</sub>-treated CX<sub>3</sub>CR1<sup>gfp/+</sup> and CX<sub>3</sub>CR1<sup>gfp/gfp</sup> mice. (<b>C</b>) Impaired expression of MHCII by CD11c<sup>+</sup>/CD11b<sup>+</sup> myeloid cells in CCl<sub>4</sub>-treated CX<sub>3</sub>CR1<sup>gfp/+</sup> and CX<sub>3</sub>CR1<sup>gfp/gfp</sup> mice. The values are expressed as means ± SD of three different cell preparations or 5–6 animals.</p>
Full article ">Figure 5
<p>A lack of CX<sub>3</sub>CR1 affects the in vitro differentiation of monocyte-derived dendritic cells (moDCs). MoDCs were obtained in vitro by a seven-day culture of bone marrow myeloid cells from either CX<sub>3</sub>CR1<sup>gfp/+</sup> or CX<sub>3</sub>CR1<sup>gfp/gfp</sup> mice with granulocyte-macrophage colony stimulating factor (GM-CSF) and interleukin-4 (IL-4). (<b>A</b>) CX<sub>3</sub>CR1 expression in Ly6G<sup>−</sup>/CD11b<sup>+</sup>/CD88<sup>+</sup>/CD11c<sup>+</sup>/MHCII<sup>high</sup> moDC originating from GM-CF/IL-4-treated bone marrow myeloid cells. (<b>B</b>) The effect of CX<sub>3</sub>CR1 on the in vitro differentiation of CD11b<sup>+</sup>/CD88<sup>+</sup>/CD11c<sup>+</sup>/MHCII<sup>high</sup> moDCs. (<b>C</b>) Effect of CX<sub>3</sub>CR1 on the expression of the transcription factors Zbtb46, IRF-4 and IRF-8 implicated in moDC differentiation. The values are expressed as means ± SD of 3–5 different cell preparations.</p>
Full article ">Figure 6
<p>The CX<sub>3</sub>CL1 antagonist CX3-AT reduces the expansion of monocyte-derived dendritic cells (moDCs) in response to hepatic injury. Liver dendritic cells were analyzed by flow cytometry in mice receiving CCl<sub>4</sub> alone or in combination with CX3-AT. (<b>A</b>,<b>B</b>) Liver distribution of CD11b<sup>+</sup>/Ly6C<sup>+</sup> moDCs and CX<sub>3</sub>CR1-expressing moDCs. (<b>C</b>) Plasma membrane expression of the maturation marker CD80 in moDCs for CCl<sub>4</sub>-treated mice receiving or not receiving CX3-AT. The values are expressed as means ± SD of three different cell preparations. (<b>D</b>) RT-PCR analysis of the hepatic transcripts for CX<sub>3</sub>CL1 and CX<sub>3</sub>CR1. The values are expressed as fold increase over control levels and are means ± SD of 6–8 animals in each experimental group.</p>
Full article ">Figure 7
<p>The CX<sub>3</sub>CL1 antagonist CX3-AT improves liver injury and inflammation in mice receiving CCl<sub>4</sub>. Parenchymal damage and lobular inflammation were analyzed in mice receiving CCl<sub>4</sub> alone and in combination with CX3-AT. (<b>A</b>) Hematoxilin/eosin staining of formalin-fixed liver sections (magnification 10× and 40×). (<b>B</b>) Circulating levels of alanine aminotransferase (ALT) and aspartate aminotransferase (AST). (<b>C</b>,<b>D</b>) RT-PCR analysis of the hepatic expression of pro-inflammatory cyto/chemokines TNF-α, CCL2 and CXCL1, as well as IL-10. The values are expressed as fold increase over control levels and are means ± SD of 6–8 animals in each experimental group.</p>
Full article ">
16 pages, 1886 KiB  
Review
How Cancer Exploits Ribosomal RNA Biogenesis: A Journey beyond the Boundaries of rRNA Transcription
by Marco Gaviraghi, Claudia Vivori and Giovanni Tonon
Cells 2019, 8(9), 1098; https://doi.org/10.3390/cells8091098 - 17 Sep 2019
Cited by 28 | Viewed by 6029
Abstract
The generation of new ribosomes is a coordinated process essential to sustain cell growth. As such, it is tightly regulated according to cell needs. As cancer cells require intense protein translation to ensure their enhanced growth rate, they exploit various mechanisms to boost [...] Read more.
The generation of new ribosomes is a coordinated process essential to sustain cell growth. As such, it is tightly regulated according to cell needs. As cancer cells require intense protein translation to ensure their enhanced growth rate, they exploit various mechanisms to boost ribosome biogenesis. In this review, we will summarize how oncogenes and tumor suppressors modulate the biosynthesis of the RNA component of ribosomes, starting from the description of well-characterized pathways that converge on ribosomal RNA transcription while including novel insights that reveal unexpected regulatory networks hacked by cancer cells to unleash ribosome production. Full article
(This article belongs to the Section Intracellular and Plasma Membranes)
Show Figures

Figure 1

Figure 1
<p>Transcriptional and post-translational regulation of rRNA transcription by oncogenic pathways. (<b>A</b>) Schematic representation of the main components of the polymerase I (PolI) pre-initiation complex (PIC) and the major activatory (upper panel, orange) and inhibitory (lower panel, purple) post-translational modifications (PTMs) induced by oncogenes or tumor suppressor genes. (<b>B</b>) Schematic representation of the convergent regulatory pathways that boost rRNA transcription upon oncogene activation (left panel, orange) or repress it through tumor suppressor genes (right panel, purple).</p>
Full article ">Figure 2
<p>Regulation of rRNA processing in cancer. Schematic representation of the 47S rRNA precursors processing pathways (grey) generating the mature 18S, 5.8S, and 28S molecules (black). The main maturation steps affected by oncogenes (orange ovals) or tumor suppressors (purple ovals) are represented.</p>
Full article ">Figure 3
<p>Small nucleolar RNAs (snoRNAs) sustain the oncogenic potential of cancer genes. (<b>A</b>) C/D box snoRNAs are required to sustain oncogene-induced self-renewal and proliferation of acute myeloid leukemia (AML) cells. (<b>B</b>) U3 and U8 snoRNAs are targets of oncogenes (orange ovals) and tumor suppressors (purple ovals) and are essential to promote oncogene-induced cell proliferation. Oncogene stimulation induces U3 stabilization and putatively prevents U3 and U8 decapping, mediated by a DCP1α/DCP2 complex, tethered inside nucleoli by PNRC1 tumor suppressor.</p>
Full article ">
22 pages, 8302 KiB  
Article
DNA Damage Changes Distribution Pattern and Levels of HP1 Protein Isoforms in the Nucleolus and Increases Phosphorylation of HP1β-Ser88
by Soňa Legartová, Gabriela Lochmanová, Zbyněk Zdráhal, Stanislav Kozubek, Jiří Šponer, Miroslav Krepl, Pavlína Pokorná and Eva Bártová
Cells 2019, 8(9), 1097; https://doi.org/10.3390/cells8091097 - 17 Sep 2019
Cited by 8 | Viewed by 5834
Abstract
The family of heterochromatin protein 1 (HP1) isoforms is essential for chromatin packaging, regulation of gene expression, and repair of damaged DNA. Here we document that γ-radiation reduced the number of HP1α-positive foci, but not HP1β and HP1γ foci, located in the vicinity [...] Read more.
The family of heterochromatin protein 1 (HP1) isoforms is essential for chromatin packaging, regulation of gene expression, and repair of damaged DNA. Here we document that γ-radiation reduced the number of HP1α-positive foci, but not HP1β and HP1γ foci, located in the vicinity of the fibrillarin-positive region of the nucleolus. The additional analysis confirmed that γ-radiation has the ability to significantly decrease the level of HP1α in rDNA promoter and rDNA encoding 28S rRNA. By mass spectrometry, we showed that treatment by γ-rays enhanced the HP1β serine 88 phosphorylation (S88ph), but other analyzed modifications of HP1β, including S161ph/Y163ph, S171ph, and S174ph, were not changed in cells exposed to γ-rays or treated by the HDAC inhibitor (HDACi). Interestingly, a combination of HDACi and γ-radiation increased the level of HP1α and HP1γ. The level of HP1β remained identical before and after the HDACi/γ-rays treatment, but HDACi strengthened HP1β interaction with the KRAB-associated protein 1 (KAP1) protein. Conversely, HP1γ did not interact with KAP1, although approximately 40% of HP1γ foci co-localized with accumulated KAP1. Especially HP1γ foci at the periphery of nucleoli were mostly absent of KAP1. Together, DNA damage changed the morphology, levels, and interaction properties of HP1 isoforms. Also, γ-irradiation-induced hyperphosphorylation of the HP1β protein; thus, HP1β-S88ph could be considered as an important marker of DNA damage. Full article
(This article belongs to the Special Issue Nucleolar Organization and Functions in Health and Disease)
Show Figures

Figure 1

Figure 1
<p>Nuclear distribution pattern of heterochromatin protein 1 (HP1) isoforms in non-irradiated and γ-irradiated human cervix adenocarcinoma (HeLa) cells. Morphology of (<b>a</b>) HP1α (see GFP-tagged HP1α, green), (<b>b</b>) HP1β (see GFP-tagged HP1β, green), and (<b>c</b>) HP1γ (see GFP-tagged HP1γ, green) was studied by laser scanning confocal microscopy. The HP1-positive foci (green) were analyzed in the vicinity of fibrillarin-positive regions of nucleoli (red fluorescence signals). DAPI staining (blue) was used for visualization of nuclei. Scale bars show 5 µm. (<b>d</b>) Analysis of the number of HP1α-, HP1β-, and HP1γ-positive foci in non-irradiated and γ-irradiated (5 Gy) cells. Cells were fixed for analysis 2 h after γ-irradiation. Fifty cell nuclei per sample were studied. Asterisk in panel (d) shows significantly different result.</p>
Full article ">Figure 2
<p>ChIP-polymerase chain reaction (ChIP-PCR) analysis of HP1α, HP1β, HP1γ, KAP1 abundance in ribosomal genes. (<b>a</b>) The analysis was performed in non-treated human cervix adenocarcinoma (HeLa) cells, and cells irradiated by 5 Gy of γ-rays. Cells were harvested 2 h after γ-irradiation. The highest density in rDNA encoding 28S rRNA and rDNA promoter region was observed for HP1γ. (<b>b</b>) Quantification of fragment densities, shown in panel (<b>a</b>), was done by ImageJ software. White asterisk shows statistically significant difference, shown by Student’s t-test at <span class="html-italic">p</span> ≤ 0.05.</p>
Full article ">Figure 3
<p>Characterization of HP1β protein in human cervix adenocarcinoma (HeLa) cells using MS. Glu-C endoproteinase was used for protein digestion. After filtering the LC-MS/MS data in Proteome Discoverer 2.2 (see Methods for details), 75% sequence coverage was obtained (marked in bold type). Identified PTMs, including phosphorylation (P) and acetylation (A) are indicated. The amino acid sequence corresponding to CD and CSD domain is shown in blue or orange, respectively, with hinge region in between.</p>
Full article ">Figure 4
<p>The proportion of signature marks on HP1β protein in non-treated, γ-irradiated, and SAHA treated human cervix adenocarcinoma (HeLa) cells. Box-plots of the relative abundance of particular phosphopeptides showing extremes, interquartile ranges, means and medians (<span class="html-italic">N</span> = 3). Precursor peak of phosphopeptides was quantified in Skyline SW. Data were normalized to the sum of selected non-modified peptides. Following post-translational modifications were analyzed: (<b>a</b>) HP1β-S88ph; (<b>b</b>) HP1β-Y163/S161ph; (<b>c</b>) HP1β-S171ph; (<b>d</b>) HP1β-S174ph; (<b>e</b>) HP1β-S171ph-S174ph. Differences between samples in normalized peptide abundances ≥1.5-fold were considered as significant, and the significance of differences was assessed using Student’s t-tests, setting the significance threshold at <span class="html-italic">P</span> &lt; 0.05 (shown by the asterisk). Cells were harvested 2 h after the treatment; 5Gy of γ-rays or 15 µM SAHA.</p>
Full article ">Figure 5
<p>Representative MS and MS/MS spectra of HP1 G81−E96 peptide carrying phosphorylation at S88 (<span class="html-italic">m/z</span> 596.586). (<b>a</b>) MS extracted ion chromatograms of the 3+ charged HP1β-S88ph peptide precursor showing changes in its abundance after γ-irradiation (5 Gy) and SAHA (15 μM) treatment. The three colored lines correspond to the ion chromatogram of the monoisotopic peptide mass and the first two isotopes. (<b>b</b>) MS/MS spectrum produced from the precursor ion of <span class="html-italic">m/z</span> 596.586. The deposition of the mass spectrometry proteomics data see at the ProteomeXchange Consortium (for detailed information, see Methodology section).</p>
Full article ">Figure 6
<p>Nuclear distribution pattern of 53BP1 foci and phosphorylated histone H2AX. (<b>a</b>) Close proximity or co-localization between 53BP1 foci (red) and phosphorylated histone H2AX (γH2AX, green) in control, γ-irradiated (5 Gy) and SAHA (15 μM) treated human cervix adenocarcinoma (HeLa) cells. Cells were harvested 2 h after the treatment. Scale bars show 5 µm. Irradiation by γ-rays increased the number of (<b>b</b>) γH2AX and (<b>c</b>) 53BP1-positive foci. (<b>d</b>) The volume of γH2AX foci in control, γ-irradiated (5 Gy) and SAHA (15 μM) treated cells remain unchanged. (<b>e</b>) SAHA treatment enlarged the volume of 53BP1-positive foci, as indicated by asterisk showing a large scale of 53BP1-foci volume. Irradiation by γ-rays reduced the number of 53BP1-positive foci (blue box). Immunofluorescent data were quantified using the ImageJ software. Fifty cell nuclei per sample were studied. The box plots are displaying the distribution of data based on the summary of five numbers: minimum, first quartile, median, third quartile, and maximum. Asterisks show significantly different results from control values, at <span class="html-italic">p</span> ≤ 0.05. Student’s t-test was used for data analysis.</p>
Full article ">Figure 7
<p>Fluorescence Lifetime Imaging-Forster Resonance Energy Transfer (FLIM-FRET) analysis of potential interaction between heterochromatin protein 1 (HP1) protein isoforms and the KAP1 protein. (<b>a</b>) Analysis of a GFP-tagged HP1α-KAP1, b] GFP-tagged HP1β-KAP1, and c] GFP-tagged HP1γ-KAP1. (<b>b</b>) An example of a nuclear distribution of HP1γ and the KAP1 protein. HP1γ-positive regions of nucleoli (green) were absent of KAP1. Abbreviation Nu means nucleoli; arrows show HP1γ foci absent of KAP1 and red frames show HP1γ foci colocalizing with KAP1. (<b>c</b>) Example of STED analysis showing the location of KAP1 inside the cell nucleoli decorated by the HP1 isoforms (see HP1β in white frame). (<b>d</b>) FLIM-FRET analysis of a] GFP-tagged HP1β ΔCD/Alexa 594-KAP1, b] GFP-tagged HP1β ΔCSD/ Alexa 594-KAP1, and c] GFP-tagged HP1β ΔHinge/Alexa 594-KAP1. E [%] means FLIM-FRET efficiency. Scale bars represent 4 µm.</p>
Full article ">Figure 8
<p>Immunoprecipitation analysis of potential interaction between HP1α, HP1β, and HP1γ. Studies of following possible interactions: (<b>a</b>) HP1α-HP1α, HP1α-HP1β, HP1α-HP1γ; HP1α-KAP1; (<b>b</b>) HP1β-HP1α, HP1β-HP1β, HP1β-HP1γ; HP1β-KAP1; (<b>c</b>) HP1γ-HP1α, HP1γ-HP1β, HP1γ-HP1γ, HP1γ-KAP1. Protein-protein interaction was studied in non-treated HeLa cells, and cells treated with SAHA (15 µM) and irradiated by γ-rays (5 Gy). Cells were harvested for analysis 2 h the treatment. (<b>d</b>) Western blot analysis of following proteins: HP1α, HP1β, HP1γ, KAP1, normalized to the total protein levels and α-tubulin, and H3K9ac, H3S10ph, γH2AX, normalized to the level of total histone H3. (<b>e</b>) Data from large panel (<b>d</b>) were quantified by ImageJ software. Levels of HP1 isoforms were normalized to the level of α-tubulin, and γH2AX or H3S10ph levels were normalized to the level of total histone H3. Profiles of HP1α, HP1β, HP1γ, KAP1, H3S10ph, and γH2AX were studied in (no. 1) non-irradiated HeLa cells, γ-irradiated Hela cells by (no. 2) 5 Gy of γ-rays/harvested after 2h; (no. 3) 2 Gy/harvested after 10 min; (no. 4) 2 Gy/harvested after 30 min; (no. 5) combination of SAHA treatment with 2 Gy/harvested after 10 min; and (no. 6) combination of SAHA treatment with 2 Gy/harvested after 30 min. A small panel in (<b>d</b>) is showing western blot data for non-treated HeLa cells, and cells irradiated by 5 Gy of γ-rays or SAHA (15 µM) treated HeLa cells.</p>
Full article ">Figure 9
<p>Effects of γ-irradiation by 5 Gy of γ-rays, SAHA treatment, and a combination of 2 Gy with SAHA treatment. The number and area of γH2AX-positive foci were measured. Data were compared with non-treated cells. Immunofluorescent data were quantified using the ImageJ software. Fluorescence intensity (FI) was analyzed, and foci with FI ≥ 100 were used for analysis. Number of repair foci are shown as a minimal and maximal value, and area of γ-positive foci is shown as mean ± standard errors (S.E.). Asterisks show statistically significant results at <span class="html-italic">p</span> ≤ 0.05. Student’s t-test was used for statistical analysis. Scale bars show 5 µm.</p>
Full article ">Figure 10
<p>Fluorescence lifetime imaging-Forster resonance energy transfer (FLIM-FRET) analysis of potential interaction between HP1α, HP1β, and HP1γ. Following interactions were studied: (<b>a</b>) a) GFP-tagged HP1α/mCherry-HP1β, b) GFP-tagged HP1β/ mCherry-HP1β, <b>c]</b> mCherry-HP1β/GFP-tagged HP1γ. E (%) means FLIM-FRET efficiency. FLIM-FRET analysis of (<b>b</b>) a) GFP-tagged HP1β ΔCD/mCherry-HP1β, b) GFP-tagged HP1β ΔCSD/mCherry-HP1β, and <b>c</b>) GFP-tagged HP1β ΔHinge/mCherry-HP1β. E (%) means FLIM-FRET efficiency. Scale bars represent 2 µm.</p>
Full article ">
23 pages, 6196 KiB  
Article
Reappraisal of Human HOG and MO3.13 Cell Lines as a Model to Study Oligodendrocyte Functioning
by Kim M. A. De Kleijn, Wieteke A. Zuure, Jolien Peijnenborg, Josje M. Heuvelmans and Gerard J. M. Martens
Cells 2019, 8(9), 1096; https://doi.org/10.3390/cells8091096 - 17 Sep 2019
Cited by 21 | Viewed by 7910
Abstract
Myelination of neuronal axons is essential for proper brain functioning and requires mature myelinating oligodendrocytes (myOLs). The human OL cell lines HOG and MO3.13 have been widely used as in vitro models to study OL (dys) functioning. Here we applied a number of [...] Read more.
Myelination of neuronal axons is essential for proper brain functioning and requires mature myelinating oligodendrocytes (myOLs). The human OL cell lines HOG and MO3.13 have been widely used as in vitro models to study OL (dys) functioning. Here we applied a number of protocols aimed at differentiating HOG and MO3.13 cells into myOLs. However, none of the differentiation protocols led to increased expression of terminal OL differentiation or myelin-sheath formation markers. Surprisingly, the applied protocols did cause changes in the expression of markers for early OLs, neurons, astrocytes and Schwann cells. Furthermore, we noticed that mRNA expression levels in HOG and MO3.13 cells may be affected by the density of the cultured cells. Finally, HOG and MO3.13 co-cultured with human neuronal SH-SY5Y cells did not show myelin formation under several pro-OL-differentiation and pro-myelinating conditions. Together, our results illustrate the difficulty of inducing maturation of HOG and MO3.13 cells into myOLs, implying that these oligodendrocytic cell lines may not represent an appropriate model to study the (dys)functioning of human (my)OLs and OL-linked disease mechanisms. Full article
(This article belongs to the Special Issue The Molecular and Cellular Basis for Parkinson's Disease)
Show Figures

Figure 1

Figure 1
<p>Normalized mRNA expression in undifferentiated HOG cells (uHOG) and HOG cells following differentiation (dHOG) with (<b>A</b>) N2.1 medium, (<b>B</b>) N2.2 medium or (<b>C</b>) T3 medium. OL: oligodendrocyte lineage, OPC: oligodendrocyte precursor, imOL: immature oligodendrocyte, mOL: mature oligodendrocyte, myOL: myelinating oligodendrocyte, NEU: neuronal, NP: neural progenitor, AST: astrocyte, iSC: immature Schwann cell, mSC: mature Schwann cell. Independent samples <span class="html-italic">t</span>-tests are based on triplicates in three independent experiments (<span class="html-italic">n</span> = 3). <span class="html-italic">p</span>-values: * &lt; 0.05, ** &lt; 0.01, *** &lt; 0.001.</p>
Full article ">Figure 2
<p>Analysis of protein expression in undifferentiated and differentiated HOG cells. (<b>A</b>) Western blot analyses of myelin basic protein (MBP) (polyclonal), myelin-oligodendrocyte glycoprotein (MOG), CNPase, MBP (monoclonal), class III β-tubulin (TUBB3), proteolipid protein 1 (PLP1) and paraoxonase 2 (PON2) in (+) human motor cortex, (1) undifferentiated HOG cells, (2) differentiated HOG cells following incubation in N2.1 medium, (3) N2.2 medium or (4) T3 medium (normalized to <span class="html-italic">GAPDH</span> expression). Western blot analysis was performed for two independent experiments (<span class="html-italic">n</span> = 2). (<b>B</b>) Example images of immunocytochemistry for CNPase and TUBB3 in undifferentiated HOG cells and HOG cells differentiated with N2.1 medium or N2.2 medium. Scale bar = 50 µm.</p>
Full article ">Figure 3
<p>Normalized mRNA expression in undifferentiated MO3.13 cells (uMO3) and MO3.13 cells following differentiation (dMO3) with (<b>A</b>) PMA, (<b>B</b>) N2.1 medium, (<b>C</b>) N2.2 medium or (<b>D</b>) T3 medium. OL: oligodendrocyte lineage, OPC: oligodendrocyte precursor, imOL: immature oligodendrocyte, mOL: mature oligodendrocyte, myOL; myelinating oligodendrocyte, NEU: neuronal, NP: neural progenitor, AST: astrocyte, NP: neural progenitor, iSC: immature Schwann cell, mSC: mature Schwann cell. Independent samples <span class="html-italic">t</span>-tests are based on triplicates in three independent experiments (<span class="html-italic">n</span> = 3). <span class="html-italic">p</span>-values: * &lt; 0.05, ** &lt; 0.01, *** &lt; 0.001.</p>
Full article ">Figure 4
<p>Analysis of protein expression in undifferentiated and differentiated MO3.13 cells. (<b>A</b>) Western blot analyses of MBP (polyclonal), MOG, CNPase, MBP (monoclonal), TUBB3, PLP1 and PON2 in (+) human motor cortex, (1) undifferentiated MO3.13 cells, (2) differentiated MO3.13 cells following incubation in N2.1 medium, (3) N2.2 medium or (4) T3 medium (normalized to GAPDH expression). Western blot analyses were performed for two independent experiments (<span class="html-italic">n</span> = 2). (<b>B</b>) Example images of immunocytochemistry for CNPase and TUBB3 in undifferentiated MO3.13 cells and MO3.13 cells differentiated with N2.1 medium or N2.2 medium. Scale bar = 50 µm.</p>
Full article ">Figure 5
<p>Normalized mRNA expression of PLP1, CNP, SOX10, vascular endothelial growth factor A (VEGF-A) and VEGF-C at various cell densities in (<b>A</b>) HOG cells and (<b>B</b>) MO3.13 cells. Standard error of the mean (SEM) is based on two technical replicates of one experiment (<span class="html-italic">n</span> = 1).</p>
Full article ">Figure 6
<p>Immunocytochemical analysis of neuronal differentiation in co-cultures of HOG or MO3.13 cells with differentiated neuronal SH-SY5Y cells. (<b>A</b>) Representative images of for expression of mature neuronal markers TUBB3 and tyrosine hydroxylase (TH) or dopamine beta-hydroxylase (DBH) in differentiated SH-SY5Y monocultures. (<b>B</b>) Representative images of Jagged1 and L1CAM surface expression and TUBB3 signals in SH-SY5Y monoculture, and SH-SY5Y + HOG and SH-SY5Y + MO3.13 co-cultures. Scale bar = 50 µm. (<b>C</b>) Quantifications of neurite caliber in low density neurite networks of differentiated SH-SY5Y cells (<span class="html-italic">n</span> = 2).</p>
Full article ">Figure 7
<p>Immunocytochemical analysis of co-cultures of HOG or MO3.13 cells with differentiated neuronal SH-SY5Y cells. (<b>A</b>) Representative image of neuronal marker neurofilament light chain (NEFL) and TUBB3 stainings in SH-SY5Y + HOG co-culture and FluoroMyelin Red (FLMred) cytoplasmic lipid droplets in living SH-SY5Y + MO3.13 co-cultures. (<b>B</b>) Representative images of TUBB3- and FLMred signals in fixed SH-SY5Y monoculture, and SH-SY5Y + HOG co-culture and SH-SY5Y + MO3.13 co-culture. These images serve as examples for the absence of FLMred overlap with the extensive neurite outgrowth (TUBB3) in our cultures. (<b>C</b>) Representative images of the absence of MBP-signals (polyclonal antibody) and the presence of TUBB3-signals in SH-SY5Y monocultures, and SH-SY5Y + HOG and SH-SY5Y + MO3.13 co-cultures. Scale bar = 50 µm.</p>
Full article ">
21 pages, 2515 KiB  
Article
Fatty Acid-Treated Induced Pluripotent Stem Cell-Derived Human Cardiomyocytes Exhibit Adult Cardiomyocyte-Like Energy Metabolism Phenotypes
by Yuichi Horikoshi, Yasheng Yan, Maia Terashvili, Clive Wells, Hisako Horikoshi, Satoshi Fujita, Zeljko J. Bosnjak and Xiaowen Bai
Cells 2019, 8(9), 1095; https://doi.org/10.3390/cells8091095 - 17 Sep 2019
Cited by 105 | Viewed by 12974
Abstract
Human induced pluripotent stem cell (iPSC)-derived cardiomyocytes (CMs) (iPSC-CMs) are a promising cell source for myocardial regeneration, disease modeling and drug assessment. However, iPSC-CMs exhibit immature fetal CM-like characteristics that are different from adult CMs in several aspects, including cellular structure and metabolism. [...] Read more.
Human induced pluripotent stem cell (iPSC)-derived cardiomyocytes (CMs) (iPSC-CMs) are a promising cell source for myocardial regeneration, disease modeling and drug assessment. However, iPSC-CMs exhibit immature fetal CM-like characteristics that are different from adult CMs in several aspects, including cellular structure and metabolism. As an example, glycolysis is a major energy source for immature CMs. As CMs mature, the mitochondrial oxidative capacity increases, with fatty acid β-oxidation becoming a key energy source to meet the heart’s high energy demand. The immaturity of iPSC-CMs thereby limits their applications. The aim of this study was to investigate whether the energy substrate fatty acid-treated iPSC-CMs exhibit adult CM-like metabolic properties. After 20 days of differentiation from human iPSCs, iPSC-CMs were sequentially cultured with CM purification medium (lactate+/glucose-) for 7 days and maturation medium (fatty acids+/glucose-) for 3–7 days by mimicking the adult CM’s preference of utilizing fatty acids as a major metabolic substrate. The purity and maturity of iPSC-CMs were characterized via the analysis of: (1) Expression of CM-specific markers (e.g., troponin T, and sodium and potassium channels) using RT-qPCR, Western blot or immunofluorescence staining and electron microscopy imaging; and (2) cell energy metabolic profiles using the XF96 Extracellular Flux Analyzer. iPSCs-CMs (98% purity) cultured in maturation medium exhibited enhanced elongation, increased mitochondrial numbers with more aligned Z-lines, and increased expression of matured CM-related genes, suggesting that fatty acid-contained medium promotes iPSC-CMs to undergo maturation. In addition, the oxygen consumption rate (OCR) linked to basal respiration, ATP production, and maximal respiration and spare respiratory capacity (representing mitochondrial function) was increased in matured iPSC-CMs. Mature iPSC-CMs also displayed a larger change in basal and maximum respirations due to the utilization of exogenous fatty acids (palmitate) compared with non-matured control iPSC-CMs. Etomoxir (a carnitine palmitoyltransferase 1 inhibitor) but not 2-deoxyglucose (an inhibitor of glycolysis) abolished the palmitate pretreatment-mediated OCR increases in mature iPSC-CMs. Collectively, our data demonstrate for the first time that fatty acid treatment promotes metabolic maturation of iPSC-CMs (as evidenced by enhanced mitochondrial oxidative function and strong capacity of utilizing fatty acids as energy source). These matured iPSC-CMs might be a promising human CM source for broad biomedical application. Full article
(This article belongs to the Special Issue Stem Cell-based Therapy and Disease Modeling)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Characterization of human induced pluripotent stem cells (iPSCs) and iPSC-derived cardiomyocytes (iPSC-CMs). (<b>A</b>) Schematic depicting the procedure for the generation of cardiomyocytes from iPSCs by temporal modulation of Wnt signaling, purification, and maturation of iPSC-CMs. Note: mTeSR1 and Roswell Park Memorial Institute <span class="html-italic">(RMPI):</span> cell culture medium; B27: culture medium supplement; CHIR-99021: highly selective inhibitor of glycogen synthase kinase 3 (GSK-3); and IWP4: inhibitor of Wnt/β-catenin signaling. (<b>B</b>) Characterization of cultured 1013 iPSCs. Phase contrast image shows that iPSCs grow as colonies (a). Confocal fluorescent images indicate that iPSCs express pluripotent stem cell-specific markers octamer-binding transcription factor (OCT4) (red) (b), and stage-specific embryonic antigen-4 (SSEA4, red) (c). Blue are cell nuclei stained with Hoechst 33342. Scale bar = 50 μm. (<b>C</b>) Characterization of the differentiated cardiomyocytes (1013 iPSC-derived CMs). iPSC-CMs (day 20) grew as a monolayer (a) and expressed cardiomyocyte-specific markers troponin T (green) (b) and sarcomeric α-actinin (red) (c). Blue are cell nuclei. Scale bar = 30 µm.</p>
Full article ">Figure 2
<p>Lactate purification of 1013 iPSC-derived CMs. (<b>A</b>) The fluorescent images of iPSC-CMs (day 31) with or without treatment of lactate-contained purification medium (no glucose) for 7 days to eliminate non-cardiomyocytes. Blue are cell nuclei stained with Hoechst 33342 and green are troponin T signals. In the purified cell culture, almost all cells with blue nuclei expressed troponin T. Scale bar = 50 µm. (<b>B</b>) The purification of iPSC-CMs increased from 75% to 98% after culturing in lactate medium. Data are presented as mean ± SEM, <span class="html-italic">n</span> = 4 * <span class="html-italic">p</span> &lt; 0.05 vs. control medium.</p>
Full article ">Figure 3
<p>The effect of fatty acid-contained cardiomyocyte maturation medium (no glucose) on the maturation of 1013 iPSC-derived CMs. (<b>A</b>) Representative immunofluorescent images of iPSC-CMs (day 34) cultured with control culture medium (a) and maturation medium for 7 days (b). A-c and A-d are the magnified images marked by yellow rectangles in A-a and A-b, respectively. Scale bar = 20 µm. (<b>B</b>) Analysis of cell area (a), perimeter (b), circularity (c), and elongation (d) of iPSC-CMs using ImageJ software. <span class="html-italic">n</span> = 50–64 * <span class="html-italic">p</span> &lt; 0.05 vs. control medium. (<b>C</b>) Representative electron microscopy images of iPSC-CMs (day 38) treated with control medium (a) and maturation medium (b) for 7 days. Scale bar = 500 nm. Maturation medium-treated iPSC-CMs showed well-aligned visible Z-lines and increased mitochondria within the cells compared with the cells cultured in control medium. Z-lines and mitochondrial are indicated by green and red arrows, respectively. (<b>D</b>) Real time PCR analysis of the effect of maturation medium on the cardiomyocyte maturation-related gene expression. <span class="html-italic">n</span> = 4 * <span class="html-italic">p</span> &lt; 0.05 vs. control medium. (iPSC-CMs day 38) (<b>E</b>) Western blot analysis of the effect of maturation medium on MYL2 (a) and MYH7 (b) protein expression in iPSC-CMs (day 38). <span class="html-italic">n</span> = 4, ** <span class="html-italic">p</span> &lt; 0.01, vs. control medium. Note: TNNI: troponin I; TNNT: troponin T; MYL: myosin light chain; MYH: myosin heavy chain; MYBPC: myosin binding protein C; SCN5A: sodium voltage-gated alpha subunit 5; KCNJ: potassium inwardly rectifying channel subunit J; KCNH: potassium voltage-gated channel subfamily H; KCNQ: potassium voltage-gated channel subfamily Q; KCND: potassium voltage-gated channel, subfamily D; CACNA1C: calcium voltage-gated channel subunit alpha1 C; GJA1: gap junction protein, alpha 1; SERCA2A: sarcoplasmic/endoplasmic reticulum Ca<sup>2+</sup>ATPase 2a; RYR2: cardiac ryanodine receptor; PPARα: peroxisome proliferator-activator receptor alpha.</p>
Full article ">Figure 4
<p>The effect of maturation medium on mitochondrial respiratory capacity of 1013 iPSC-derived CMs. (<b>A</b>) The diagram depicts the trace of oxygen consumption rate (OCR) on cells after sequentially administration of 10 µM oligomycin (ATP synthase inhibitor), 2 µM FCCP (uncoupler of oxidative phosphorylation in mitochondria), and 10 µM antimycin A (electron transport chain blocker) to the culture. The profiles of fundamental parameters of mitochondrial function measured are basal respiration, ATP production, maximal respiration and spare respiratory capacity that were marked with different color in the schematic. (<b>B</b>) Representative two OCR traces of iPSC-CMs (day 34) treated with either control medium or maturation medium for 3 days, respectively, in response to oligomycin, FCCP, and antimycin A. (<b>C</b>) OCR parameters representing mitochondrial function in maturation medium-treated iPSC-CMs were significantly higher. <span class="html-italic">n</span> = 4, * <span class="html-italic">p</span> &lt; 0.05 vs. control medium.</p>
Full article ">Figure 5
<p>The effect of exogenous fatty acids (palmitate) pre-treatment on the OCR of 1013 iPSC-derived CMs (day 34) cultured with either control medium or maturation medium. (<b>A</b>) The changed basal and maximum OCR in both control and matured iPSC-CMs in response to exogenous palmitate. iPSC-CMs were pretreated with bovine serum albumin (BSA; as control), palmitate, or etomoxir (ETO, a specific irreversible inhibitor of carnitine palmitoyltransferase 1 inhibitor). Basal and maximum respiration capacity was significantly increased in the matured iPSC-CMs compared to control cells due to the utilization of palmitate. Pretreatment with palmitate-BSA (blue) resulted in the increased basal and maximal respiration in non-matured iPSC-CMs compared with BSA alone treatment group (black) (a). Matured iPSC-CMs exhibit the greater increase in both basal respiration (4.4 fold higher vs. non-matured iPSC-CMs) and maximal respiration (2.5 fold higher vs. non-matured iPSC-CMs) in response to palmitate-BSA (blue) (b and c), indicating the stronger ability to oxidize exogenously added palmitate in the matured iPSC-CMs. <span class="html-italic">n</span> = 4 * <span class="html-italic">p</span> &lt; 0.05. (<b>B</b>) The effect of ETO on the OCR of matured iPSC-CMs pre-treated with BSA or palmitate. Representative OCR traces of iPSC-CMs pre-treated with BSA or palmitate together with or without ETO in response to oligomycin, FCCP, and antimycin A (a). ETO completely inhibited the palmitate pretreatment-mediated increases of mitochondrial respiration in matured iPSC-CMs (b). <span class="html-italic">n</span> = 4 * <span class="html-italic">p</span> &lt; 0.05. (<b>C</b>) The effect of 2-deoxyglucose (2-DG) on the matured iPSC-CMs that were pre-treated with BSA or palmitate. Representative OCR traces of iPSC-CMs pre-treated with BSA or palmitate together with or without 2-DG (glucose analogue, a competitive glycolytic inhibitor) in response to oligomycin, FCCP, and antimycin A (a). 2-DG did not attenuate the palmitate-mediated increase of OCR in matured iPSC-CMs (b), suggesting that the palmitate-mediated increased OCR in palmitate-pretreated iPSC-CMs is resulted from the fatty acid effect. <span class="html-italic">n</span> = 4 * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 6
<p>The effect of maturation medium on glycolytic function of iPSC-CMs (day 34) via analysis of the extracellular acidification rate (ECAR) of 1013 iPSC-derived CMs. (<b>A</b>-a) This diagram depicts (1) the representative traces of ECAR on the cells after sequentially adding 5.5 µM glucose, 10 µM oligomycin, and 25 mM 2-DG to the culture and (2) the profiles of key parameters of glycolytic function are glycolysis (blue), glycolytic capacity (green), glycolytic reserve (yellow), and non-glycolytic acidification (red). (<b>A</b>-b) Representative two ECAR traces of iPSC-CMs treated with control and maturation medium for 3 days, respectively. (<b>A</b>-c) The quantified ECAR shows that maturation medium increased glycolysis, glycolytic capacity, and glycolytic reserve of iPSC-CMs, suggesting that matured iPSC-CMs have a higher capacity to switch to glucose utilization when fatty acid oxidation is compromised. <span class="html-italic">n</span> = 4, * <span class="html-italic">p</span> &lt; 0.05 vs. control medium. (<b>B</b>) The effects of different concentrations of glucose on glycolytic function of the control medium-treated iPSC-CMs. There is no significant difference in glycolytic function for each glucose concentration. (<b>C</b>-a) Representative traces of ECAR of matured iPSC-CMs in response to various glucose concentrations. (<b>C</b>-b) The glycolysis and glycolytic capacity exhibited the response to the glucose in a dose-dependent manner. <span class="html-italic">n</span> = 4, * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 7
<p>The effect of the maturation medium on protein expression, and metabolism of H3083 iPSC-derived CMs (<b>A</b>) Western blot analysis of the effect of maturation medium treatment for 7 days on the expression of MYL2 and MYH7 proteins in iPSC-CMs (day 38). <span class="html-italic">n</span> = 4 * <span class="html-italic">p</span> &lt; 0.05, vs. control medium. (<b>B</b>) The effect of maturation medium on OCR and ECAR of iPSC-CMs (day 34). * <span class="html-italic">p</span> &lt; 0.05 vs. control medium.</p>
Full article ">
18 pages, 3844 KiB  
Article
Maternal Protein Restriction Modulates Angiogenesis and AQP9 Expression Leading to a Delay in Postnatal Epididymal Development in Rat
by Talita de Mello Santos, Marilia Martins Cavariani, Dhrielly Natália Pereira, Bruno César Schimming, Luiz Gustavo de Almeida Chuffa and Raquel Fantin Domeniconi
Cells 2019, 8(9), 1094; https://doi.org/10.3390/cells8091094 - 17 Sep 2019
Cited by 5 | Viewed by 3704
Abstract
The maternal nutritional status is essential to the health and well-being of the fetus. Maternal protein restriction during the perinatal stage causes sperm alterations in the offspring that are associated with epididymal dysfunctions. Vascular endothelial growth factor (VEGF) and its receptor, VEGFr-2, as [...] Read more.
The maternal nutritional status is essential to the health and well-being of the fetus. Maternal protein restriction during the perinatal stage causes sperm alterations in the offspring that are associated with epididymal dysfunctions. Vascular endothelial growth factor (VEGF) and its receptor, VEGFr-2, as well as aquaporins (AQPs) are important regulators of angiogenesis and the epididymal microenvironment and are associated with male fertility. We investigated the effects of maternal protein restriction on epididymal angiogenesis and AQP expression in the early stages of postnatal epididymal development. Pregnant rats were divided into two experimental groups that received either a normoprotein (17% protein) or low-protein diet (6% protein) during gestation and lactation. At postnatal day (PND)7 and PND14, male offspring were euthanized, the epididymides were subjected to morphometric and microvascular density analyses and to VEGF-A, VEGF-r2, AQP1 and AQP9 expression analyses. The maternal low-protein diet decreased AQP9 and VEGFr-2 expression, decreased epididymal microvascularity and altered the morphometric features of the epididymal epithelium; no changes in AQP1 expression were observed at the beginning of postnatal epididymal development. Maternal protein restriction alters microvascularization and affects molecules involved in the epidydimal microenvironment, resulting in morphometric alterations related to a delay in the beginning of epididymis postnatal development. Full article
(This article belongs to the Special Issue Aquaporins 2019)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Illustration of the experimental design.</p>
Full article ">Figure 2
<p>Expression of AQP1, AQP9, VEGFr-2 and VEGFa in the epididymis of NP and LP animals at PND7. Data are expressed as the mean ± S.E.M. * <span class="html-italic">p</span> &lt; 0.05. Mann–Whitney test.</p>
Full article ">Figure 3
<p>Expression of AQP1, AQP9, VEGFr-2 and VEGFa in the epididymis of NP and LP animals at PND14. Data are expressed as the mean ± S.E.M. * <span class="html-italic">p</span> &lt; 0.05. Mann–Whitney test.</p>
Full article ">Figure 4
<p>Epididymis sections of the initial segment (IS), caput regions, corpus and cauda from the NP and LP animals at PND14 subjected to AQP1 immunostaining. NP = normoprotein animals; LP = low-protein animals. (<b>A</b>) (IS and Caput from NP group); (<b>B</b>) (IS and Caput from LP group); (<b>C</b>) (Corpus from NP group); (<b>D</b>) (Corpus from LP group); (<b>E</b>) (Cauda from NP group); (<b>F</b>) (Cauda from LP group). Arrow = Positive staining for AQP1 in the vascular endothelium.</p>
Full article ">Figure 5
<p>Epididymis sections of the initial segment (IS), caput regions, corpus and cauda from the NP and LP animals at PND14 subjected to AQP9 immunostaining. NP = normoprotein animals; LP = low-protein animals. (<b>A</b>) (IS and Caput from NP group); (<b>B</b>) (IS and Caput from LP group); (<b>C</b>) (Corpus from NP group); (<b>D</b>) (Corpus from LP group); (<b>E</b>) (Cauda from NP group); (<b>F</b>) (Cauda from LP group). Arrow = Positive staining for AQP9.</p>
Full article ">Figure 6
<p>The microvascular densities in the epididymis of the NP and LP animals at PND14. Data are expressed as the mean ± S.E.M. * <span class="html-italic">p</span> &lt; 0.05. Mann–Whitney test.</p>
Full article ">Figure 7
<p>Representative image that illustrates the impact of maternal protein restriction in the early stage of postnatal epididymal development in rat. Maternal protein restriction decreases microvascular density by decreasing the VEGFr-2 expression in the endothelial cells. With the decrease in the number of vessels, the arrival of growth factors and hormones, including androgens, also decreases. Blood vessels are present around the mesenchymal cells that surround the epithelium, and before the first wave of testicular fluid, the arrival of hormones occurs exclusively through blood supply. When differentiated, the principal cells begin to express AQP9 in the apical region. Because AQP9 is downregulated in these cells, the luminal microenvironment may not be adequately regulated. AQP9 = aquaporin 9; AQP1 = aquaporin 1; VEGFa = vascular endothelial growth factor; VEGFr-2 = vascular endothelial growth factor receptor type 2, LM: luminal microenvironment.</p>
Full article ">
19 pages, 3987 KiB  
Article
Single Cell Mass Cytometry of Non-Small Cell Lung Cancer Cells Reveals Complexity of In Vivo and Three-Dimensional Models over the Petri-Dish
by Róbert Alföldi, József Á. Balog, Nóra Faragó, Miklós Halmai, Edit Kotogány, Patrícia Neuperger, Lajos I. Nagy, Liliána Z. Fehér, Gábor J. Szebeni and László G. Puskás
Cells 2019, 8(9), 1093; https://doi.org/10.3390/cells8091093 - 16 Sep 2019
Cited by 21 | Viewed by 7420
Abstract
Single cell genomics and proteomics with the combination of innovative three-dimensional (3D) cell culture techniques can open new avenues toward the understanding of intra-tumor heterogeneity. Here, we characterize lung cancer markers using single cell mass cytometry to compare different in vitro cell culturing [...] Read more.
Single cell genomics and proteomics with the combination of innovative three-dimensional (3D) cell culture techniques can open new avenues toward the understanding of intra-tumor heterogeneity. Here, we characterize lung cancer markers using single cell mass cytometry to compare different in vitro cell culturing methods: two-dimensional (2D), carrier-free, or bead-based 3D culturing with in vivo xenografts. Proliferation, viability, and cell cycle phase distribution has been investigated. Gene expression analysis enabled the selection of markers that were overexpressed: TMEM45A, SLC16A3, CD66, SLC2A1, CA9, CD24, or repressed: EGFR either in vivo or in long-term 3D cultures. Additionally, TRA-1-60, pan-keratins, CD326, Galectin-3, and CD274, markers with known clinical significance have been investigated at single cell resolution. The described twelve markers convincingly highlighted a unique pattern reflecting intra-tumor heterogeneity of 3D samples and in vivo A549 lung cancer cells. In 3D systems CA9, CD24, and EGFR showed higher expression than in vivo. Multidimensional single cell proteome profiling revealed that 3D cultures represent a transition from 2D to in vivo conditions by intermediate marker expression of TRA-1-60, TMEM45A, pan-keratin, CD326, MCT4, Gal-3, CD66, GLUT1, and CD274. Therefore, 3D cultures of NSCLC cells bearing more putative cancer targets should be used in drug screening as the preferred technique rather than the Petri-dish. Full article
(This article belongs to the Special Issue Single Cell Analysis)
Show Figures

Figure 1

Figure 1
<p>Different culture conditions of A549 human adenocarcinoma cells for 4 days (<b>A</b>) and for 9 days (<b>B</b>). Cells were seeded at the same surface/volume ratio regarding the different culture conditions as described in Materials and Methods. Images were taken by the HoloMonitor M3 instrument using phase contrast X10 objective. Scale bar: 150 µm.</p>
Full article ">Figure 2
<p>Proliferation rate, viability and apoptosis of A549 cells under different culture conditions. (<b>A</b>) Cells grow slowly in 3D culture than in 2D cultures on culture day 4 (3D spehorids <span class="html-italic">p</span> ≤ 0.0001; RAFT, 3D Cytodex3 and 3D Nutrisphere <span class="html-italic">p</span> ≤ 0.01), but on day 9 only the 3D Spheroid (<span class="html-italic">p</span> ≤ 0.01) and 3D Cytodex3 (<span class="html-italic">p</span> ≤ 0.05) showed significantly lower cell numbers than 2D cultures. (n = 3, paired t-test). The viability of cells (AnnV−/PI− living population) remained around 80–90% at day 4 (<b>B</b>) and day 9 (<b>C</b>), only RAFT culturing resulted in 50% decrease in viability at day 9. Insert shows representative dot plots of single cells for SSC-FSC (Side scatter-Forward scatter, B left insert) and for quadrants detecting living cells (AnnV−/PI−), early apoptotic cells (AnnV+/PI-), late apoptotic cells (AnnV+/PI+) and necrotic cells (AnnV−/PI+) (B right insert). Data are mean ± SD of three replicates.</p>
Full article ">Figure 3
<p>Cell cycle phase distribution of A549 cells cultured under different conditions at the 4<sup>th</sup> (<b>A</b>) and the 9<sup>th</sup> culturing days (<b>B</b>). Insert shows PI signal of single cells FL2-A (area)-FL2-W (width) gating out aggregates (<b>A</b> left insert) and a representative image of cell cycle phases (<b>A</b> right insert). Data are mean ± SD of three replicates except RAFT where 12 collagen discoids were pooled.</p>
Full article ">Figure 4
<p>Gene expression changes after 4 (d4) and 9 days (d9) of culture under different 2D and 3D conditions. Selected genes having cell surface localized protein product and well characterized antibody available on the market: TMEM5A, SLC16A3 (MCT4), CEACAM5, SLC2A1 (GLUT1), and CA9 showed 16–32 times overexpression in 3D Cytodex3 and 3D Nutrisphere compared to 2D monolayer. Data are mean ± SD of three replicates.</p>
Full article ">Figure 5
<p>Long-term (day 9, d9) maintenance of 3D cultures (3D Spheroid, 3D Nutrisphere, 3D Cytodex3) mimic the in vivo situation better. The highlighted genes (black squares) were selected for subsequent analysis by single cell mass cytometry. Cluster analysis (hierarchical) was performed by Gene Cluster 3.0 software from expression data (ΔΔCt, log2 values) of presented samples.</p>
Full article ">Figure 6
<p>Representative multidimensional visualization of stochastic neighbor embedding (viSNE) analysis of 12 protein markers at single cell resolution in 2D, 3D (3D Cytodex3 or 3D Nutrisphere) cultures and in vivo (early or late stage) tumors. The analysis was performed within 4.5 × 10<sup>4</sup> HLA-A,B,C positive cells in case of all conditions in order to identify human A549 cells in the xenografts and exclude murine stromal cells. (iterations = 1000, perplexity = 30, theta = 0.5).</p>
Full article ">Figure 7
<p>Merging viSNE graphs of multiparametric single cell mass cytometry data (12 parameters) of 2D, 3D and in vivo samples delineates a map with three different islands of 2D, 3D, and in vivo conditions.</p>
Full article ">Figure 8
<p>(<b>A</b>) Trajectories tend to localize to early and late stage tumors in the case of TRA-1-60, TMEM45A, pan-keratin, CD326, MCT4, GAL-3, CD66, and GLUT1. Percentage of HLA-A,B,C positive cells (A549) from a representative experiment were plotted on trajectories among five different conditions: (I) 2D monolayer, (II) 3D Cytodex3, (III) 3D Nutrisphere, (IV) early and (V) late stage in vivo adenocarcinoma. (<b>B</b>) Heatmap of mass cytometry data regarding protein density at single cell resolution among the five different conditions normalized to 2D monolayer (green: low, red: high expression).</p>
Full article ">
13 pages, 6062 KiB  
Article
Expression of FGF8, FGF18, and FGFR4 in Gastroesophageal Adenocarcinomas
by Gerd Jomrich, Xenia Hudec, Felix Harpain, Daniel Winkler, Gerald Timelthaler, Thomas Mohr, Brigitte Marian and Sebastian F. Schoppmann
Cells 2019, 8(9), 1092; https://doi.org/10.3390/cells8091092 - 16 Sep 2019
Cited by 18 | Viewed by 4106
Abstract
Even though distinctive advances in the field of esophageal cancer therapy have occurred over the last few years, patients’ survival rates remain poor. FGF8, FGF18, and FGFR4 have been identified as promising biomarkers in a number of cancers; however no data exist on [...] Read more.
Even though distinctive advances in the field of esophageal cancer therapy have occurred over the last few years, patients’ survival rates remain poor. FGF8, FGF18, and FGFR4 have been identified as promising biomarkers in a number of cancers; however no data exist on expression of FGF8, FGF18, and FGFR4 in adenocarcinomas of the esophago-gastric junction (AEG). A preliminary analysis of the Cancer Genome Atlas (TCGA) database on FGF8, FGF18, and FGFR4 mRNA expression data of patients with AEG was performed. Furthermore, protein levels of FGF8, FGF18, and FGFR4 in diagnostic biopsies and post-operative specimens in neoadjuvantly treated and primarily resected patients using immunohistochemistry were investigated. A total of 242 patients was analyzed in this study: 87 patients were investigated in the TCGA data set analysis and 155 patients in the analysis of protein expression using immunohistochemistry. High protein levels of FGF8, FGF18, and FGFR4 were detected in 94 (60.7%), 49 (31.6%) and 84 (54.2%) patients, respectively. Multivariable Cox proportional hazard regression models revealed that high expression of FGF8 was an independent prognostic factor for diminished overall survival for all patients and for neoadjuvantly treated patients. By contrast, FGF18 overexpression was significantly associated with longer survival rates in neoadjuvantly treated patients. In addition, FGF8 protein level correlated with Mandard regression due to neoadjuvant therapy, indicating potential as a predictive marker. In summary, FGF8 and FGF18 are promising candidates for prognostic factors in adenocarcinomas of the esophago-gastric junction and new potential targets for new anti-cancer therapies. Full article
(This article belongs to the Special Issue Fibroblast Growth Factor Receptor (FGFR) Signaling Pathway in Tumor)
Show Figures

Figure 1

Figure 1
<p>Kaplan-Meier curves of overall survival of patients with adenocarcinomas of the esophago-gastric junction. (<b>a</b>–<b>c</b>) Patients from TCGA data set analysis: high FGF8, FGF18, and FGFR4 expression compared with those with low/absent FGF8, FGF18, and FGFR4 expression. (<b>d</b>–<b>f</b>) Patients from the immunohistological analysis: high FGF8, FGF18, and FGFR4 expression compared with those with low/absent FGF8, FGF18, and FGFR4 expression.</p>
Full article ">Figure 2
<p>Specimen of adenocarcinomas of the esophago-gastric junction stained for (<b>a</b>) FGF8, (<b>b</b>) FGF18 and (<b>c</b>) FGFR4. Positive staining was found in the tumor cells and to a lesser degree in the microenvironment. FGF8 and FGFR4 expression were primarily found in the nucleus, while FGF18 expression was mainly found in the cytoplasm. For quantitative evaluation, only epithelial cells were investigated. Corresponding sections stained by CK7 (<b>d</b>), Ki67 (<b>e</b>), and negative control (<b>f</b>). (The bar corresponds to 50 µm.) Original magnification ×400 all).</p>
Full article ">Figure 3
<p>Representative high (<b>a</b>–<b>c</b>) and low (<b>d</b>–<b>f</b>) expressing tumor section of FGF8 (<b>a</b> and <b>d</b>), FGF18 (<b>b</b> and <b>e</b>), and FGFR4 (<b>c</b> and <b>f</b>).</p>
Full article ">
14 pages, 1781 KiB  
Article
Expression of FGFR1–4 in Malignant Pleural Mesothelioma Tissue and Corresponding Cell Lines and its Relationship to Patient Survival and FGFR Inhibitor Sensitivity
by Gregor Vlacic, Mir A. Hoda, Thomas Klikovits, Katharina Sinn, Elisabeth Gschwandtner, Katja Mohorcic, Karin Schelch, Christine Pirker, Barbara Peter-Vörösmarty, Jelena Brankovic, Balazs Dome, Viktoria Laszlo, Tanja Cufer, Ales Rozman, Walter Klepetko, Bettina Grasl-Kraupp, Balazs Hegedus, Walter Berger, Izidor Kern and Michael Grusch
Cells 2019, 8(9), 1091; https://doi.org/10.3390/cells8091091 - 16 Sep 2019
Cited by 10 | Viewed by 4009
Abstract
Malignant pleural mesothelioma (MPM) is a devastating malignancy with limited therapeutic options. Fibroblast growth factor receptors (FGFR) and their ligands were shown to contribute to MPM aggressiveness and it was suggested that subgroups of MPM patients could benefit from FGFR-targeted inhibitors. In the [...] Read more.
Malignant pleural mesothelioma (MPM) is a devastating malignancy with limited therapeutic options. Fibroblast growth factor receptors (FGFR) and their ligands were shown to contribute to MPM aggressiveness and it was suggested that subgroups of MPM patients could benefit from FGFR-targeted inhibitors. In the current investigation, we determined the expression of all four FGFRs (FGFR1–FGFR4) by immunohistochemistry in tissue samples from 94 MPM patients. From 13 of these patients, we were able to establish stable cell lines, which were subjected to FGFR1–4 staining, transcript analysis by quantitative RT-PCR, and treatment with the FGFR inhibitor infigratinib. While FGFR1 and FGFR2 were widely expressed in MPM tissue and cell lines, FGFR3 and FGFR4 showed more restricted expression. FGFR1 and FGFR2 showed no correlation with clinicopathologic data or patient survival, but presence of FGFR3 in 42% and of FGFR4 in 7% of patients correlated with shorter overall survival. Immunostaining in cell lines was more homogenous than in the corresponding tissue samples. Neither transcript nor protein expression of FGFR1–4 correlated with response to infigratinib treatment in MPM cell lines. We conclude that FGFR3 and FGFR4, but not FGFR1 or FGFR2, have prognostic significance in MPM and that FGFR expression is not sufficient to predict FGFR inhibitor response in MPM cell lines. Full article
(This article belongs to the Special Issue Fibroblast Growth Factor Receptor (FGFR) Signaling Pathway in Tumor)
Show Figures

Figure 1

Figure 1
<p>Expression of FGFR1–4 in MPM tissue. (<b>A</b>) Representative images of MPM tissue specimens stained for FGFR1 (score 2, epithelioid), FGFR2 (score 2, epithelioid), FGFR3 (score 1, epithelioid), and FGFR4 (score 1, epithelioid). Scale bar = 25 µm. (<b>B</b>) Distribution of staining intensities of FGFR1–4 in 94 MPM tissue specimens.</p>
Full article ">Figure 2
<p>Correlation of FGFR expression with histology and patient prognosis. (<b>A</b>) Percentage of epithelioid, biphasic, and sarcomatoid tumors within the different staining groups of FGFR1 (upper panel), FGFR3 (middle panel), and FGFR4 (lower panel). (<b>B</b>) Kaplan–Meier curves for overall survival of MPM patients with different staining scores of FGFR1 (upper panels), FGFR3 (middle panel), and FGFR4 (lower panel).</p>
Full article ">Figure 3
<p>Sensitivity of patient-derived cell lines to the FGFR inhibitor infigratinib. (<b>A</b>) MPM cell lines were incubated with increasing concentrations of infigratinib or vehicle (DMSO) as control and cell number was determined after 72 h. Dose–response curves were calculated with GraphPad Prism. Three sensitive cell lines (IC<sub>50</sub> &lt; 1 µM), one intermediate cell line (1 µM &lt; IC<sub>50</sub> &lt; 10 µM), and three resistant cell lines (IC<sub>50</sub> &gt; 10 µM) are shown in green, black, and red, respectively. (<b>B</b>) Infigratinib IC<sub>50</sub> values of the cell lines were plotted against the IHC scores for FGFR1, FGFR3, or the sum of FGFR1–3 of the corresponding tumors from which the cell lines were established.</p>
Full article ">Figure 4
<p>Expression of FGFR1–4 in patient-derived cell lines. (<b>A</b>) Representative images from cell blocks stained for FGFR1 (Meso208), FGFR2 (Meso161), FGFR3 (VMC45), and FGFR4 (Meso205). Scale bar = 100 µm in overview images and 10 µm in high magnification insets. (<b>B</b>) Total mRNA was isolated from logarithmically growing cell lines and subjected to cDNA synthesis. Quantitative RT-PCR was performed by Taqman assays for FGFR1–4. Expression levels were plotted as 2<sup>−dCt</sup> × 10<sup>5</sup> normalized to the house-keeping genes GAPDH and beta-actin. (<b>C</b>) Infigratinib IC<sub>50</sub> values were plotted as function of FGFR1 mRNA expression level. (<b>D</b>) BLU9931 IC<sub>50</sub> values were plotted as function of FGFR4 mRNA expression levels.</p>
Full article ">
20 pages, 886 KiB  
Review
Therapeutic Approaches Targeting Nucleolus in Cancer
by Pietro Carotenuto, Annalisa Pecoraro, Gaetano Palma, Giulia Russo and Annapina Russo
Cells 2019, 8(9), 1090; https://doi.org/10.3390/cells8091090 - 16 Sep 2019
Cited by 60 | Viewed by 6007
Abstract
The nucleolus is a distinct sub-cellular compartment structure in the nucleus. First observed more than 200 years ago, the nucleolus is detectable by microscopy in eukaryotic cells and visible during the interphase as a sub-nuclear structure immersed in the nucleoplasm, from which it [...] Read more.
The nucleolus is a distinct sub-cellular compartment structure in the nucleus. First observed more than 200 years ago, the nucleolus is detectable by microscopy in eukaryotic cells and visible during the interphase as a sub-nuclear structure immersed in the nucleoplasm, from which it is not separated from any membrane. A huge number of studies, spanning over a century, have identified ribosome biogenesis as the main function of the nucleolus. Recently, novel functions, independent from ribosome biogenesis, have been proposed by several proteomic, genomic, and functional studies. Several works have confirmed the non-canonical role for nucleoli in regulating important cellular processes including genome stability, cell-cycle control, the cellular senescence, stress responses, and biogenesis of ribonucleoprotein particles (RNPs). Many authors have shown that both canonical and non-canonical functions of the nucleolus are associated with several cancer-related processes. The association between the nucleolus and cancer, first proposed by cytological and histopathological studies showing that the number and shape of nucleoli are commonly altered in almost any type of cancer, has been confirmed at the molecular level by several authors who demonstrated that numerous mechanisms occurring in the nucleolus are altered in tumors. Recently, therapeutic approaches targeting the nucleolus in cancer have started to be considered as an emerging “hallmark” of cancer and several therapeutic interventions have been developed. This review proposes an up-to-date overview of available strategies targeting the nucleolus, focusing on novel targeted therapeutic approaches. Finally, a target-based classification of currently available treatment will be proposed. Full article
(This article belongs to the Special Issue Nucleolar Organization and Functions in Health and Disease)
Show Figures

Figure 1

Figure 1
<p>The nucleolar stress pathway. Nucleolus can be exposed to a variety of cellular stressors that disrupt ribosome biogenesis activating a complex cellular response namely “nucleolar stress”. This stress pathway is mediated by several ribosomal proteins RPs and/or nucleolar proteins and its activation results in cell cycle arrest, apoptosis, DNA damage and senescence. Dysregulation of this response is known to contribute to the development of cancer.</p>
Full article ">Figure 2
<p>Targeting Nucleolar Function in Cancer Therapeutics. Schematic representation of anti-cancer drugs targeting nucleolar structures and/or functions.</p>
Full article ">
27 pages, 5139 KiB  
Article
Fisetin, a 3,7,3′,4′-Tetrahydroxyflavone Inhibits the PI3K/Akt/mTOR and MAPK Pathways and Ameliorates Psoriasis Pathology in 2D and 3D Organotypic Human Inflammatory Skin Models
by Jean Christopher Chamcheu, Stephane Esnault, Vaqar M. Adhami, Andrea L. Noll, Sergette Banang-Mbeumi, Tithi Roy, Sitanshu S. Singh, Shile Huang, Konstantin G. Kousoulas and Hasan Mukhtar
Cells 2019, 8(9), 1089; https://doi.org/10.3390/cells8091089 - 15 Sep 2019
Cited by 59 | Viewed by 9179
Abstract
Psoriasis is a chronic immune-mediated skin disease that involves the interaction of immune and skin cells, and is characterized by cytokine-driven epidermal hyperplasia, deviant differentiation, inflammation, and angiogenesis. Because the available treatments for psoriasis have significant limitations, dietary products are potential natural sources [...] Read more.
Psoriasis is a chronic immune-mediated skin disease that involves the interaction of immune and skin cells, and is characterized by cytokine-driven epidermal hyperplasia, deviant differentiation, inflammation, and angiogenesis. Because the available treatments for psoriasis have significant limitations, dietary products are potential natural sources of therapeutic molecules, which can repair the molecular defects associated with psoriasis and could possibly be developed for its management. Fisetin (3,7,3′,4′-tetrahydroxyflavone), a phytochemical naturally found in pigmented fruits and vegetables, has demonstrated proapoptotic and antioxidant effects in several malignancies. This study utilized biochemical, cellular, pharmacological, and tissue engineering tools to characterize the effects of fisetin on normal human epidermal keratinocytes (NHEKs), peripheral blood mononuclear cells (PBMC), and CD4+ T lymphocytes in 2D and 3D psoriasis-like disease models. Fisetin treatment of NHEKs dose- and time-dependently induced differentiation and inhibited interleukin-22-induced proliferation, as well as activation of the PI3K/Akt/mTOR pathway. Fisetin treatment of TNF-α stimulated NHEKs also significantly inhibited the activation of p38 and JNK, but had enhanced effect on ERK1/2 (MAPK). In addition, fisetin treatment significantly decreased the secretion of Th1/Th-17 pro-inflammatory cytokines, particularly IFN-γ and IL-17A by 12-O-tetradecanolylphorbol 13-acetate (TPA)-stimulated NHEKs and anti-CD3/CD28-activated human PBMCs. Furthermore, we established the in vivo relevance of fisetin functions, using a 3D full-thickness human skin model of psoriasis (FTRHSP) that closely mimics in vivo human psoriatic skin lesions. Herein, fisetin significantly ameliorated psoriasis-like disease features, and decreased the production of IL-17 by CD4+ T lymphocytes co-cultured with FTRHSP. Collectively, our data identify the prodifferentiative, antiproliferative, and anti-inflammatory effects of fisetin, via modulation of the PI3K-Akt-mTOR and p38/JNK pathways and the production of cytokines in 2D and 3D human skin models of psoriasis. These results suggest that fisetin has a great potential to be developed as an effective and inexpensive agent for the treatment of psoriasis and other related inflammatory skin disorders. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Fisetin at low doses (&lt; 20 µM) does not significantly affect primary normal human epidermal keratinocyte (NHEK) growth and viability and does not induce apoptosis. (<b>A</b>/<b>B</b>) Relative number of viable NHEK, immortalized keratinocytes (HaCaT), and A431 cancer cells treated with or without fisetin (1–80 µM) for 24 h (<b>A</b>) and 48 h as determined by MTT assay. Mean of percentage viability of NHEK, HaCaT, and A431 cell lines plotted against the indicated doses of fisetin. Experiments were performed three times with each concentration done in octuplicate wells, and the IC<sub>50</sub> values calculated from these plots are shown. (<b>C</b>) Effects of fisetin on the percentage of cells population in the different phases of cell cycle, and indication of late apoptosis or necrosis (i.e., PI-positive) cells were only seen in cells treated with higher fisetin concentration. (<b>D</b>) Levels or percentahes of cell cycle distribution in fisetin-treated cells as assessed by flow cytometry analysis. Bar graphs represent mean ± SD of results from three independent experiments performed in triplicate. Statistical significance was determined using one-way ANOVA and Dunn’s multiple comparison test and significance was considered when <span class="html-italic">p</span> &lt; 0.05 (*), as compared with the control. (<b>E</b>) and (<b>F</b>) Effect of different concentrations of fisetin on the expression of markers of apoptosis including caspase-3, -8, and -9, PARP (85 kDa and 116 kDa) and Bcl-2 family of proteins (Bcl2, Bax, and Bak) on cells harvested after 48 h of treatment as analyzed by Western blotting. Equal protein loading was confirmed using β-actin as loading control. (<b>F</b>) Numerical data above the blots represent relative quantitative density values for the blots normalized with an internal loading control. The Western blot data shown are representative immunoblots of two to three independent experiments with similar results.</p>
Full article ">Figure 2
<p>Fisetin treatment increased the expression of markers of epidermal keratinocytes’ differentiation and AP-1 transcription factor in NHEK. (<b>A</b>,<b>B</b>) The bar graphs represent dose-dependent induction of the differentiation marker transglutaminase (TGase) (a terminal differentiation marker) activity in NHEK treated with or without different doses of fisetin for 24 h, and the linear plot of the hydroxamate analysis of TGase activity. Data represent means ± SEM of three independent experiments, each performed in quadruplicate. Significance was assessed for control cells versus fisetin-treated cells or positive control, determined by one-way analysis of variance, as denoted by * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, and **** <span class="html-italic">p</span> &lt; 0.0001. Dose- and time-dependent increases in the protein expression of (<b>C</b>) early and late differentiation markers, and (<b>D</b>,<b>E</b>) nuclear protein expression of members of AP-1 factors subunits including Jun (C-Jun, Jun B, Jun D) and Fos (C-Fos, Fra-2, Fos-B) in NHEKs when compared to control cultures after 48 h, as analyzed by Western blotting. Panels (<b>C</b>–<b>E</b>) are representative of 2–3 experiments with similar results. Equal protein loading was confirmed using β-actin and lamin as loading controls, and the numerical data above the blots represent the quantitative relative densitometry values for the blots normalized to their respective loading controls.</p>
Full article ">Figure 3
<p>Fisetin significantly suppresses the IL-22-induced activation of PI3K/Akt/mTOR signaling in NHEK. Western blot determined protein expression levels of (<b>A</b>,<b>B</b>) PI3-K (p110 and p85), and (<b>C</b>,<b>D</b>) phosphorylation of Akt (at Ser473 and Thr308), and (<b>E</b>) phosphorylation of mTOR (at Ser2448 and Ser2481), and (<b>F</b>) phosphorylation of p-p70S6K (Thr389), as compared with untreated cells (control) and IL-22-treated cells only. (<b>B</b>,<b>D</b>,<b>G</b>) graphs of the relative intensity of normalized protein components, where each bar depict means +/-SD of three different experiments. **, *** and **** indicate <span class="html-italic">p</span> &lt; 0.01, <span class="html-italic">p</span> &lt; 0.001, and <span class="html-italic">p</span> &lt; 0.0001 vs. control for IL-22-treated only or vs. IL-22 for fisetin-treated cells.</p>
Full article ">Figure 4
<p>Fisetin regulates the PI3K/Akt/mTOR and MAPK signaling pathways in TNF-α-activated NHEK. NHEK cultures were pretreated for 24  h with/without 10 μM fisetin and then stimulated with 10 ng/mL TNF-α for 30 or 60 min prior to immunoblot analysis and quantification. (<b>A</b>) Western blot bands showing the protein expression levels of; p-p38, p-JNK, PI3K (p110a and p85), p-p90RSK, p-Akt(Ser473), p-ERK(p44/42) and (phospho-S6). Proteins were quantitated using the Bio-Rad Image Lab v6 software compared with untreated and TNF-α-activated control cells and normalized to Rab11 or vinculin as loading controls. (<b>B</b>,<b>C</b>) Quantitative analysis of normalized target protein expressions, and the plotted values are mean ± SD of each dataset from three independent experiments performed in triplicates. Significance of comparisons were made for TNF-α-stimulated control cells versus fisetin-treated at the 30 and 60 min time points, corresponding to solid lines, and also between unstimulated control cells versus TNF-α-stimulated cells indicated by the broken lines, as denoted by <span class="html-italic">* p</span> &lt; 0.05, <span class="html-italic">** p</span> &lt; 0.01, <span class="html-italic">*** p</span> &lt; 0.001, and <span class="html-italic">**** p</span> &lt; 0.0001.</p>
Full article ">Figure 5
<p>Fisetin reduces the inflammatory response in 12-O-tetradecanoyl-phorbol-13- acetate (TPA)-activated NHEK. NHEKs were pretreated with or without different concentrations of fisetin (10 and 20 µM) for 24 h, before stimulation with or without 100 ng/mL of TPA. The release of pro-inflammatory cytokines (IL-1 α, IL-1 β, IL-6, IL-8 (CXCL8), TNF-α, and TGF-α) in condition culture media was measured by Procarta-based multiplex immunoassay. Fisetin pretreatment suppressed TPA-induced secretion of these proinflammatory mediators to values similar to that of 1, 25 dihydroxyvitamin D3 (Vitamin D3) pretreatment. Values are the mean ± SD of results from three independent experiments each performed in quadruplicate. Statistical significance was determined using one-way ANOVA and Tukey’s multiple comparison test. Significance of comparisons were made for TPA-stimulated alone cells versus TPA-stimulated and fisetin/vitamin D<sub>3</sub>-treated cells, and also between unstimulated control cells versus fisetin/vitamin D3-treated cells, as denoted by <span class="html-italic">* p</span> &lt; 0.05, <span class="html-italic">** p</span> &lt; 0.01, <span class="html-italic">*** p</span> &lt; 0.001, and <span class="html-italic">**** p</span> &lt; 0.0001.</p>
Full article ">Figure 6
<p>Fisetin suppresses the expression and secretion of IFN-γ and IL-17 in activated human peripheral blood mononuclear cells (PBMC). PBMC were prepared from the circulating blood from three different healthy blood donors, and were cultured without (resting) or with anti-CD3 plus anti-CD8 (a-CD3/a-CD28), either treated with 10 µM of fisetin or the vehicle alone (-) for 10 min before activation. (<b>A</b>) PBMC were activated for 6 h before measurement of the type-1 and type-17 proinflammatory cytokines IFN-γ (<span class="html-italic">IFNG</span>) and IL-17A (<span class="html-italic">IL17A</span>) mRNA expression levels. Fold change in mRNA levels compared to resting cells, whose expression level was fixed at 1, were determined. Means ± SE are shown, and log10 transformation followed by paired <span class="html-italic">t</span>-test was used to compare mRNA levels in a-CD3/a-CD28-activated PBMC with or without fisetin. (<b>B</b>) PBMC were activated for 48 h in the indicated conditions and secreted cytokines (IFN-γ, IL-17A, and IL-4) present in the cell culture media were measured by ELISA. Means ± SD are shown and the paired <span class="html-italic">t</span>-test was used to compare values between fisetin and no fisetin (–) for a-CD3/a-CD28-activated PBMC.</p>
Full article ">Figure 7
<p>Fisetin modulated psoriasis-like features in a three-dimensional (3D) full-thickness reconstituted human skin model of psoriasis (FTRHSP). (<b>A</b>) Protocol for the establishment/generation of the 3D full-thickness reconstituted human skin model of psoriasis (FTRHSP), and for the evaluation of the therapeutic effects of compounds (fisetin or known control agent vitamin D<sub>3</sub> (Vit-D<sub>3</sub>)) on psoriasis-like pathologic features. Candidate compounds were added on top of the tissue after five days of lifting to the air‒liquid interface and at the same day in wells containing preactivated or not T lymphocytes at the base of the engineered reconstructs and co-cultured. These were continually cultivated for an additional five days in the presence or absence of these agents prior to harvest, with media replenishment every alternate day. (<b>B</b>) Photomicrographs of modified Mayers’s hematoxylin- and eosin-stained control RHSE, FTRHSP, and FTRHSP treated with/without fisetin/Vit-D<sub>3</sub> reconstructs, scale bar; top panel 20 μM, and bottom panel 50 µM. (<b>C</b>) Bar graphs showing quantification of changes in the thicknesses of; the viable epidermis (left panel), which significantly decreased in treated group vs. FTRHSP, and stratum corneum (right panel), which significantly increased in treated group vs. FTRHSP. The FTRHSP was generated and treated under different conditions (including with or without test agents fisetin or Vit-D<sub>3</sub>), and analyzed as detailed methods section. Significant differences between means ± standard deviation were determined as denoted by <span class="html-italic">* p</span> &lt; 0.05, <span class="html-italic">** p</span> &lt; 0.01, and <span class="html-italic">*** p</span> &lt; 0.001 vs. control RHSE.</p>
Full article ">Figure 8
<p>Fisetin modulates psoriasis-like features in the 3D full-thickness reconstituted human skin model of psoriasis (FTRHSP). Immunofluorescence staining was performed on FTRHSP sections by incubating with primary antibody against targets overnight at 4 °C followed by incubation with specific Alexa Flour 488-labeled secondary antibodies for 2 h at room temperature in the dark. Samples were mounted in mounting medium containing 4′,6-diamidino-2-phenylindole (DAPI) and analyzed microscopically. Immunofluorescence staining analysis of differential staining showing the protein expressions of early (K10) and late (involucrin) differentiation markers in control RHSE and FTRHSP under different treatment conditions. Stainings are shown in red and green, respectively, and DAPI in blue and results are representative images of two independent experiments each performed in quadruplicate in comparison to the control and treated FTRHSP. Scale bar = 20 μm.</p>
Full article ">Figure 9
<p>Representative immunofluorescent photomicrographs showing the protein expression levels of differentiation (filaggrin) and desmosomal (desmoglein-1) protein markers in control RHSE and FTRHSP under different treatment conditions versus fisetin or Vit-D<sub>3</sub>-treated FTRHSP tissues. Results are representative of three independent experiments each performed in quadruplicate and comparing control RHSE vs. FTRHSP treated groups. Data green (Dsgl-1); red (filaggrin), blue (DAPI) and mixed is merged representation. Dsgl-1 = desmoglein-1. Scale bar = 20 μm.</p>
Full article ">Figure 10
<p>Fisetin modulated the expression of makers of psoriasis-like inflammation and mTOR activity in 3D human skin model of psoriasis. Immunofluorescence staining for p-p70S6K and Psoriasin was performed on FTRHSP sections by incubating with primary antibody against targets overnight at 4 °C followed by incubation with specific Alexa Flour 488-labeled secondary antibodies for 2 h at room temperature in the dark. Samples were mounted in mounting medium containing 4′,6-diamidino-2-phenylindole (DAPI) and analyzed microscopically. (<b>A</b>–<b>E</b>) the antimicrobial peptide (psoriasin) and mTOR activation effector, p-p70S6K stainings are shown in red and green, respectively, and DAPI in blue. Representative pictures are shown. Scale bar = 20 μm. The yellow arrows delineate the thickness of the stratum corneum, while the white arrows indicate the thickness of the viable epidermis. (<b>F</b>,<b>G</b>) For densitometric quantification, each color image was separated into its green, blue, and red channel components using ImageJ software (National Institutes of Health, Bethesda, MD, USA), and green and red channels were used to analyze mean fluorescent intensity. Data shown here are mean fluorescence intensity ± SEM. (<b>H</b>) Differential protein expression levels of IL-17A secreted in the conditioned media of FTRHSP cultures. Pro-inflammatory cytokines IL-17A in conditioned media, were analyzed when 3D FTRHSP were treated without and with fisetin (10–30 µM) treatment or 0.1 µM Vit-D<sub>3</sub> for five days following activation by co-culturing with anti-CD3/CD28 activated CD4+ T cells for seven days as described in <a href="#sec2dot12-cells-08-01089" class="html-sec">Section 2.12</a>. Significant differences were determined and significance of comparisons were made for the expression levels of the target proteins of FTRHSP tissues versus fisetin or Vitamin D<sub>3</sub>-treated FTRHSP, corresponding to the solid lines, and also between nonactivated T cell-exposed control RHSE versus FTRHSP tissues, indicated by the broken lines, as denoted by <span class="html-italic">* p</span> &lt; 0.05, <span class="html-italic">** p</span> &lt; 0.01, and <span class="html-italic">*** p</span> &lt; 0.001.</p>
Full article ">
21 pages, 1518 KiB  
Review
ER-Mitochondria Communication in Cells of the Innate Immune System
by Dmitry Namgaladze, Vera Khodzhaeva and Bernhard Brüne
Cells 2019, 8(9), 1088; https://doi.org/10.3390/cells8091088 - 15 Sep 2019
Cited by 41 | Viewed by 11470
Abstract
In cells the interorganelle communication comprises vesicular and non-vesicular mechanisms. Non-vesicular material transfer predominantly takes place at regions of close organelle apposition termed membrane contact sites and is facilitated by a growing number of specialized proteins. Contacts of the endoplasmic reticulum (ER) and [...] Read more.
In cells the interorganelle communication comprises vesicular and non-vesicular mechanisms. Non-vesicular material transfer predominantly takes place at regions of close organelle apposition termed membrane contact sites and is facilitated by a growing number of specialized proteins. Contacts of the endoplasmic reticulum (ER) and mitochondria are now recognized to be essential for diverse biological processes such as calcium homeostasis, phospholipid biosynthesis, apoptosis, and autophagy. In addition to these universal roles, ER-mitochondria communication serves also cell type-specific functions. In this review, we summarize the current knowledge on ER-mitochondria contacts in cells of the innate immune system, especially in macrophages. We discuss ER- mitochondria communication in the context of macrophage fatty acid metabolism linked to inflammatory and ER stress responses, its roles in apoptotic cell engulfment, activation of the inflammasome, and antiviral defense. Full article
Show Figures

Figure 1

Figure 1
<p>Endoplasmic reticulum (ER)-mitochondrial regulation of lipid overload-induced stress responses. Excessive incorporation of saturated fatty acids (SFA) into ER phospholipids or accumulation of free cholesterol in the ER induce ER stress response mediated by activating transcription factor 6 (ATF6), protein kinase RNA-activated (PKR)-like ER kinase (PERK) and inositol requiring enzyme 1 (IRE1) sensor proteins. Concomitantly, SFA undergo fatty acid β-oxidation (FAO), which attenuates SFA incorporation into phosphatidylserine (PS) and phosphatidylethanolamine (PE) and dampens ER stress. Furthermore, fatty acids induce Drp1-dependent mitochondrial fragmentation, which attenuates mitochondrial ROS formation and activation of pro-inflammatory c-Jun N-terminal kinase (JNK) signaling. ER stress also provokes inositol triphosphate receptors (IP3R) activation and mitochondrial Ca<sup>2+</sup> overload. Under ER stress conditions, Ca<sup>2+</sup>/calmodulin-dependent protein kinase IIγ (CAMKIIγ) may translocate to mitochondria and undergo Ca<sup>2+</sup>- and ROS-dependent activation, promoting apoptosis. CPT: Carnitine palmitoyltransferase; PA: Phosphatidic acid; PSS: Phosphatidylserine synthase; PSD: Phosphatidylserine decarboxylase.</p>
Full article ">Figure 2
<p>ER-mitochondrial communication in NOD-like receptor protein 3 (NLRP3) inflammasome regulation. NLRP3 inflammasome activation requires two signals; a priming signal (e.g. toll-like receptor (TLR) activation) induces the expression of pro-IL-1β and NLRP3, which occurs in a mROS-dependent manner. Priming signals also induce mROS-dependent cardiolipin translocation to the outer mitochondrial membrane (OMM), where it recruits NLRP3 and caspase-1. An activation signal via e.g. P2X7 receptor ligation promotes K<sup>+</sup> efflux, mitochondrial Ca<sup>2+</sup> increase, mROS elevation and release of oxidized mitochondrial DNA (mtDNA). These events contribute to the recruitment of ASC to NLRP3 and caspase-1 and formation of the activated oligomeric inflammasome complex, which cleaves and activates caspase-1, followed by the subsequent cleavage and secretion of interleukin (IL)-1β. Mitochondria-associated membranes (MAM) provide the platform for inflammasome assembly, which may be facilitated by the interaction of NLRP3, mitochondrial antiviral-signaling protein (MAVS) and mitofusin 2 (Mfn2). Movements of mitochondria and ER, driven by a loss of NAD<sup>+</sup>, SIRT2 inactivation, and tubulin acetylation may also facilitate apposition of NLRP3 and ASC. ER stress activates the NLRP3 inflammasome through IRE1α. Mitophagy may eliminate mROS-producing mitochondria and thus, dampen NLRP3 inflammasome activation.</p>
Full article ">Figure 3
<p>Communication of ER and mitochondrial in sensing cytosolic nucleic acids. The response to RNA viruses is orchestrated by the sensor retinoic acid inducible gene I (RIG-I), which upon dsRNA recognition translocates to mitochondria- and the MAM-located adaptor MAVS to activate the interferon regulatory factor 3 (IRF3) transcription factor. MAVS may be negatively regulated by their interaction with Mfn2 and Gp78. Sensing of cytosolic DNA involves activation of the cyclic GMP-AMP (cGAMP) synthase (cGAS)-stimulator of interferon genes (STING) cascade, whereas activated STING translocates from the ER to Golgi to activate IRF3. The cross-talk between DNA and RNA sensing may involve direct interaction of STING and MAVS. cGAS may also sense mtDNA following mitochondrial damage by translocating to mitochondria.</p>
Full article ">
20 pages, 3922 KiB  
Article
The Anti-Apoptotic Effect of ASC-Exosomes in an In Vitro ALS Model and Their Proteomic Analysis
by Roberta Bonafede, Jessica Brandi, Marcello Manfredi, Ilaria Scambi, Lorenzo Schiaffino, Flavia Merigo, Ermanna Turano, Bruno Bonetti, Emilio Marengo, Daniela Cecconi and Raffaella Mariotti
Cells 2019, 8(9), 1087; https://doi.org/10.3390/cells8091087 - 14 Sep 2019
Cited by 66 | Viewed by 6521
Abstract
Stem cell therapy represents a promising approach in the treatment of several neurodegenerative disorders, including amyotrophic lateral sclerosis (ALS). The beneficial effect of stem cells is exerted by paracrine mediators, as exosomes, suggesting a possible potential use of these extracellular vesicles as non-cell [...] Read more.
Stem cell therapy represents a promising approach in the treatment of several neurodegenerative disorders, including amyotrophic lateral sclerosis (ALS). The beneficial effect of stem cells is exerted by paracrine mediators, as exosomes, suggesting a possible potential use of these extracellular vesicles as non-cell based therapy. We demonstrated that exosomes isolated from adipose stem cells (ASC) display a neuroprotective role in an in vitro model of ALS. Moreover, the internalization of ASC-exosomes by the cells was shown and the molecules and the mechanisms by which exosomes could exert their beneficial effect were addressed. We performed for the first time a comprehensive proteomic analysis of exosomes derived from murine ASC. We identified a total of 189 proteins and the shotgun proteomics analysis revealed that the exosomal proteins are mainly involved in cell adhesion and negative regulation of the apoptotic process. We correlated the protein content to the anti-apoptotic effect of exosomes observing a downregulation of pro-apoptotic proteins Bax and cleaved caspase-3 and upregulation of anti-apoptotic protein Bcl-2 α, in an in vitro model of ALS after cell treatment with exosomes. Overall, this study shows the neuroprotective effect of ASC-exosomes after their internalization and their global protein profile, that could be useful to understand how exosomes act, demonstrating that they can be employed as therapy in neurodegenerative diseases. Full article
(This article belongs to the Section Intracellular and Plasma Membranes)
Show Figures

Figure 1

Figure 1
<p>TEM and western blot analysis of adipose stem cells (ASC)-exosomes. Electron microscopy shows vesicles with characteristic morphology and size of exosomes. Scale bar, 100 nm (<b>A</b>). The blots show western blot detection of the expression of HSP70 (70 kDa), CD9 (25 kDa) and CD81 (26 kDa) in exosomes (EXO); ASC lysates (ASC) was used as positive control (<b>B</b>).</p>
Full article ">Figure 2
<p>Gene ontology (GO) enrichment of the ASC-exosomes identified proteins according to Database for Annotation, Visualization and Integrated Discovery (DAVID) functional annotation. The top 10 enriched biological process (<b>A</b>) molecular function (<b>B</b>) and cellular component (<b>C</b>) are reported. The percentage represents the portion of the genes encoding the proteins with the corresponding gene ontology biological processes (GOBPs), gene ontology molecular functions (GOMFs) or gene ontology cellular components (GOCCs) in the ASC-exosomes proteins.</p>
Full article ">Figure 3
<p>Cytoscape based ClueGo/CluePedia pathway analysis and visualization. Enriched pathways were obtained from the Kyoto Encyclopaedia of Genes and Genome (KEGG) database. Terms are grouped based on shared genes (kappa score) showed in different colors. The size of nodes indicated the degree of significance. The most significant term defined the name of the group.</p>
Full article ">Figure 4
<p>Protein network of identified ASC-exosomes proteins. Schematic view of known and predicted protein interactions according to the STRING database (v. 10). Each node represents a protein, and each edge represents an interaction. Only interactions with the medium confidence score (0.4) are shown. Interactions include physical and functional associations, showing the evidence view.</p>
Full article ">Figure 5
<p>Western blot analysis of phospho-Akt and <span class="html-italic">SOD1</span> expression in ASC-exosomes and ASC. The blots show western blot detection of the expression of phospho-Akt (60 kDa) (<b>A</b>) and SOD1 (16 kDa) (<b>B</b>) in exosomes (EXO); ASC lysates were used as positive control. Amido Black staining was used as total loading control (<b>C</b>).</p>
Full article ">Figure 6
<p>Zymogram assay of SOD1 protein. The assay shows that exosomes (EXO) contain the active form of SOD1 protein. The ASC and deprived ASC (ASC-) were used as the positive control (<b>A</b>). Comassie blue staining was used as total loading control (<b>B</b>).</p>
Full article ">Figure 7
<p>Acridine orange/propidium iodide (AO/PI) double staining on NSC-34(G93A) cells. Apoptotic and live cells were visualized after AO/PI double staining. The green fluorescence staining by AO indicate live cells, while orange/red fluorescence indicate the PI staining that bound to DNA after damaged membranes. The image shows cells in a basal condition (no cell death was detected and nucleus is uniformly distributed), cells after H<sub>2</sub>O<sub>2</sub> treatment (the nucleus is located in bias and apoptosis-associated changes of cell membranes can be detected, indicating a process of disintegration) and cells after treatment with H<sub>2</sub>O<sub>2</sub> and exosomes (H<sub>2</sub>O<sub>2</sub> + EXO) in which a rescue of cells from death is detected, with an increase in cell viability compared to cells after H<sub>2</sub>O<sub>2</sub> treatment Magnification 20× (<b>A</b>). The graph shows the percentage of cell viability of NSC-34(G93A) cells in basal condition and after H<sub>2</sub>O<sub>2</sub> and ASC-exosomes treatment (H<sub>2</sub>O<sub>2</sub> + EXO). Cell viability significant increased after ASC-exosomes treatment. One-way ANOVA and Bonferroni post-hoc analysis were performed between all the experimental conditions (*** <span class="html-italic">p</span> &lt; 0.001) (<b>B</b>).</p>
Full article ">Figure 8
<p>TEM images of cells treated with exosomes-ultra-small superparamagnetic iron oxide nanoparticles (USPIO). TEM images showed no damaged cell after ASC-exosomes treatment; magnification 4400×, scale bar 1 µm (<b>A</b>). In (<b>B</b>) note a representative image of phospholipidic membrane structure contained high electron-density particles, whose dimension are attributable to USPIO nanoparticles used to label ASC-exosomes; magnification 46,000×, scale bar 100 nm. In (<b>C</b>), a higher magnification of the section squared in (<b>B</b>) is shown; magnification 140,000×, scale bar 50 nm.</p>
Full article ">Figure 9
<p>Expression profile of apoptotic markers in NSC-34(G93A) cells. The blots show western blot analysis of Cleaved Caspase 3 (<b>A</b>), Bax (<b>B</b>) and Bcl-2 α (<b>C</b>) proteins performed in NSC-34(G93A) cells (used as control, CNTR), NSC-34 (G93A) cells treated with H<sub>2</sub>O<sub>2</sub> and NSC-34 (G93A) cells treated with H<sub>2</sub>O<sub>2</sub> and exosomes (EXO). Amido Black staining was used as total loading control (<b>D</b>).</p>
Full article ">
15 pages, 1410 KiB  
Article
Regulation and Function of C-Type Natriuretic Peptide (CNP) in Gonadotrope-Derived Cell Lines
by Samantha M Mirczuk, Andrew J Lessey, Alice R Catterick, Rebecca M Perrett, Christopher J Scudder, Jordan E Read, Victoria J Lipscomb, Stijn J Niessen, Andrew J Childs, Craig A McArdle, Imelda M McGonnell and Robert C Fowkes
Cells 2019, 8(9), 1086; https://doi.org/10.3390/cells8091086 - 14 Sep 2019
Cited by 9 | Viewed by 4139
Abstract
C-type natriuretic peptide (CNP) is the most conserved member of the mammalian natriuretic peptide family, and is implicated in the endocrine regulation of growth, metabolism and reproduction. CNP is expressed throughout the body, but is particularly abundant in the central nervous system and [...] Read more.
C-type natriuretic peptide (CNP) is the most conserved member of the mammalian natriuretic peptide family, and is implicated in the endocrine regulation of growth, metabolism and reproduction. CNP is expressed throughout the body, but is particularly abundant in the central nervous system and anterior pituitary gland. Pituitary gonadotropes are regulated by pulsatile release of gonadotropin releasing hormone (GnRH) from the hypothalamus, to control reproductive function. GnRH and CNP reciprocally regulate their respective signalling pathways in αT3-1 gonadotrope cells, but effects of pulsatile GnRH stimulation on CNP expression has not been explored. Here, we examine the sensitivity of the natriuretic peptide system in LβT2 and αT3-1 gonadotrope cell lines to continuous and pulsatile GnRH stimulation, and investigate putative CNP target genes in gonadotropes. Multiplex RT-qPCR assays confirmed that primary mouse pituitary tissue express Nppc, Npr2 (encoding CNP and guanylyl cyclase B (GC-B), respectively) and Furin (a CNP processing enzyme), but failed to express transcripts for Nppa or Nppb (encoding ANP and BNP, respectively). Pulsatile, but not continuous, GnRH stimulation of LβT2 cells caused significant increases in Nppc and Npr2 expression within 4 h, but failed to alter natriuretic peptide gene expression in αT3-1 cells. CNP enhanced expression of cJun, Egr1, Nr5a1 and Nr0b1, within 8 h in LβT2 cells, but inhibited Nr5a1 expression in αT3-1 cells. Collectively, these data show the gonadotrope natriuretic peptide system is sensitive to pulsatile GnRH signalling, and gonadotrope transcription factors are putative CNP-target genes. Such findings represent additional mechanisms by which CNP may regulate reproductive function. Full article
Show Figures

Figure 1

Figure 1
<p>Expression profiling of the natriuretic peptide system in primary mouse endocrine tissues. (<b>A</b>) Murine mRNA sequences obtained from NCBI Nucleotide (<a href="https://www.ncbi.nlm.nih.gov/nuccore" target="_blank">https://www.ncbi.nlm.nih.gov/nuccore</a>), were imported into express Designer Software (Beckman Coulter), from which multiplex primers were designed using the following parameters: maximum PCR product = 300 nt, minimum PCR product = 100 nt, minimum separation size = 7 nt. Multiplex PCR reactions were performed, using specific primers for <span class="html-italic">Nppa</span>, <span class="html-italic">Nppb</span>, <span class="html-italic">Nppc</span>, <span class="html-italic">Npr1</span>, <span class="html-italic">Npr2</span>, <span class="html-italic">Npr3</span>, <span class="html-italic">Furin</span>, <span class="html-italic">Corin</span>, <span class="html-italic">ActB</span>, <span class="html-italic">Gapdh</span> and <span class="html-italic">Rpl19</span> and an internal positive control KanR. As shown (<a href="#cells-08-01086-f001" class="html-fig">Figure 1</a>A), capillary gel electrophoresis was used to separate specific PCR products (blue peaks), and compared alongside the appropriate size standard (red peaks, 140–420 nt). (<b>B</b>) RNA was isolated from range of tissues from 12 week old male and female C57/B6 mice (heart, adrenal, adipose, brain, pituitary, kidney, liver, testis and ovaries; n = 5 to 8). Data shown are means (n = 5 to 8) of relative gene expression (normalized to <span class="html-italic">ActB</span>; red indicates low level expression, green indicates high level expression, white indicates no transcript detected).</p>
Full article ">Figure 2
<p>Natriuretic peptide expression profile in gonadotrope-derived cell lines. mRNA expression of natriuretic peptide components in untreated αT3-1 and LβT2 cells (<span class="html-italic">Nppa</span> and <span class="html-italic">Nppb</span> were not detected). Data shown are means ± SEM (n = 5 individual RNA extractions) of relative gene expression (normalized to <span class="html-italic">ActB</span>).</p>
Full article ">Figure 3
<p>Effect of continuous or pulsatile GnRH stimulation on natriuretic peptide gene expression in LβT2 and αT3-1 cell lines. (<b>A</b>) Schematic of pulsatile or continuous GnRH treatment protocol, including wash periods. LβT2 cells (<b>B</b>) or αT3-1 cells (<b>C</b>) were treated with 0 or 100 nM GnRH, for either 4 h continuously, or as 5 min pulses every hour for 4 h, before extracting RNA and performing multiplex RT-qPCR to examine alterations in gene expression profiling of natriuretic peptide system (<span class="html-italic">Nppc</span>, <span class="html-italic">Furin</span>, <span class="html-italic">Corin</span>, <span class="html-italic">Npr1</span>, <span class="html-italic">Npr2</span>, <span class="html-italic">Npr3</span>). Data shown are means ± SEM (n = 6 to 9 individual RNA extractions) of relative gene expression (normalized to <span class="html-italic">ActB</span>); *** <span class="html-italic">p</span> &lt; 0.001, ** <span class="html-italic">p</span> &lt; 0.01, significantly different from basal, or continuous vs pulsatile, as indicated).</p>
Full article ">Figure 4
<p>Effect of continuous or pulsatile GnRH stimulation on gonadotrope transcription factor gene expression in LβT2 and αT3-1 cell lines. (<b>A</b>) Comparison of gonadotrope transcription factor expression (<span class="html-italic">cFos</span>, <span class="html-italic">cJun</span>, <span class="html-italic">Egr1</span>, <span class="html-italic">Nr5a1</span>, <span class="html-italic">Nr0b1</span>) in LβT2 and αT3-1 cells. Data shown are means (n = 5 to 8 individual RNA extractions) of relative gene expression (normalized to <span class="html-italic">ActB</span>; red indicates low level expression, green indicates high level expression. (<b>B</b>,<b>C</b>) LβT2 cells (<b>B</b>) or αT3-1 cells (<b>C</b>) were treated with 0 or 100 nM GnRH, for either 4 h continuously, or as 5 min pulses every hour for 4 h, before extracting RNA and performing multiplex RT-qPCR for gonadotrope transcription factors. Data shown are means ± SEM (n = 6 to 9 individual RNA extractions) of relative gene expression (normalized to <span class="html-italic">ActB</span>); *** <span class="html-italic">p</span> &lt; 0.001, ** <span class="html-italic">p</span> &lt; 0.01, * <span class="html-italic">p</span> &lt; 0.05, significantly different from basal, or continuous vs pulsatile, as indicated).</p>
Full article ">Figure 5
<p>Effect of CNP on gonadotrope transcription factor gene expression in LβT2 and αT3-1 cells. LβT2 (<b>A</b>) and αT3-1 (<b>B</b>) cells were stimulated with 0 or 100 nM CNP for 0, 4, 8, and 24 h, before extracting RNA and performing multiplex RT-qPCR for gonadotrope transcription factors. Data shown are means ± SEM (n = 6 individual RNA extractions) of relative gene expression (normalized to <span class="html-italic">ActB</span>); **** <span class="html-italic">p</span> &lt; 0.0001, *** <span class="html-italic">p</span> &lt; 0.001, ** <span class="html-italic">p</span> &lt; 0.01, * <span class="html-italic">p</span> &lt; 0.05, significantly different from basal (0 h).</p>
Full article ">
12 pages, 1741 KiB  
Article
Empagliflozin Protects HK-2 Cells from High Glucose-Mediated Injuries via a Mitochondrial Mechanism
by Wen-Chin Lee, You-Ying Chau, Hwee-Yeong Ng, Chiu-Hua Chen, Pei-Wen Wang, Chia-Wei Liou, Tsu-Kung Lin and Jin-Bor Chen
Cells 2019, 8(9), 1085; https://doi.org/10.3390/cells8091085 - 14 Sep 2019
Cited by 53 | Viewed by 10926
Abstract
Empagliflozin is known to retard the progression of kidney disease in diabetic patients. However, the underlying mechanism is incompletely understood. High glucose induces oxidative stress in renal tubules, eventually leading to mitochondrial damage. Here, we investigated whether empagliflozin exhibits protective functions in renal [...] Read more.
Empagliflozin is known to retard the progression of kidney disease in diabetic patients. However, the underlying mechanism is incompletely understood. High glucose induces oxidative stress in renal tubules, eventually leading to mitochondrial damage. Here, we investigated whether empagliflozin exhibits protective functions in renal tubules via a mitochondrial mechanism. We used human proximal tubular cell (PTC) line HK-2 and employed western blotting, terminal deoxynucleotidyl transferase dUTP nick end labelling assay, fluorescence staining, flow cytometry, and enzyme-linked immunosorbent assay to investigate the impact of high glucose and empagliflozin on cellular apoptosis, mitochondrial morphology, and functions including mitochondrial membrane potential (MMP), reactive oxygen species (ROS) production, and adenosine triphosphate (ATP) generation. We found that PTCs were susceptible to high glucose-induced mitochondrial fragmentation, and empagliflozin ameliorated this effect via the regulation of mitochondrial fission (FIS1 and DRP1) and fusion (MFN1 and MFN2) proteins. Empagliflozin reduced the high glucose-induced cellular apoptosis and improved mitochondrial functions by restoring mitochondrial ROS production, MMP, and ATP generation. Our results suggest that empagliflozin may protect renal PTCs from high glucose-mediated injuries through a mitochondrial mechanism. This could be one of the novel mechanisms explaining the benefits demonstrated in EMPA-REG OUTCOME trial. Full article
Show Figures

Figure 1

Figure 1
<p>Empagliflozin has no negative effects on the viability of HK-2 cells. Cell viability was measured with crystal violet assay and the absorbance was analyzed at 570 nm using a microplate reader. Data were obtained from three independent experiments and are expressed as mean ± SEM. Only high glucose (30 mM) treatment led to a decrease in cell viability. * <span class="html-italic">P</span> &lt; 0.05 versus the normal glucose (5 mM) treatment group. Glu, glucose; Empa, empagliflozin.</p>
Full article ">Figure 2
<p>Empagliflozin rescues high glucose-induced mitochondrial fragmentation in human PTCs. HK-2 cells were cultured in 5 mM glucose (<b>A</b>), 30 mM glucose (<b>B</b>), 30 mM glucose with 100 nM empagliflozin (<b>C</b>), and 30 mM glucose with 500 nM empagliflozin (<b>D</b>). Mitochondria were stained with Mitotracker red. Magnified images of indexed mitochondria (arrow) in each treatment group are shown in (<b>E</b>–<b>H</b>). (<b>I</b>) The mitochondrial fission rate was higher in high glucose condition. Empagliflozin rescued this effect. Data were obtained from three independent experiments and are expressed as 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>Impact of empagliflozin on high glucose-induced alterations in the expression levels of mitochondrial fusion/fission proteins. (<b>A</b>–<b>E</b>) The expression of MFN1, MFN2, DRP1, and FIS1 was analyzed with western blotting and normalized to the level of β-actin. Data are expressed as mean ± SEM. * <span class="html-italic">P</span> &lt; 0.05, ** <span class="html-italic">P</span> &lt; 0.001.</p>
Full article ">Figure 4
<p>Empagliflozin reduces the high glucose-induced apoptosis of HK-2 cells. (<b>A</b>–<b>D</b>) Fluorescence images show positive TUNEL staining in the four treatment groups. (<b>E</b>) Quantitative analysis of TUNEL assay shows that high glucose markedly induces apoptosis and empagliflozin ameliorates this effect. Data are expressed as mean ± SEM. ** <span class="html-italic">P</span> &lt; 0.001.</p>
Full article ">Figure 5
<p>Empagliflozin improves the mitochondrial function of high glucose-treated HK-2 cells. (<b>A</b>) Cellular ROS production, (<b>B</b>) mitochondrial ROS production, (<b>C</b>) MMP, and (<b>D</b>) ATP generation in the four treatment groups. Data were obtained from three independent experiments and are expressed as mean ± SEM. * <span class="html-italic">P</span> &lt; 0.05, ** <span class="html-italic">P</span> &lt; 0.001.</p>
Full article ">
12 pages, 2074 KiB  
Article
Effect of Adenomatous Polyposis Coli Loss on Tumorigenic Potential in Pancreatic Ductal Adenocarcinoma
by Jennifer M. Cole, Kaitlyn Simmons and Jenifer R. Prosperi
Cells 2019, 8(9), 1084; https://doi.org/10.3390/cells8091084 - 14 Sep 2019
Cited by 9 | Viewed by 3926
Abstract
Loss of the Adenomatous Polyposis Coli (APC) tumor suppressor in colorectal cancer elicits rapid signaling through the Wnt/β-catenin signaling pathway. In contrast to this well-established role of APC, recent studies from our laboratory demonstrated that APC functions through Wnt-independent pathways to [...] Read more.
Loss of the Adenomatous Polyposis Coli (APC) tumor suppressor in colorectal cancer elicits rapid signaling through the Wnt/β-catenin signaling pathway. In contrast to this well-established role of APC, recent studies from our laboratory demonstrated that APC functions through Wnt-independent pathways to mediate in vitro and in vivo models of breast tumorigenesis. Pancreatic ductal adenocarcinoma (PDAC) has an overall median survival of less than one year with a 5-year survival rate of 7.2%. APC is lost in a subset of pancreatic cancers, but the impact on Wnt signaling or tumor development is unclear. Given the lack of effective treatment strategies for pancreatic cancer, it is important to understand the functional implications of APC loss in pancreatic cancer cell lines. Therefore, the goal of this project is to study how APC loss affects Wnt pathway activation and in vitro tumor phenotypes. Using lentiviral shRNA, we successfully knocked down APC expression in six pancreatic cancer cell lines (AsPC-1, BxPC3, L3.6pl, HPAF-II, Hs 766T, MIA PaCa-2). No changes were observed in localization of β-catenin or reporter assays to assess β-catenin/TCF interaction. Despite this lack of Wnt/β-catenin pathway activation, the majority of APC knockdown cell lines exhibit an increase in cell proliferation. Cell migration assays showed that the BxPC-3 and L3.6pl cells were impacted by APC knockdown, showing faster wound healing in scratch wound assays. Interestingly, APC knockdown had no effect on gemcitabine treatment, which is the standard care for pancreatic cancer. It is important to understand the functional implications of APC loss in pancreatic cancer cells lines, which could be used as a target for therapeutics. Full article
Show Figures

Figure 1

Figure 1
<p>Quantitative PCR demonstrates that APC is knocked down in the cell lines transfected with APC shRNA constructs. Using lentiviral-mediated MISSION shRNA we verified <span class="html-italic">APC</span> knock down in six pancreatic ductal adenocarcinoma cell lines using quantitative PCR and relative expression of <span class="html-italic">APC</span> revealed gene knockdown in AsPC-1, BxPC-3, HPAF-II, Hs 766T, L3.6pl, and MIA PaCa-2 cell lines compared to CTL (<a href="#cells-08-01084-f001" class="html-fig">Figure 1</a><b>A</b>–<b>F</b>). *<span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 2
<p>APC knockdown did not increase nuclear β-catenin. (<b>A</b>) In AsPC-1, BxPC-3, HPAF-II, Hs 766T, L3.6pl and MIA PaCa-2 CTL cells, β-catenin is localized at the cell–cell junctions. After the knockdown of APC in each of those cell lines: APC shRNA (representative), there were no changes in beta catenin localization. (<b>B</b>) In the APC knockdown PDAC cell lines, we observed no alterations in Wnt/β-catenin signaling using β-catenin/TCF reporter assays.</p>
Full article ">Figure 3
<p>Proliferation changes are observed in some cell lines that have been transfected with APC shRNA constructs compared to controls. Proliferation was measured in total cell counts at 24, 48 and 72 h post-plating for each cell line. (<b>B</b> and <b>E</b>) In the BxPC-3 and L3.6pl cells, knockdown of APC resulted in a significant increase in cell proliferation at 48 h. (<b>D</b>) In the Hs 766T cells, knockdown of APC resulted in a significant increase in cell proliferation at 48 and 72 h. (<b>A</b>, <b>C</b>, and <b>F</b>) The AsPC-1, HPAF-II and MIA PaCa-2 cell lines showed no significant differences in cell proliferation of the APC shRNA cells lines compared to CTL. *<span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 4
<p>Migration of cells in wound healing assays revealed faster migration with APC knockdown in some pancreatic cancer cell lines. (<b>A</b>) The AsPC-1 APC shRNA2 knockdown cells migrated slower than the control at 24 and 48 h post wounding. (<b>B</b> and <b>E</b>) Only the BxPC-3 and L3.6pl cells were impacted by APC knockdown, showing faster wound healing. Specifically, at 48 h post-wounding, the APC shRNA1 and APC shRNA2 cell lines were significantly more closed than the parent BxPC-3 cells. (<b>E</b>) In L3.6pl, there were significant differences in percent filled area in APC shRNA1 and APC shRNA2 at 24 h and APC shRNA2 at 48 h compared to CTL. (<b>A</b>, <b>C</b>, <b>D</b>, and <b>F</b>) In AsPC-1, HPAF-II, Hs766T, and MIA PaCa-2 cells, there were no significant differences in percent filled area in APC shRNA cells versus CTL. *<span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 5
<p>Effect of gemcitabine on PDAC cell lines. (<b>A</b>) In AsPC-1 cells, we saw significant decreases in total cell numbers in controls and APC shRNA1, but not in APC shRNA2 when treated with 1 µM gemcitabine for 48 h. (<b>B</b>) In BxPC-3 cells, 1 µM gemcitabine decreased total cell numbers in all cell lines. (<b>C</b>) In HPAF-II cells, 1 µM gemcitabine decreased total cell numbers in APC shRNA1, and APC shRNA2, but not in the CTL. (<b>D</b>) In Hs 766T cells, 48 h treatment with 1 µM gemcitabine resulted in significant decreases in total cell numbers in controls and APC shRNA1, but not in APC shRNA2. (<b>E</b>) In L3.6pl cells, 1 µM gemcitabine decreased total cell numbers in CTL APC shRNA1, and APC shRNA2. (<b>F</b>) In MIA PaCa-2 cells, 1 µM gemcitabine decreased total cell numbers in CTL, APC shRNA1, and APC shRNA2. *<span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">
13 pages, 902 KiB  
Review
Improving Cancer Immunotherapy by Targeting the Hypoxic Tumor Microenvironment: New Opportunities and Challenges
by Muhammad Zaeem Noman, Meriem Hasmim, Audrey Lequeux, Malina Xiao, Caroline Duhem, Salem Chouaib, Guy Berchem and Bassam Janji
Cells 2019, 8(9), 1083; https://doi.org/10.3390/cells8091083 - 14 Sep 2019
Cited by 161 | Viewed by 10149
Abstract
Initially believed to be a disease of deregulated cellular and genetic expression, cancer is now also considered a disease of the tumor microenvironment. Over the past two decades, significant and rapid progress has been made to understand the complexity of the tumor microenvironment [...] Read more.
Initially believed to be a disease of deregulated cellular and genetic expression, cancer is now also considered a disease of the tumor microenvironment. Over the past two decades, significant and rapid progress has been made to understand the complexity of the tumor microenvironment and its contribution to shaping the response to various anti-cancer therapies, including immunotherapy. Nevertheless, it has become clear that the tumor microenvironment is one of the main hallmarks of cancer. Therefore, a major challenge is to identify key druggable factors and pathways in the tumor microenvironment that can be manipulated to improve the efficacy of current cancer therapies. Among the different tumor microenvironmental factors, this review will focus on hypoxia as a key process that evolved in the tumor microenvironment. We will briefly describe our current understanding of the molecular mechanisms by which hypoxia negatively affects tumor immunity and shapes the anti-tumor immune response. We believe that such understanding will provide insight into the therapeutic value of targeting hypoxia and assist in the design of innovative combination approaches to improve the efficacy of current cancer therapies, including immunotherapy. Full article
(This article belongs to the Special Issue Tumor Microenvironment: Interaction and Metabolism)
Show Figures

Figure 1

Figure 1
<p>The hypoxic tumor microenvironment and its impact on anti-tumor immunity. (<b>A</b>) Hypoxia is established in the tumor microenvironment due to an increase in tumor cell proliferation, and a decrease in oxygen supply. Non-hypoxic tumor regions displayed normal blood vessels covered by well-organized endothelial cells and pericytes. In hypoxic tumor regions, the pressure of oxygen is low which arises from oxygen diffusion limitations due to disorganized, chaotic tumor microvasculature network with leaky vessels. (<b>B</b>) Under hypoxia, the stabilization of hypoxia-inducible factor (HIF)-1α in cells upregulates the expression of PD-L1 in hypoxic tumor cells and PD-L1 and VISTA in hypoxic MDSCs. The increased expression of PD-L1 and VISTA results in an inhibition of T cell proliferation and T cell mediated lysis. (<b>C</b>) HIF-1α is also involved in the upregulation of cluster of differentiation 47 (CD47) on the surface of tumor cells. Following the binding of CD47 to signal regulatory protein α (SIRPα), expressed on the surface of macrophages, tumor cells provide a strong “don’t eat me signal” to block phagocytosis property of macrophages. (<b>D</b>) The activation of autophagy in hypoxic tumor cells impairs tumor cell susceptibility to CTL and NK-mediated lysis by at least two distinct mechanisms involving the degradation of NK-derived Granzyme B and the stabilization of pSTAT3. Other hypoxia-dependent, but autophagy-independent, mechanisms are described in this review including the overexpression of NANOG and miR-210 targeting PTPN1, HOXA1, and TP53I11. (<b>E</b>) Hypoxia upregulates the expression of HLA-G on the surface of tumor cells. The upregulated HLA-G binds to ILT2, ILT4 and KIR2DL4 expressed by several immune cells (B and T cells, NK cells, myelomonocytic cells, dendritic cells, monocytes and macrophages) leading to tumor escape from immune surveillance.</p>
Full article ">
20 pages, 4450 KiB  
Article
Endoglin Protein Interactome Profiling Identifies TRIM21 and Galectin-3 as New Binding Partners
by Eunate Gallardo-Vara, Lidia Ruiz-Llorente, Juan Casado-Vela, María J. Ruiz-Rodríguez, Natalia López-Andrés, Asit K. Pattnaik, Miguel Quintanilla and Carmelo Bernabeu
Cells 2019, 8(9), 1082; https://doi.org/10.3390/cells8091082 - 13 Sep 2019
Cited by 25 | Viewed by 6022
Abstract
Endoglin is a 180-kDa glycoprotein receptor primarily expressed by the vascular endothelium and involved in cardiovascular disease and cancer. Heterozygous mutations in the endoglin gene (ENG) cause hereditary hemorrhagic telangiectasia type 1, a vascular disease that presents with nasal and gastrointestinal bleeding, skin [...] Read more.
Endoglin is a 180-kDa glycoprotein receptor primarily expressed by the vascular endothelium and involved in cardiovascular disease and cancer. Heterozygous mutations in the endoglin gene (ENG) cause hereditary hemorrhagic telangiectasia type 1, a vascular disease that presents with nasal and gastrointestinal bleeding, skin and mucosa telangiectases, and arteriovenous malformations in internal organs. A circulating form of endoglin (alias soluble endoglin, sEng), proteolytically released from the membrane-bound protein, has been observed in several inflammation-related pathological conditions and appears to contribute to endothelial dysfunction and cancer development through unknown mechanisms. Membrane-bound endoglin is an auxiliary component of the TGF-β receptor complex and the extracellular region of endoglin has been shown to interact with types I and II TGF-β receptors, as well as with BMP9 and BMP10 ligands, both members of the TGF-β family. To search for novel protein interactors, we screened a microarray containing over 9000 unique human proteins using recombinant sEng as bait. We find that sEng binds with high affinity, at least, to 22 new proteins. Among these, we validated the interaction of endoglin with galectin-3, a secreted member of the lectin family with capacity to bind membrane glycoproteins, and with tripartite motif-containing protein 21 (TRIM21), an E3 ubiquitin-protein ligase. Using human endothelial cells and Chinese hamster ovary cells, we showed that endoglin co-immunoprecipitates and co-localizes with galectin-3 or TRIM21. These results open new research avenues on endoglin function and regulation. Full article
(This article belongs to the Special Issue TGF-beta/BMP Signaling Pathway)
Show Figures

Figure 1

Figure 1
<p>Protein–protein association between galectin-3 and endoglin. (<b>A</b>–<b>C</b>). Co-immunoprecipitation of galectin-3 and endoglin. CHO-K1 cells were transiently transfected with pcEXV-Ø (Ø), pcEXV–HA–EngFL (Eng) and pcDNA3.1–Gal-3 (Gal3) expression vectors. (<b>A</b>) Total cell lysates (TCL) were analyzed by SDS-PAGE under reducing conditions, followed by Western blot (WB) analysis using specific antibodies to endoglin, galectin-3 and β-actin (loading control). Cell lysates were subjected to immunoprecipitation (IP) with anti-endoglin (<b>B</b>) or anti-galectin-3 (<b>C</b>) antibodies, followed by SDS-PAGE under reducing conditions and WB analysis with anti-endoglin or anti-galectin-3 antibodies, as indicated. Negative controls with an IgG2b (<b>B</b>) and IgG1 (<b>C</b>) were included. (<b>D</b>) Protein-protein interactions between galectin-3 and endoglin using Bio-layer interferometry (BLItz). The Ni–NTA biosensors tips were loaded with 7.3 µM recombinant human galectin-3/6xHis at the C-terminus (LGALS3), and protein binding was measured against 0.1% BSA in PBS (negative control) or 4.1 µM soluble endoglin (sEng). Kinetic sensorgrams were obtained using a single channel ForteBioBLItzTM instrument.</p>
Full article ">Figure 2
<p>Galectin-3 and endoglin co-localize in human endothelial cells. Human umbilical vein-derived endothelial cell (HUVEC) monolayers were fixed with paraformaldehyde, permeabilized with Triton X-100, incubated with the mouse mAb P4A4 anti-endoglin, washed, and incubated with a rabbit polyclonal anti-galectin-3 antibody (PA5-34819). Galectin-3 and endoglin were detected by immunofluorescence upon incubation with Alexa 647 goat anti-rabbit IgG (red staining) and Alexa 488 goat anti-mouse IgG (green staining) secondary antibodies, respectively. (<b>A</b>) Single staining of galectin-3 (red) and endoglin (green) at the indicated magnifications. (<b>B</b>) Merge images plus DAPI (nuclear staining in blue) show co-localization of galectin-3 and endoglin (yellow color). Representative images of five different experiments are shown.</p>
Full article ">Figure 3
<p>Protein–protein association between TRIM21 and endoglin. (<b>A</b>–<b>E</b>) Co-immunoprecipitation of TRIM21 and endoglin. A,B. HUVEC monolayers were lysed and total cell lysates (TCL) were subjected to SDS-PAGE under reducing (for TRIM21 detection) or nonreducing (for endoglin detection) conditions, followed by Western blot (WB) analysis using antibodies to endoglin, TRIM21 or β-actin (<b>A</b>). HUVECs lysates were subjected to immunoprecipitation (IP) with anti-TRIM21 or negative control antibodies, followed by WB analysis with anti-endoglin (<b>B</b>). C,D. CHO-K1 cells were transiently transfected with pDisplay–HA–Mock (Ø), pDisplay–HA–EngFL (<b>E</b>) or pcDNA3.1–HA–hTRIM21 (T) expression vectors, as indicated. Total cell lysates (TCL) were subjected to SDS-PAGE under nonreducing conditions and WB analysis using specific antibodies to endoglin, TRIM21, and β-actin (<b>C</b>). Cell lysates were subjected to immunoprecipitation (IP) with anti-TRIM21 or anti-endoglin antibodies, followed by SDS-PAGE under reducing (upper panel) or nonreducing (lower panel) conditions and WB analysis with anti-TRIM21 or anti-endoglin antibodies. Negative controls of appropriate IgG were included (<b>D</b>). E. CHO-K1 cells were transiently transfected with pcDNA3.1–HA–hTRIM21 and pDisplay–HA–Mock (Ø), pDisplay–HA–EngFL (FL; full-length), pDisplay–HA–EngEC (EC; cytoplasmic-less) or pDisplay–HA–EngTMEC (TMEC; cytoplasmic-less) expression vectors, as indicated. Cell lysates were subjected to immunoprecipitation with anti-TRIM21, followed by SDS-PAGE under reducing conditions and WB analysis with anti-endoglin antibodies, as indicated. The asterisk indicates the presence of a nonspecific band. Mr, molecular reference; Eng, endoglin; TRIM, TRIM21. (<b>F</b>) Protein–protein interactions between TRIM21 and endoglin using Bio-layer interferometry (BLItz). The Ni–NTA biosensors tips were loaded with 5.4 µM recombinant human TRIM21/6xHis at the N-terminus (R052), and protein binding was measured against 0.1% BSA in PBS (negative control) or 4.1 µM soluble endoglin (sEng). Kinetic sensorgrams were obtained using a single channel ForteBioBLItzTM instrument.</p>
Full article ">Figure 4
<p>TRIM21 and endoglin co-localize in human endothelial cells. HUVEC monolayers were fixed with paraformaldehyde, permeabilized with Triton X-100, incubated with the mouse mAb P4A4 anti-endoglin, washed and incubated with a rabbit monoclonal anti-TRIM21 antibody (#92043). TRIM21 and endoglin were detected by immunofluorescence upon incubation with Alexa 647 goat anti-rabbit IgG (red staining) and Alexa 488 goat anti-mouse IgG (green staining) secondary antibodies, respectively. (<b>A</b>) Single staining of TRIM21 (red) and endoglin (green) at the indicated magnifications. (<b>B</b>) Merge images plus DAPI (nuclear staining in blue) show co-localization of TRIM21 and endoglin (yellow color). Representative images of four different experiments are shown.</p>
Full article ">
Previous Issue
Next Issue
Back to TopTop