[go: up one dir, main page]

Next Issue
Volume 12, October-1
Previous Issue
Volume 12, September-1
 
 

Cells, Volume 12, Issue 18 (September-2 2023) – 126 articles

Cover Story (view full-size image): Chronic alcohol abuse leads to alterations in the gastrointestinal microbiota that are associated with behavioral, physiological, and immunological effects. However, the direct effects of alcohol-associated changes in the microbiome are ill-defined. To address this, we developed a humanized alcohol-microbiota mouse model, which allows us to systematically evaluate the immunological effects of chronic alcohol abuse mediated by changes in the intestinal microbiota. In this study, we discuss the effects of human alcohol-associated microbiota on pulmonary host defense against bacterial pneumonia. Our findings highlight the importance of considering both the direct effects of alcohol and alcohol-induced changes in the microbiota when investigating the mechanisms behind alcohol-related disorders and treatment strategies. 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:
20 pages, 17068 KiB  
Article
Cognitive Functions, Neurotransmitter Alterations, and Hippocampal Microstructural Changes in Mice Caused by Feeding on Western Diet
by Raly James Perez Custodio, Zaynab Hobloss, Maiju Myllys, Reham Hassan, Daniela González, Jörg Reinders, Julia Bornhorst, Ann-Kathrin Weishaupt, Abdel-latif Seddek, Tahany Abbas, Adrian Friebel, Stefan Hoehme, Stephan Getzmann, Jan G. Hengstler, Christoph van Thriel and Ahmed Ghallab
Cells 2023, 12(18), 2331; https://doi.org/10.3390/cells12182331 - 21 Sep 2023
Cited by 3 | Viewed by 2622
Abstract
Metabolic Dysfunction Associated Steatotic Liver Disease (MASLD) is the most common chronic liver disease in Western countries. It is becoming increasingly evident that peripheral organ-centered inflammatory diseases, including liver diseases, are linked with brain dysfunctions. Therefore, this study aims to unravel the effect [...] Read more.
Metabolic Dysfunction Associated Steatotic Liver Disease (MASLD) is the most common chronic liver disease in Western countries. It is becoming increasingly evident that peripheral organ-centered inflammatory diseases, including liver diseases, are linked with brain dysfunctions. Therefore, this study aims to unravel the effect of MASLD on brain histology, cognitive functions, and neurotransmitters. For this purpose, mice fed for 48 weeks on standard (SD) or Western diet (WD) were evaluated by behavioral tests, followed by sacrifice and analysis of the liver-brain axis including histopathology, immunohistochemistry, and biochemical analyses. Histological analysis of the liver showed features of Metabolic Dysfunction-Associated Steatohepatitis (MASH) in the WD-fed mice including lipid droplet accumulation, inflammation, and fibrosis. This was accompanied by an elevation of transaminase and alkaline phosphatase activities, increase in inflammatory cytokine and bile acid concentrations, as well as altered amino acid concentrations in the blood. Interestingly, compromised blood capillary morphology coupled with astrogliosis and microgliosis were observed in brain hippocampus of the WD mice, indicating neuroinflammation or a disrupted neurovascular unit. Moreover, attention was impaired in WD-fed mice along with the observations of impaired motor activity and balance, enhanced anxiety, and stereotyped head-twitch response (HTR) behaviors. Analysis of neurotransmitters and modulators including dopamine, serotonin, GABA, glutamate, and acetylcholine showed region-specific dysregulation in the brain of the WD-fed mice. In conclusion, the induction of MASH in mice is accompanied by the alteration of cellular morphology and neurotransmitter expression in the brain, associated with compromised cognitive functions. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Induction of steatohepatitis by 48 weeks of Western diet (WD) and standard diet (SD). (<b>A</b>) Body weight and liver-to-body-weight ratio. (<b>B</b>) Liver enzyme activities of alanine aminotransferase (ALT), aspartate aminotransferase (AST), and alkaline phosphatase (ALP) in plasma. (<b>C</b>) Hematoxylin-eosin (HE) staining, immunostaining with antibodies directed against the pan-leukocyte marker CD45, and visualization of fibrosis by Sirius red. Data are presented as the mean ± SE of 12 mice per test. Mann-Whitney U; ***: <span class="html-italic">p</span>  &lt;  0.001.</p>
Full article ">Figure 2
<p>Clinical chemistry and immunostaining of arginase1 and glutamine synthetase in mice after 48 weeks of Western diet (WD) and standard diet (SD). (<b>A</b>) TNF-α, IL-10, IL-12, and IL-6 in plasma. (<b>B</b>) Sum bile acids (BA) from plasma taken from the portal vein, hepatic vein, and left heart chamber. (<b>C</b>) Ammonia (plasma). (<b>D</b>) Urea (plasma). (<b>E</b>) Arginine (plasma). (<b>F</b>). Glutamine and glutamate (plasma). (<b>G</b>) Quantification of arginase1 and glutamine synthetase immunostainings from whole slide scans. (<b>H</b>). Representative immunostainings for arginase1 (periportal zonation) and glutamine synthetase (GS) (pericentral zonation). Data are presented as the mean ± SE of 12 mice per test. (<b>A</b>) Mann-Whitney U, (<b>B</b>–<b>H</b>) two-way ANOVA with Tukey’s multiple comparisons; *: <span class="html-italic">p</span> ≤ 0.05, **: <span class="html-italic">p</span>  &lt;  0.01; ***: <span class="html-italic">p</span>  &lt;  0.001.</p>
Full article ">Figure 3
<p>Motor, anxiety, impulsivity, and stereotypical behaviors of mice following 48 weeks of Western diet (WD) and standard diet (SD). (<b>A</b>,<b>B</b>) Total distance (cm) and movement duration (s) during the Open-field test (OFT). (<b>C</b>,<b>D</b>) Latency of fall (s) and falling frequency (#) in the Rotarod test (RT). (<b>E</b>,<b>F</b>) The center and peripheral field explorations (s) in the OFT. (<b>G</b>,<b>H</b>) Entry and time spent (%) in the open arms during the Elevated-plus maze test (EPMT). (<b>I</b>,<b>J</b>) Latency of fall (s) and falling frequency (#) in the Cliff-avoidance test (CAT). (<b>K</b>) Total counts of Head-twitch responses (HTR). (<b>L</b>) Total counts of jumps. (<b>M</b>) Total counts of buried marbles during the Marble-burying test (MBT). Data are presented as the mean ± SE of 20 mice per test. (<b>A</b>–<b>F</b>) two-way ANOVA with Tukey’s multiple comparisons, (<b>G</b>–<b>M</b>) two-tailed unpaired <span class="html-italic">t</span>-test. *: <span class="html-italic">p</span> ≤ 0.05, **: <span class="html-italic">p</span>  &lt;  0.01; ***: <span class="html-italic">p</span>  &lt;  0.001. Representative tracks, heatmaps, and photographs (marble burying) are shown.</p>
Full article ">Figure 4
<p>Cognitive behaviors of mice after 48 weeks of Western diet (WD) and standard diet (SD). (<b>A</b>,<b>B</b>) Investigation time (s) and discrimination index in the Novel-object recognition test (NORT). (<b>C</b>,<b>D</b>) Recognition and preference indices in the Object-based recognition test (OBAT). (<b>E</b>,<b>F</b>) Arms entry (#) and spontaneous alternations (%) in the Y-maze test (YMT). (<b>G</b>–<b>I</b>) Latency (s), errors, and repeat errors (#) during the acquisition phase of the Barnes maze test (BMT). (<b>J</b>) Visits per quadrant during the Probe trial fifth day of BMT. Data are presented as the mean ± SE of 20 mice per test. Compared to SD group, (<b>A</b>–<b>F</b>) two-tailed unpaired <span class="html-italic">t</span>-test, (<b>D</b>,<b>E</b>) two-way ANOVA with Tukey’s multiple comparisons. **: <span class="html-italic">p</span>  &lt;  0.01; ***: <span class="html-italic">p</span>  &lt;  0.001. Representative tracks and heatmaps per experiment are shown.</p>
Full article ">Figure 5
<p>Neurotransmitter levels in specific brain regions of mice 48 weeks after Western diet (WD) and standard diet (SD). Data are presented as mean ± SE of 11 mice. Mann-Whitney U; *: <span class="html-italic">p</span> ≤ 0.05, **: <span class="html-italic">p</span>  &lt;  0.01.</p>
Full article ">Figure 6
<p>Immunostaining for serotonin on tissue sections of the hippocampus and cortex of the brain of mice 48 weeks after Western diet (WD) and standard diet (SD).</p>
Full article ">Figure 7
<p>Histological alterations of hippocampal tissue after 48 weeks of Western diet (WD). (<b>A</b>) Immunostaining for the endothelial marker CD31, the astrocyte marker GFAP, and the microglia marker lba1. DAPI visualizes nuclei. (<b>B</b>) 3D reconstructed individual structures: microvessel segments (from the area indicated in (<b>A</b>)), astrocyte, microglial cell, and the intertwined network of all three structures. (<b>C</b>–<b>E</b>) Quantifications of microvessels, astrocytes, and microglia. Data are presented as mean ± SE of 12 mice. Mann-Whitney U; *: <span class="html-italic">p</span> ≤ 0.05, **: <span class="html-italic">p</span>  &lt;  0.01.</p>
Full article ">
69 pages, 4604 KiB  
Review
Insight and Recommendations for Fragile X-Premutation-Associated Conditions from the Fifth International Conference on FMR1 Premutation
by Flora Tassone, Dragana Protic, Emily Graves Allen, Alison D. Archibald, Anna Baud, Ted W. Brown, Dejan B. Budimirovic, Jonathan Cohen, Brett Dufour, Rachel Eiges, Nicola Elvassore, Lidia V. Gabis, Samantha J. Grudzien, Deborah A. Hall, David Hessl, Abigail Hogan, Jessica Ezzell Hunter, Peng Jin, Poonnada Jiraanont, Jessica Klusek, R. Frank Kooy, Claudine M. Kraan, Cecilia Laterza, Andrea Lee, Karen Lipworth, Molly Losh, Danuta Loesch, Reymundo Lozano, Marsha R. Mailick, Apostolos Manolopoulos, Veronica Martinez-Cerdeno, Yingratana McLennan, Robert M. Miller, Federica Alice Maria Montanaro, Matthew W. Mosconi, Sarah Nelson Potter, Melissa Raspa, Susan M. Rivera, Katharine Shelly, Peter K. Todd, Katarzyna Tutak, Jun Yi Wang, Anne Wheeler, Tri Indah Winarni, Marwa Zafarullah and Randi J. Hagermanadd Show full author list remove Hide full author list
Cells 2023, 12(18), 2330; https://doi.org/10.3390/cells12182330 - 21 Sep 2023
Cited by 16 | Viewed by 6243
Abstract
The premutation of the fragile X messenger ribonucleoprotein 1 (FMR1) gene is characterized by an expansion of the CGG trinucleotide repeats (55 to 200 CGGs) in the 5’ untranslated region and increased levels of FMR1 mRNA. Molecular mechanisms leading to fragile [...] Read more.
The premutation of the fragile X messenger ribonucleoprotein 1 (FMR1) gene is characterized by an expansion of the CGG trinucleotide repeats (55 to 200 CGGs) in the 5’ untranslated region and increased levels of FMR1 mRNA. Molecular mechanisms leading to fragile X-premutation-associated conditions (FXPAC) include cotranscriptional R-loop formations, FMR1 mRNA toxicity through both RNA gelation into nuclear foci and sequestration of various CGG-repeat-binding proteins, and the repeat-associated non-AUG (RAN)-initiated translation of potentially toxic proteins. Such molecular mechanisms contribute to subsequent consequences, including mitochondrial dysfunction and neuronal death. Clinically, premutation carriers may exhibit a wide range of symptoms and phenotypes. Any of the problems associated with the premutation can appropriately be called FXPAC. Fragile X-associated tremor/ataxia syndrome (FXTAS), fragile X-associated primary ovarian insufficiency (FXPOI), and fragile X-associated neuropsychiatric disorders (FXAND) can fall under FXPAC. Understanding the molecular and clinical aspects of the premutation of the FMR1 gene is crucial for the accurate diagnosis, genetic counseling, and appropriate management of affected individuals and families. This paper summarizes all the known problems associated with the premutation and documents the presentations and discussions that occurred at the International Premutation Conference, which took place in New Zealand in 2023. Full article
Show Figures

Figure 1

Figure 1
<p>Molecular mechanisms leading to <span class="html-italic">FMR1</span>-PM-associated conditions. Three nonexclusive models are proposed for how CGG repeats contribute to the pathogenesis of PM conditions, including FXTAS. First, the cotranscriptional R-loop formation, which compromises genomic stability and triggers a DNA-damage response that can activate inflammatory cascades [<a href="#B116-cells-12-02330" class="html-bibr">116</a>,<a href="#B117-cells-12-02330" class="html-bibr">117</a>]. Second, CGG-repeat RNAs can elicit a gain-of-function toxicity through RNA gelation into nuclear foci and sequestration of various rCGG-repeat-binding proteins, leading to their functional depletion [<a href="#B25-cells-12-02330" class="html-bibr">25</a>,<a href="#B26-cells-12-02330" class="html-bibr">26</a>]. Third, repeat-associated non-AUG (RAN)-initiated translation generates potentially toxic proteins that accumulate within intranuclear neuronal inclusions in FXTAS patients. The relative contribution from each mechanism to downstream sequelae, such as mitochondrial dysfunction and neuronal death, and their potential synergies in disease pathogenesis, are areas of ongoing research in the field. Adapted from [<a href="#B118-cells-12-02330" class="html-bibr">118</a>].</p>
Full article ">Figure 2
<p>FXPAC involvement across the lifespan.</p>
Full article ">Figure 3
<p>(<b>a</b>–<b>c</b>) H&amp;E and ubiquitin staining (brown). Inclusions in astrocytes (<b>a</b>), neurons (<b>b</b>), and Purkinje cells (<b>c</b>); (<b>d</b>) H&amp;E. White-matter disease in cerebellum; (<b>e</b>) Cortical atrophy and venrticulomegalia; (<b>f</b>) Perl’s staining. Iron deposition in capillaries; (<b>g</b>,<b>h</b>) Iba1 staining. Activated microglia; (<b>i</b>,<b>k</b>) GFAP staining. Activated astrocytes; (<b>j</b>) Iba1 staining. Senescent microglia; (<b>l</b>) H&amp;E. Microbleeding; (<b>m</b>,<b>n</b>) H&amp;E and ubiquitin staining (brown). Inclusions in endothelial cells. Arrows and asterisks are indicating the pathology of interest. Arrowhead in (<b>c</b>) points to nucleolus.</p>
Full article ">Figure 4
<p>Lived experience perspective. Results from an anonymous Survey Monkey questionnaire distributed via email to the member based of Fragile X New Zealand (<span class="html-italic">n</span> = 38). (<b>a</b>) Binary data (blue = yes; gray = no); (<b>b</b>) Preferred terminology by survey respondents; (<b>c</b>,<b>d</b>). Example quotes in response to query about (<b>c</b>) what is helpful in research and (<b>d</b>) what is unhelpful in research.</p>
Full article ">Figure 5
<p>Lived experience perspectives from member based of NFXF and FXAA. (<b>a</b>) Results from an anonymous Survey Monkey questionnaire distributed via email to the member based of NFXF and FXAA (<span class="html-italic">n</span> = 255); (<b>b</b>) Example quotes.</p>
Full article ">Figure 6
<p>Attendees at the 5th International Conference on <span class="html-italic">FMR1</span> Premutation.</p>
Full article ">
14 pages, 1088 KiB  
Review
Emerging Roles of Ubiquitination in Biomolecular Condensates
by Peigang Liang, Jiaqi Zhang and Bo Wang
Cells 2023, 12(18), 2329; https://doi.org/10.3390/cells12182329 - 21 Sep 2023
Cited by 1 | Viewed by 1824
Abstract
Biomolecular condensates are dynamic non-membrane-bound macromolecular high-order assemblies that participate in a growing list of cellular processes, such as transcription, the cell cycle, etc. Disturbed dynamics of biomolecular condensates are associated with many diseases, including cancer and neurodegeneration. Extensive efforts have been devoted [...] Read more.
Biomolecular condensates are dynamic non-membrane-bound macromolecular high-order assemblies that participate in a growing list of cellular processes, such as transcription, the cell cycle, etc. Disturbed dynamics of biomolecular condensates are associated with many diseases, including cancer and neurodegeneration. Extensive efforts have been devoted to uncovering the molecular and biochemical grammar governing the dynamics of biomolecular condensates and establishing the critical roles of protein posttranslational modifications (PTMs) in this process. Here, we summarize the regulatory roles of ubiquitination (a major form of cellular PTM) in the dynamics of biomolecular condensates. We propose that these regulatory mechanisms can be harnessed to combat many diseases. Full article
(This article belongs to the Special Issue Advances in Ubiquitination and Deubiquitination Research)
Show Figures

Figure 1

Figure 1
<p>Stress-granule dynamics are tightly regulated by ubiquitination. Cells under stress inhibit the translation initiation of mRNAs, leading to stress granules’ formation that involves the phase separation of RBPs (such as G3BP1, T-cell intracellular antigen 1 (TIA-1), ELAV-like protein 1 (HuR), TIA1-related protein (TIAR) etc.) and mRNAs. Both the assembly and disassembly of stress granules are regulated by PTMs, including ubiquitination. The disassembly of stress granules is enhanced through the K63-linked ubiquitination of G3BP1, subsequently fostering the interaction between ubiquitin chains and VCP. Mutations in RBPs or prolonged stress cause the transition of stress granules into pathological aggregates. Ubiquitination governs the clearance of these aggregates, with K48-linked ubiquitin chains directing degradation through the ubiquitin-proteasome system (UPS), whereas K63-linked ubiquitination is associated with the autophagic degradation pathways. Created with BioRender.com.</p>
Full article ">Figure 2
<p>The process of autophagic cargo segregation is driven by poly-ubiquitin chain-induced p62 phase separation. p62 protein forms oligomers via the PB1 domain and also binds to ubiquitin through the UBA domain. Upon reaching the threshold concentration required for LLPS, the formation of p62 condensates occurs. Subsequently, other client proteins are recruited to these p62 condensates that further mature and ultimately are broken down through autophagy. Created with BioRender.com.</p>
Full article ">
13 pages, 2693 KiB  
Article
Novel Variant in CEP250 Causes Protein Mislocalization and Leads to Nonsyndromic Autosomal Recessive Type of Progressive Hearing Loss
by Minjin Kang, Jung Ah Kim, Mee Hyun Song, Sun Young Joo, Se Jin Kim, Seung Hyun Jang, Ho Lee, Je Kyung Seong, Jae Young Choi, Heon Yung Gee and Jinsei Jung
Cells 2023, 12(18), 2328; https://doi.org/10.3390/cells12182328 - 21 Sep 2023
Cited by 2 | Viewed by 1283
Abstract
Genetic hearing loss is the most common hereditary sensorial disorder. Though more than 120 genes associated with deafness have been identified, unveiled causative genes and variants of diverse types of hearing loss remain. Herein, we identified a novel nonsense homozygous variant in CEP250 [...] Read more.
Genetic hearing loss is the most common hereditary sensorial disorder. Though more than 120 genes associated with deafness have been identified, unveiled causative genes and variants of diverse types of hearing loss remain. Herein, we identified a novel nonsense homozygous variant in CEP250 (c.3511C>T; p.Gln1171Ter) among the family members with progressive moderate sensorineural hearing loss in nonsyndromic autosomal recessive type but without retinal degeneration. CEP250 encodes C-Nap1 protein belonging to the CEP protein family, comprising 30 proteins that play roles in centrosome aggregation and cell cycle progression. The nonsense variant in CEP250 led to the early truncating protein of C-Nap1, which hindered centrosome localization; heterologous expression of CEP250 (c.3511C>T) in NIH3T3 cells within cilia expression condition revealed that the truncating C-Nap1 (p.Gln1171Ter) was not localized at the centrosome but was dispersed in the cytosol. In the murine adult cochlea, Cep250 was expressed in the inner and outer hair cells. Knockout mice of Cep250 showed significant hair cell degeneration and progressive hearing loss in auditory brainstem response. In conclusion, a nonsense variant in CEP250 results in a deficit of centrosome localization and hair cell degeneration in the cochlea, which is associated with the progression of hearing loss in humans and mice. Full article
(This article belongs to the Special Issue Cell Death in Health and Disease)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Discovery of new causative gene CEP250 mutations in patients with hearing loss through whole-exome sequencing methods. (<b>A</b>) Whole-exome sequencing analysis in the YUHL cohort was performed and <span class="html-italic">CEP250</span> was found as a novel deafness gene in YUHL 251 family. (<b>B</b>) Pedigree and found mutations of YUHL251 (−21, 42-year-old female; −22, 40-year-old female; −23, 35-year-old female. (<b>C</b>) Hearing tests showed high-frequency sensorineural hearing loss in YUHL251-21. Red line (circle; air conduction, [ or &lt;; bone conduction] refers to the right ear threshold and blue line (cross; air threshold,] or &gt;; bone conduction) refers to the left ear threshold. (<b>D</b>) The autosomal recessive <span class="html-italic">CEP250</span> c.3511C&gt;T, p.Gln1171Ter gene mutation was discovered through whole-exome sequencing. (<b>E</b>) Evolutionary conservation of altered amino acid residues (<span class="html-italic">Homo sapiens</span> (human), <span class="html-italic">Rattus norvegicus</span> (rat), <span class="html-italic">Mus musculus</span> (mouse), <span class="html-italic">Gallus gallus</span> (chicken), <span class="html-italic">Xenopus tropicalis</span> (frog)). (<b>F</b>) Schematic diagram of pathogenic variants in <span class="html-italic">CEP250</span>. Abbreviations: YUHL, Yonsei University Hearing Loss. Asterisk (*) denotes a stop codon.</p>
Full article ">Figure 2
<p>Expression of p.Gln1171Ter C-Nap1 in NIH3T3 cell line. (<b>A</b>,<b>B</b>) Immunoblotting analysis using Myc (<b>A</b>) and CEP250 (<b>B</b>) antibodies. <span class="html-italic">CEP250</span> WT was expressed at 255 kDa (full length) and ~160 kDa (short form). CEP250 p.Gln1171Ter was detected at ~150 kDa, corresponding to the early truncated protein. (<b>C</b>) Immunocytochemistry was performed in NIH3T3 cells transfected with wild-type CEP250 and p.Gln1171Ter. It was stained using γ-tubulin (red), Dapi (blue), and C-Myc (green) antibodies in NIH3T3 cells. White arrows point to centrosome.</p>
Full article ">Figure 3
<p>Cilia expression in NIH3T3 cells transfected with CEP250 and p.Gln1171Ter variant in immunostaining. Staining with Acetyl-⍺-tubulin (green), C-Myc (red, CEP250) antibody to confirm whether the expression of CEP250 p.Gln1171Ter mutation affects the production and growth of primary cilia. Both CEP250 WT and p.Gln1171Ter mutations confirmed the production of primary cilia. Green arrows point to primary cilia.</p>
Full article ">Figure 4
<p>Expression and localization of Cep250 in the mouse cochlea. (<b>A</b>,<b>B</b>) <span class="html-italic">Cep250</span> transcriptional expression in the organ of Corti at P1 (<b>A</b>) and P30 (<b>B</b>). <span class="html-italic">Myo7a</span> and <span class="html-italic">Atoh1</span> are depicted as hair cell markers. HC, hair cell; SC, supporting cell. (<b>C</b>) <span class="html-italic">Cep250</span> transcript expression in the spiral ganglion neuron at different ages. <span class="html-italic">Calb2</span>, <span class="html-italic">Calb1</span>, and <span class="html-italic">Lypd1</span> are markers for type I, II, and III neurons, respectively. SGN, spiral ganglion neuron. (<b>D</b>,<b>E</b>) Immunohistochemistry was performed for wildtype and <span class="html-italic">Cep250</span> knockout mice (P30). Cep250 (red) was expressed in the inner hair cells (IHCs), outer hair cells, and pillar cells in the organ of Corti of wildtype mice but not <span class="html-italic">Cep250</span> knockout mice (<b>D</b>). Cep250 (red) is expressed in the spiral ganglion (dashed area) of wildtype mice but not in that of the <span class="html-italic">Cep250</span> knockout mice (<b>E</b>). Dapi (blue) staining. WT, wildtype; Het, heterozygous <span class="html-italic">Cep250</span> knockout mice; KO, homozygous <span class="html-italic">Cep250</span> knockout mice.</p>
Full article ">Figure 5
<p>Auditory brainstem response (ABR) test in the <span class="html-italic">Cep250</span> knockout mice. ABR thresholds were measured at the stimuli of click and tone sounds. The number of <span class="html-italic">Cep250</span> (+/+), (+/−), and (−/−) mice was 5, 11, and 5, respectively. ****, <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 6
<p>The cell survival rate in the <span class="html-italic">Cep250</span> knockout mice at the age of 17 weeks. (<b>A</b>) Whole-mount immunostaining with phalloidin (green), anti-Cep250 antibody (red), and Dapi (blue) was performed through the whole cochlear turn. Tonotopic frequencies are noted. (<b>B</b>) The basal turn of the <span class="html-italic">Cep250</span> knockout mouse showed a higher degeneration rate of the outer hair cell compared to that of WT mice. (<b>C</b>) Inner hair cell counts in apex, mid, and basal turns in CEP250 knockout mice. (<b>D</b>) Outer hair cell counts in apex, mid, and basal turns in <span class="html-italic">Cep250</span> knockout mice. IHC, inner hair cells; OHC, outer hair cells; ns, not significant; ****, <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">
15 pages, 1752 KiB  
Review
NLRP3 Inflammasome as a Potentially New Therapeutic Target of Mesenchymal Stem Cells and Their Exosomes in the Treatment of Inflammatory Eye Diseases
by Carl Randall Harrell, Valentin Djonov, Ana Antonijevic and Vladislav Volarevic
Cells 2023, 12(18), 2327; https://doi.org/10.3390/cells12182327 - 21 Sep 2023
Cited by 5 | Viewed by 1712
Abstract
Due to their potent immunoregulatory and angio-modulatory properties, mesenchymal stem cells (MSCs) and their exosomes (MSC-Exos) have emerged as potential game-changers in regenerative ophthalmology, particularly for the personalized treatment of inflammatory diseases. MSCs suppress detrimental immune responses in the eyes and alleviate ongoing [...] Read more.
Due to their potent immunoregulatory and angio-modulatory properties, mesenchymal stem cells (MSCs) and their exosomes (MSC-Exos) have emerged as potential game-changers in regenerative ophthalmology, particularly for the personalized treatment of inflammatory diseases. MSCs suppress detrimental immune responses in the eyes and alleviate ongoing inflammation in ocular tissues by modulating the phenotype and function of all immune cells that play pathogenic roles in the development and progression of inflammatory eye diseases. MSC-Exos, due to their nano-sized dimension and lipid envelope, easily bypass all barriers in the eyes and deliver MSC-sourced bioactive compounds directly to target cells. Although MSCs and their exosomes offer a novel approach to treating immune cell-driven eye diseases, further research is needed to optimize their therapeutic efficacy. A significant number of experimental studies is currently focused on the delineation of intracellular targets, which crucially contribute to the immunosuppressive and anti-inflammatory effects of MSCs and MSC-Exos. The activation of NLRP3 inflammasome induces programmed cell death of epithelial cells, induces the generation of inflammatory phenotypes in eye-infiltrated immune cells, and enhances the expression of adhesion molecules on ECs facilitating the recruitment of circulating leukocytes in injured and inflamed eyes. In this review article, we summarize current knowledge about signaling pathways that are responsible for NLRP3 inflammasome-driven intraocular inflammation and we emphasize molecular mechanisms that regulate MSC-based modulation of NLRP3-driven signaling in eye-infiltrated immune cells, providing evidence that NLRP3 inflammasome should be considered a potentially new therapeutic target for MSCs and MSC-Exo-based treatment of inflammatory eye diseases. Full article
(This article belongs to the Special Issue Updates on Mesenchymal Stem Cells-Derived Extracellular Vesicles)
Show Figures

Figure 1

Figure 1
<p>Molecular mechanisms and signaling pathways involved in the activation of NLRP3 inflammasome in eye-infiltrated immune cells. The activation of the NLRP3 inflammasome in immune cells is a tightly regulated, multi-step process. It requires 2 steps: priming and activation. The priming step of NLRP3 inflammasome activation is elicited by the activation of pattern recognition receptors (PRRs), including Toll-like receptors (TLR). These membrane-bound receptors are expressed on immune cells that recognize pathogen-associated molecular patterns (PAMPs) of microbial antigens or damage-associated molecular patterns (DAMPs) released from injured cells. Signals generated from activated TLR-2, TLR-4, TLR-5, and TLR-6 initiate phosphorylation and consequent activation of extracellular signal-regulated kinase (ERK), c-Jun N-terminal kinase (JNK), and p38 mitogen-activated protein kinase (MAPK), which, in turn, phosphorylate c-Jun, ATF-2, and CREB transcription factors that bind to the promoter regions of NLRP3 and pro-IL-1β, leading to their transcriptional upregulation.</p>
Full article ">Figure 2
<p>Molecular mechanisms responsible for NLRP3-dependent generation of detrimental immune response in inflamed eyes. Activation of NLRP3 inflammasome results in increased production of IL-1β and IL-18 in tissue-infiltrated macrophages and neutrophils. IL-1β and IL-18 induce massive recruitment of circulating leukocytes in inflamed tissues, increased expansion of inflammatory Th1 and Th17 lymphocytes, enhanced production of inflammatory cytokines and chemokines in neutrophils and macrophages, attenuated proliferation of immunosuppressive T regulatory cells (Tregs), and increased antibody production in plasma cells.</p>
Full article ">Figure 3
<p>The beneficial effects of MSCs and MSC-derived exosomes in the attenuation of NLRP3-driven eye injury and inflammation. MSCs secrete a wide array of immunosuppressive cytokines and immunoregulatory factors, such as IL-10, transforming growth factor-β (TGF-β), prostaglandin E2 (PGE2), and stanniocalcin (STC)-1, which efficiently suppress NLRP3 inflammasome activation and inhibited production of IL-1β and IL-18 in eye-infiltrated macrophages. MSC-derived PGE2 was found mainly responsible for the MSC-dependent generation of tolerogenic phenotype in DCs, which was followed by attenuated activation of naïve T cells. MSC-sourced IDO and Kynurenine were considered mainly responsible for the MSC-dependent regulation of T/NK/NKT cells’ phenotype and function.</p>
Full article ">
20 pages, 7512 KiB  
Article
Human Preadipocytes Differentiated under Hypoxia following PCB126 Exposure during Proliferation: Effects on Differentiation, Glucose Uptake and Adipokine Profile
by Zeinab El Amine, Jean-François Mauger and Pascal Imbeault
Cells 2023, 12(18), 2326; https://doi.org/10.3390/cells12182326 - 21 Sep 2023
Cited by 1 | Viewed by 1187
Abstract
Persistent organic pollutants (POPs) accumulation and hypoxia are two factors proposed to adversely alter adipose tissue (AT) functions in the context of excess adiposity. Studies have shown that preadipocytes exposure to dioxin and dioxin-like POPs have the greatest deleterious impact on rodent and [...] Read more.
Persistent organic pollutants (POPs) accumulation and hypoxia are two factors proposed to adversely alter adipose tissue (AT) functions in the context of excess adiposity. Studies have shown that preadipocytes exposure to dioxin and dioxin-like POPs have the greatest deleterious impact on rodent and immortalized human preadipocyte differentiation, but evidence on human preadipocytes is lacking. Additionally, hypoxia is known to strongly interfere with the dioxin-response pathway. Therefore, we tested the effects of pre-differentiation polychlorinated biphenyl (PCB)126 exposure at 10 µM for 3 days and subsequent differentiation under hypoxia on human subcutaneous adipocytes (hSA) differentiation, glucose uptake and expression of selected metabolism- and inflammation-related genes. Pre-differentiation PCB126 exposure lowered the adenosine triphosphate (ATP) content, glucose uptake and leptin expression of mature adipocytes but had limited effects on differentiation under normoxia (21% O2). Under hypoxia (3% O2), preadipocytes ability to differentiate was significantly reduced as reflected by significant decreased lipid accumulation and downregulation of key adipocyte genes such as peroxisome proliferator-activated receptor gamma (PPARγ) and adiponectin. Hypoxia increased glucose uptake and glucose transporter 1 (GLUT1) expression but abolished the adipocytes insulin response and GLUT4 expression. The expression of pro-inflammatory adipokine interleukin-6 (IL-6) was slightly increased by both PCB126 and hypoxia, while IL-8 expression was significantly increased only following the PCB126-hypoxia sequence. These observations suggest that PCB126 does not affect human preadipocyte differentiation, but does affect the subsequent adipocytes population, as reflected by lower ATP levels and absolute glucose uptake. On the other hand, PCB126 and hypoxia exert additive effects on AT inflammation, an important player in the development of chronic diseases such as type 2 diabetes and cardiovascular diseases. Full article
(This article belongs to the Special Issue The Adipose Tissue: From “Cinderella” to “Lion King” Organ)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Differentiating preadipocytes exposed to either DMSO (vector) or 10 µM PCB126 during proliferation (3 days). The top 4 pictures were taken after the first 24 h of differentiation under either 21% or 3% O<sub>2</sub>. Bottom 4 pictures were taken after 14 days of differentiation under either 21% or 3% O<sub>2</sub>. All pictures were taken using a light microscope at 5× magnification.</p>
Full article ">Figure 2
<p>Effect of 10 µM PCB126 and hypoxia on cellular ATP content and media LDH activity. Top panels: Cell ATP content relative to control condition (DMSO, normoxia) 24 h following induction. Mid panels: Cell ATP content relative to control condition after 14 days of differentiation. Bottom panels: Media LDH activity measured after 14 days of differentiation. Main panels (left-hand side) illustrate the complete model and minor panels on the right-hand side summarize the separate main effects of PCB126 (upper minor panel) and hypoxia (lower minor panel). Main effect of PCB126 or hypoxia at * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 or *** <span class="html-italic">p</span> &lt; 0.001. Results are expressed as mean ± SEM of three separate experiments for each treatment group.</p>
Full article ">Figure 3
<p>Effect of PCB126 and hypoxia on VEGFA (top panels) and CYP1A1 (bottom panels) expression in human adipocytes. Preadipocytes were treated with DMSO (vector) or 10 µM PCB126 for 3 days (pre-differentiation exposure), then PCB126 was removed from media and differentiation was induced. At this point, cells were subjected to 21 or 3% O<sub>2</sub> for 14 days. VEGFA and CYP1A1 gene expression was measured using RT-qPCR. Main panels (left-hand side) illustrate the complete model and minor panels on the right-hand side summarize the separate main effects of PCB126 (upper minor panel) and hypoxia (lower minor panel). Bars not sharing a common letter are statistically different at <span class="html-italic">p</span> &lt; 0.001. Main effect of PCB126 or hypoxia at *** <span class="html-italic">p</span> &lt; 0.001. Results are expressed as mean ± SEM of three separate experiments for each treatment group.</p>
Full article ">Figure 4
<p>Effect of PCB126 and hypoxia on PPARγ (top panels), adiponectin (middle panels) and leptin (bottom panels) expression in human adipocytes. Preadipocytes were treated with DMSO (vector) or 10 µM PCB126 for 3 days (pre-differentiation exposure), then PCB126 was removed from media and differentiation was induced. At this point, cells were subjected to 21 or 3% O<sub>2</sub> for 14 days. Gene expression of PPARγ, adiponectin and leptin was measured using RT-qPCR. Main panels (left-hand side) illustrate the complete model and minor panels on the right-hand side summarize the separate main effects of PCB126 (upper minor panel) and hypoxia (lower minor panel). Bars not sharing a common letter are statistically different at <span class="html-italic">p</span> &lt; 0.001. Main effect of PCB126 or hypoxia at *** <span class="html-italic">p</span> &lt; 0.001. Results are expressed as mean ± SEM of three separate experiments for each treatment group.</p>
Full article ">Figure 5
<p>Effect of PCB126 and hypoxia on Il-6 (top panels) and IL-8 (bottom panels) expression in human adipocytes. Preadipocytes were treated with DMSO (vector) or 10 µM PCB126 for 3 days (pre-differentiation exposure), then PCB126 was removed from media and differentiation was induced. At this point, cells were subjected to 21 or 3% O<sub>2</sub> for 14 days. Gene expression of IL-6 and IL-8 was measured using RT-qPCR. Main panels (left-hand side) illustrate the complete model and minor panels on the right-hand side summarize the separate main effects of PCB126 (upper minor panel) and hypoxia (lower minor panel). Bars not sharing a common letter are statistically different at <span class="html-italic">p</span> &lt; 0.05. Main effect of PCB126 or hypoxia at ** <span class="html-italic">p</span> &lt; 0.01 or *** <span class="html-italic">p</span> &lt; 0.001. Results are expressed as mean ± SEM of three separate experiments for each treatment group.</p>
Full article ">Figure 6
<p>Relative TG content of human differentiated preadipocytes exposed to 10 µM PCB126 or DMSO for 3 days during proliferation followed by differentiation for 14 days under either 21% or 3% O<sub>2</sub>. Top panels illustrate relative TG content compared to DMSO/normoxia. Bottom panels illustrate TG content normalized to ATP, relative to DMSO/normoxia. Main panels (left-hand side) illustrate the complete model and minor panels on the right-hand side summarize the separate main effects of PCB126 (upper minor panel) and hypoxia (lower minor panel). Main effect of PCB126 or hypoxia at *** <span class="html-italic">p</span> &lt; 0.001. Results are expressed as mean ± SEM of three separate experiments for each treatment group.</p>
Full article ">Figure 7
<p>Early effect of PCB126 and hypoxia on basal glucose uptake in human adipocytes. Preadipocytes were treated with DMSO (vector) or 10 µM PCB126 for 3 days (pre-differentiation exposure), then PCB126 was removed from media and differentiation was induced. At this point, cells were subjected to 21 or 3% O<sub>2</sub> for 24 h, after which glucose uptake was assessed. Main panels (left-hand side) illustrate the complete model and minor panels on the right-hand side summarize the separate main effects of PCB126 (upper minor panel) and hypoxia (lower minor panel). Main effect of PCB126 or hypoxia at *** <span class="html-italic">p</span> &lt; 0.001. Results are expressed as mean ± SEM of three separate experiments for each treatment group.</p>
Full article ">Figure 8
<p>Effect of PCB126 and hypoxia on basal glucose uptake in human differentiated preadipocytes on day 14. Preadipocytes were treated with DMSO (vector) or 10 µM PCB126 for 3 days (pre-differentiation exposure), then PCB126 was removed from media and differentiation was induced. At this point, cells were subjected to 21 or 3% O<sub>2</sub> for 14 days. Glucose uptake was assessed at the end of the 14-day differentiation period. Main panels (left-hand side) illustrate the complete model and minor panels on the right-hand side summarize the separate main effects of PCB126 (upper minor panel) and hypoxia (lower minor panel). Main effect of PCB126 or hypoxia at *** <span class="html-italic">p</span> &lt; 0.001. Results are expressed as mean ± SEM of three separate experiments for each treatment group.</p>
Full article ">Figure 9
<p>Effect of PCB126 and hypoxia on insulin-stimulated glucose uptake in human adipocytes. Preadipocytes were treated with DMSO (vector) or 10 µM PCB126 for 3 days (pre-differentiation exposure), then PCB126 was removed from media and differentiation was induced. At this point, cells were subjected to 21 or 3% O<sub>2</sub> for 14 days and measurement was done on day 14 post induction. Main panels (left-hand side) illustrate the complete model and minor panels on the right-hand side summarize the separate main effects of PCB126 (upper minor panel) and hypoxia (lower minor panel). Main effect of PCB126 or hypoxia at *** <span class="html-italic">p</span> &lt; 0.001. Results are expressed as mean ± SEM of three separate experiments for each treatment group.</p>
Full article ">Figure 10
<p>Effect of PCB126 and hypoxia on GLUT1 (top panel) and GLUT4 (bottom panel) expression in human adipocytes. Preadipocytes were treated with DMSO (vector) or 10 µM PCB126 for 3 days (pre-differentiation exposure), then PCB126 was removed from media and differentiation was induced. At this point, cells were subjected to 21 or 3% O<sub>2</sub> for 14 days. Main panels (left-hand side) illustrate the complete model and minor panels on the right-hand side summarize the separate main effects of PCB126 (upper minor panel) and hypoxia (lower minor panel). Main effect of PCB126 or hypoxia at *** <span class="html-italic">p</span> &lt; 0.001. Results are expressed as mean ± SEM of three separate experiments for each treatment group.</p>
Full article ">
15 pages, 2989 KiB  
Article
Establishment and Characterization of Continuous Satellite Muscle Cells from Olive Flounder (Paralichthys olivaceus): Isolation, Culture Conditions, and Myogenic Protein Expression
by Sathish Krishnan, Selvakumari Ulagesan, Josel Cadangin, Ji-Hye Lee, Taek-Jeong Nam and Youn-Hee Choi
Cells 2023, 12(18), 2325; https://doi.org/10.3390/cells12182325 - 21 Sep 2023
Cited by 1 | Viewed by 1791
Abstract
Olive flounder (Paralichthys olivaceus) muscle satellite cells (OFMCs) were obtained by enzymatic primary cell isolation and the explant method. Enzymatic isolation yielded cells that reached 80% confluence within 8 days, compared to 15 days for the explant method. Optimal OFMC growth [...] Read more.
Olive flounder (Paralichthys olivaceus) muscle satellite cells (OFMCs) were obtained by enzymatic primary cell isolation and the explant method. Enzymatic isolation yielded cells that reached 80% confluence within 8 days, compared to 15 days for the explant method. Optimal OFMC growth was observed in 20% fetal bovine serum at 28 °C with 0.8 mM CaCl2 and the basic fibroblast growth factor (BFGF) to enhance cell growth. OFMCs have become permanent cell lines through the spontaneous immortalization crisis at the 20th passage. Olive flounder skeletal muscle myoblasts were induced into a mitogen-poor medium containing 2% horse serum for differentiation; they fused to form multinucleate myotubes. The results indicated complete differentiation of myoblasts into myotubes; we also detected the expression of the myogenic regulatory factors myoD, myogenin, and desmin. Upregulation (Myogenin, desmin) and downregulation (MyoD) of muscle regulation factors confirmed the differentiation in OFMCs. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) <span class="html-italic">Paralichthys olivaceus</span> (red color rectangle indicates the place where muscle tissue was collected). (<b>b</b>) Live and dead cells of olive flounder muscle cells stained with calcine AM and propidium iodide. Scale bar: 200 µm.</p>
Full article ">Figure 2
<p>Phase contrast micrographs of olive flounder muscle satellite cells attachment and proliferation from days 0 to 6. Scale bar: 100 µm.</p>
Full article ">Figure 3
<p>Explant primary cell isolation. Radial movement of cells from explant tissue. Scale bar—100 µm.</p>
Full article ">Figure 4
<p>Optimal growth conditions for OFMCs. (<b>a</b>) FBS concentration, (<b>b</b>) temperature, (<b>c</b>) salt concentration, and (<b>d</b>) growth factors. ns—non-significant; *—significance (<span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 5
<p>Four representative images of spontaneous immortalization of OFMC cells. (<b>a</b>) Large-scale apoptosis and senescence occurred in the OFMC cells. Green circle—sporadic cluster; blue circle—detaching and defragmented cell. (<b>b</b>) The remaining cells initiated to overcome the crisis. Red arrow—sporadic cells started proliferating into myocytes. (<b>c</b>) Cells began to actively divide again. (<b>d</b>) The formation of a confluent cell monolayer Scale bar: 100 µm.</p>
Full article ">Figure 6
<p>Differentiation of OFMCs into myotubes. (<b>a</b>) Representative images of undifferentiated (myoblasts) and differentiated cells (myotubes) incubated with 2% horse serum. Magnification, 10×; Scale bar—100 μm. (<b>b</b>) MyoD, myogenin, and desmin protein expression levels in undifferentiated and differentiated cells. Results are means ± standard deviation of three independent experiments. ** <span class="html-italic">p</span> &lt; 0.05 vs. corresponding control group (D-0). CON, control; D, day; UD, undifferentiation, D-0. (<b>c</b>) FITC-phalloidin and DAPI were used to visualize actin and nucleus staining, and the colocalization of phalloidin and DAPI staining is indicated in the merged images. Magnification 10×; scale bar—100 μm. green- cytoskeleton staining, blue-nucleus staining.</p>
Full article ">
21 pages, 1044 KiB  
Review
Regulatory Mechanisms That Guide the Fetal to Postnatal Transition of Cardiomyocytes
by Patrick G. Burgon, Jonathan J. Weldrick, Omar Mohamed Sayed Ahmed Talab, Muhammad Nadeer, Michail Nomikos and Lynn A. Megeney
Cells 2023, 12(18), 2324; https://doi.org/10.3390/cells12182324 - 21 Sep 2023
Cited by 1 | Viewed by 1673
Abstract
Heart disease remains a global leading cause of death and disability, necessitating a comprehensive understanding of the heart’s development, repair, and dysfunction. This review surveys recent discoveries that explore the developmental transition of proliferative fetal cardiomyocytes into hypertrophic postnatal cardiomyocytes, a process yet [...] Read more.
Heart disease remains a global leading cause of death and disability, necessitating a comprehensive understanding of the heart’s development, repair, and dysfunction. This review surveys recent discoveries that explore the developmental transition of proliferative fetal cardiomyocytes into hypertrophic postnatal cardiomyocytes, a process yet to be well-defined. This transition is key to the heart’s growth and has promising therapeutic potential, particularly for congenital or acquired heart damage, such as myocardial infarctions. Although significant progress has been made, much work is needed to unravel the complex interplay of signaling pathways that regulate cardiomyocyte proliferation and hypertrophy. This review provides a detailed perspective for future research directions aimed at the potential therapeutic harnessing of the perinatal heart transitions. Full article
Show Figures

Figure 1

Figure 1
<p>The transitional program: Reprogramming embryonic cardiomyocytes into postnatal cardiomyocytes. (<b>A</b>) Embryonic heart growth is primarily due to the high proliferative capacity of fetal cardiomyocytes, whereas postnatal heart growth is associated with binucleated cardiomyocyte hypertrophy. (<b>B</b>) Linking the fetal cardiogenomic program (green) to the postnatal cardiogenomic program (red). The transitional program (yellow) mediates the passage of proliferative fetal cardiomyocytes to exit the cell cycle prematurely to form binucleated hypertrophic cardiomyocytes.</p>
Full article ">Figure 2
<p>The transition of mammalian cardiomyocyte (CM) growth signaling pathways that convert proliferative fetal cardiomyocytes into hypertrophic adult cardiomyocytes. Top panel. The Hippo pathway is the primary fetal cardiomyogenic hyperplasia program (green) that regulates fetal CM number prior to birth (0 d). Birth triggers the perinatal transition program (yellow) that facilitates the conversion of fetal CM into adult hypertrophic CMs (red). MicroRNAs are excellent candidate molecules that propagate the CM perinatal transition. The mTOR pathway is the principal pathway that controls CM physiological hypertrophy. 0 d = time of birth and 14 d = 14 days post-birth. Bottom Panel: Highlights the differences between fetal (green) and adult (red) cardiomyocytes.</p>
Full article ">
16 pages, 2284 KiB  
Review
Neuroimmune Interactions in Fetal Alcohol Spectrum Disorders: Potential Therapeutic Targets and Intervention Strategies
by Sayani Mukherjee, Prashant Tarale and Dipak K. Sarkar
Cells 2023, 12(18), 2323; https://doi.org/10.3390/cells12182323 - 21 Sep 2023
Viewed by 1897
Abstract
Fetal alcohol spectrum disorders (FASD) are a set of abnormalities caused by prenatal exposure to ethanol and are characterized by developmental defects in the brain that lead to various overt and non-overt physiological abnormalities. Growing evidence suggests that in utero alcohol exposure induces [...] Read more.
Fetal alcohol spectrum disorders (FASD) are a set of abnormalities caused by prenatal exposure to ethanol and are characterized by developmental defects in the brain that lead to various overt and non-overt physiological abnormalities. Growing evidence suggests that in utero alcohol exposure induces functional and structural abnormalities in gliogenesis and neuron–glia interactions, suggesting a possible role of glial cell pathologies in the development of FASD. However, the molecular mechanisms of neuron–glia interactions that lead to the development of FASD are not clearly understood. In this review, we discuss glial cell pathologies with a particular emphasis on microglia, primary resident immune cells in the brain. Additionally, we examine the involvement of several neuroimmune molecules released by glial cells, their signaling pathways, and epigenetic mechanisms responsible for FASD-related alteration in brain functions. Growing evidence suggests that extracellular vesicles (EVs) play a crucial role in the communication between cells via transporting bioactive cargo from one cell to the other. This review emphasizes the role of EVs in the context of neuron–glia interactions during prenatal alcohol exposure. Finally, some potential applications involving nutritional, pharmacological, cell-based, and exosome-based therapies in the treatment of FASD are discussed. Full article
(This article belongs to the Special Issue Alcohol and Neuroimmunology)
Show Figures

Figure 1

Figure 1
<p>A schematic diagram showing ethanol-induced alterations in glial cell functions. Early ethanol exposure during development causes a shift from resting microglia and astrocytes to activated microglia and astrocyte phenotypes. These inflammatory glial phenotypes secrete extracellular vesicles containing inflammatory molecules, including cytokines, chemokines, and miRNAs, which are detrimental to neurons associated with regulation of stress-axis function. Ethanol exposure also alters oligodendrocyte functions, which has long-term effects on neuronal myelination, affecting synaptic plasticity. These events may result in FASD-associated cognitive, behavioral, and motor impairments.</p>
Full article ">Figure 2
<p>A schematic diagram showing ethanol-induced changes in microglial cell functions. Early ethanol exposure during the developmental period causes phenotypic shift from resting microglia to activated microglia (M1). The most common signaling pathways that are upregulated during alcohol-induced microglial priming are TLR2/4, chemokine, cytokine, complement, MOR, NLRP3, NF-kB/TNF-α, and ROS signaling. The most common epigenetic alterations include decrease in DNMT1/3A, SIRT1, miR153, and MeCP2. All these altered signaling mechanisms trigger the release of neurotoxic factors from primed microglia and cause apoptotic death of neurons.</p>
Full article ">Figure 3
<p>A schematic diagram showing the early intervention, therapeutic, and management strategies to improve the cognitive function in FASD. Extracellular vesicles may be considered as potential biomarkers as well as therapeutic vesicles for FASD.</p>
Full article ">
22 pages, 7521 KiB  
Article
TRPV1 Channels Are New Players in the Reticulum–Mitochondria Ca2+ Coupling in a Rat Cardiomyoblast Cell Line
by Nolwenn Tessier, Mallory Ducrozet, Maya Dia, Sally Badawi, Christophe Chouabe, Claire Crola Da Silva, Michel Ovize, Gabriel Bidaux, Fabien Van Coppenolle and Sylvie Ducreux
Cells 2023, 12(18), 2322; https://doi.org/10.3390/cells12182322 - 20 Sep 2023
Cited by 5 | Viewed by 2060
Abstract
The Ca2+ release in microdomains formed by intercompartmental contacts, such as mitochondria-associated endoplasmic reticulum membranes (MAMs), encodes a signal that contributes to Ca2+ homeostasis and cell fate control. However, the composition and function of MAMs remain to be fully defined. Here, [...] Read more.
The Ca2+ release in microdomains formed by intercompartmental contacts, such as mitochondria-associated endoplasmic reticulum membranes (MAMs), encodes a signal that contributes to Ca2+ homeostasis and cell fate control. However, the composition and function of MAMs remain to be fully defined. Here, we focused on the transient receptor potential vanilloid 1 (TRPV1), a Ca2+-permeable ion channel and a polymodal nociceptor. We found TRPV1 channels in the reticular membrane, including some at MAMs, in a rat cardiomyoblast cell line (SV40-transformed H9c2) by Western blotting, immunostaining, cell fractionation, and proximity ligation assay. We used chemical and genetic probes to perform Ca2+ imaging in four cellular compartments: the endoplasmic reticulum (ER), cytoplasm, mitochondrial matrix, and mitochondrial surface. Our results showed that the ER Ca2+ released through TRPV1 channels is detected at the mitochondrial outer membrane and transferred to the mitochondria. Finally, we observed that prolonged TRPV1 modulation for 30 min alters the intracellular Ca2+ equilibrium and influences the MAM structure or the hypoxia/reoxygenation-induced cell death. Thus, our study provides the first evidence that TRPV1 channels contribute to MAM Ca2+ exchanges. Full article
Show Figures

Figure 1

Figure 1
<p>Expression and intracellular localization of TRPV1 in SV40-transformed H9c2 cells. (<b>A</b>) RT-PCR obtained RNA expression of TRPV1 in SV40-transformed H9c2 cells and mouse tissues (brain, dorsal root ganglion (DRG), heart). (<b>B</b>) Immunoblot analysis of TRPV1, voltage-dependent anion channel (VDAC), and glucose-regulated protein 75 (GRP75) on subcellular fractions from SV40-transformed H9c2 cells. (<b>C</b>–<b>H</b>) Double-staining immunofluorescence was applied to SV40-transformed H9c2 cells using antibodies against TRPV1 ((<b>C</b>,<b>F</b>); green signal) and IP3R (inositol 1,4,5-trisphosphate receptor; (<b>D</b>); red signal) or GRP75 ((<b>G</b>); red signal). Both color channels were merged to demonstrate co-distribution (yellow signal) of both immunofluorescence staining signals (<b>E</b>,<b>H</b>). Scale bar = 10 µm. (<b>I</b>–<b>N</b>) Representative confocal microscopy images of the TRPV1-IP3R (<b>I</b>), TRPV1-VDAC (<b>J</b>), TRPV1-GRP75 (<b>K</b>), TRPV1-GRIM19 (genes associated with retinoid–IFN-induced mortality-19; (<b>L</b>)), TRPV1-ANT (adenine nucleotide translocase; (<b>M</b>)), and TRPV1-CYPF (cyclophilin F; (<b>N</b>)) interactions in SV40-transformed H9c2 cells by proximity ligation assay. Scale bar = 50 µm.</p>
Full article ">Figure 2
<p>Acute effect of TRPV1 modulation on Ca<sup>2+</sup> homeostasis. (<b>A</b>,<b>B</b>) Reticular Ca<sup>2+</sup> concentration ([Ca<sup>2+</sup>]<sub>r</sub>). (<b>A</b>) Time traces showing [Ca<sup>2+</sup>]<sub>r</sub> measured with erGAP1 probe during ionomycin (1 µM; black line), resiniferatoxin (RTX) stimulation (10 µM; pink line), or 5′-iodoresiniferatoxin (iRTX) stimulation (10 µM; blue line). (<b>B</b>) Scatter plots representing reticular Ca<sup>2+</sup> content assessed by ionomycin (1 µM; black; n = 156), RTX (10 µM; pink; n = 69), or iRTX (10 µM; blue; n = 42) stimulation. (<b>C</b>,<b>D</b>) Cytosolic Ca<sup>2+</sup> concentration ([Ca<sup>2+</sup>]<sub>c</sub>). (<b>C</b>) Time traces showing [Ca<sup>2+</sup>]<sub>c</sub> measured with Fura-2 AM probe during ionomycin (1 µM; black line), RTX (10 µM; pink line), or iRTX (10 µM; blue line) stimulation. (<b>D</b>) Scatter plots representing cytosolic Ca<sup>2+</sup> content assessed by ionomycin (1 µM; black; n = 156), RTX (10 µM; pink; n = 105), or iRTX (10 µM; blue; n = 74) stimulation. (<b>E</b>–<b>G</b>) Mitochondrial Ca<sup>2+</sup> concentration ([Ca<sup>2+</sup>]<sub>m</sub>). (<b>E</b>) Time traces showing [Ca<sup>2+</sup>]<sub>m</sub> measured with CMV-mito-R-GECO1 probe during NaATP (100 µM; black line), RTX (10 µM; pink line), or iRTX (10 µM; blue line) stimulation. (<b>F</b>) Scatter plots representing mitochondrial Ca<sup>2+</sup> content assessed by NaATP (100 µM; black; n = 21), RTX (10 µM; pink; n = 32), or iRTX (10 µM; blue; n = 31) stimulation. (<b>G</b>) Scatter plots representing mitochondrial total Ca<sup>2+</sup> content assessed by ionomycin (1 µM) after NaATP (black; n = 21), RTX (pink; n = 32), or iRTX (blue; n = 31) stimulation. (<b>H</b>,<b>I</b>) Ca<sup>2+</sup> concentration in mitochondrial hot spots ([Ca<sup>2+</sup>]<sub>hot spots</sub>). (<b>H</b>) Time traces showing [Ca<sup>2+</sup>]<sub>hot spots</sub> measured with N33D3cpv probe during RTX (10 µM; pink line), iRTX (10 µM; blue line), or NaATP (100 µM; black line) stimulation. (<b>I</b>) Scatter plots representing Ca<sup>2+</sup> content in mitochondrial hot spots assessed by NaATP (100 µM; black; n = 58), RTX (10 µM; pink; n= 40), or iRTX (10 µM; blue; n = 43) stimulation. Data are from at least three independent experiments. Statistics: * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001; # <span class="html-italic">p</span> &lt; 0.05, RTX vs. iRTX Mann–Whitney test.</p>
Full article ">Figure 3
<p>Effects of 30 min prolonged TRPV1 modulation on Ca<sup>2+</sup> homeostasis. (<b>A</b>–<b>C</b>) Reticular Ca<sup>2+</sup> concentration ([Ca<sup>2+</sup>]<sub>r</sub>. (<b>A</b>) Time traces showing [Ca<sup>2+</sup>]<sub>r</sub> measured with erGAP1 probe during thapsigargin stimulation (2 µM) from control cells (Ctrl, black line) and 30 min-pretreated cells with RTX (10 µM; pink line) or iRTX (10 µM; blue line). (<b>B</b>) Scatter plots representing the steady-state [Ca<sup>2+</sup>]<sub>r</sub> concentration and (<b>C</b>) the total reticular Ca<sup>2+</sup> content assessed by thapsigargin (2 µM) from control cells (black; n = 45) and 30 min-pretreated cells with RTX (10 µM; pink; n = 16) or iRTX (10 µM; blue; n = 18). (<b>D</b>–<b>F</b>) Cytosolic Ca<sup>2+</sup> concentration ([Ca<sup>2+</sup>]<sub>c</sub>. (<b>D</b>) Time traces showing [Ca<sup>2+</sup>]<sub>c</sub> measured with Fura-2 AM probe during thapsigargin stimulation (2 µM) from control cells (black line) and 30 min-pretreated cells with RTX (10 µM; pink line) or iRTX (10 µM; blue line). (<b>E</b>) Scatter plots representing the steady-state [Ca<sup>2+</sup>]<sub>c</sub> and (<b>F</b>) the total cytosolic Ca<sup>2+</sup> content assessed by thapsigargin (2 µM) from control cells (black; n = 150) and 30 min-pretreated cells with RTX (10 µM; pink; n = 92) or iRTX (10 µM; blue; n = 103). (<b>G</b>–<b>I</b>) Mitochondrial Ca<sup>2+</sup> concentration ([Ca<sup>2+</sup>]<sub>m</sub>. (<b>G</b>) Time traces showing [Ca<sup>2+</sup>]<sub>m</sub> measured with 4mtD3cvp probe during NaATP stimulation (100 µM) from control cells (black line) and 30 min-pretreated cells with RTX (10 µM; pink line) or iRTX (10 µM; blue line). (<b>H</b>) Scatter plots representing the steady-state [Ca<sup>2+</sup>]<sub>m</sub> and (<b>I</b>) the total mitochondrial Ca<sup>2+</sup> content assessed by NaATP (100 µM) from control cells (black; n = 151) and 30 min-pretreated cells with RTX (10 µM; pink; n = 32) or iRTX (10 µM; blue; n = 53). (<b>J</b>–<b>L</b>) Ca<sup>2+</sup> concentration in mitochondrial hot spots ([Ca<sup>2+</sup>]<sub>hot spots</sub>). (<b>J</b>) Time traces showing [Ca<sup>2+</sup>]<sub>hot spots</sub> measured with N33D3cpv probe during NaATP stimulation (100 µM) from control cells (black line) and 30 min-pretreated cells with RTX (10 µM; pink line) or iRTX (10 µM; blue line). (<b>K</b>) Scatter plots representing the steady-state [Ca<sup>2+</sup>]<sub>hot spots</sub> and (<b>L</b>) total Ca<sup>2+</sup> content in mitochondrial hot spots by NaATP (100 µM) from control cells (black; n = 42) and 30 min-pretreated cells with RTX (10 µM; pink; n = 28) or iRTX (10 µM; blue; n = 44). Data are from at least three independent experiments. Statistics: * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 4
<p>Effects of 30 min prolonged TRPV1 modulation on interactions between endoplasmic reticulum and mitochondria. (<b>A</b>–<b>C</b>) Representative confocal microscopy images of in situ IP3R–VDAC interactions depicted as red dots (<b>A</b>) in control cells and 30 min-pretreated cells with (<b>B</b>) RTX (10 µM) or (<b>C</b>) iRTX (10 µM). Nuclei appear in blue. Scale bar: 50 µm. (<b>D</b>) Quantification of the interactions per cell presented as a fold of control; n = 30–31 cells. (<b>E</b>–<b>J</b>) Ultrastructural analysis by electron microscopy of reticulum–mitochondria interactions in control cells (black; n = 83) and in 30 min-pretreated cells with RTX (10 µM; pink; n = 73) or iRTX (10 µM; blue; n = 79). Schematics of the different parameters measured are explained in <a href="#app1-cells-12-02322" class="html-app">Figure S5</a>. (<b>E</b>) Quantification of reticulum–mitochondria interface expressed as a percentage of the mitochondrial circumference. (<b>F</b>) Frequency distribution of reticulum–mitochondria interactions. (<b>G</b>) Mean of the reticulum–mitochondria interaction width. (<b>H</b>–<b>J</b>) Representative images of electron microscopy in control cells (<b>H</b>) and 30 min-pretreated cells with RTX (<b>I</b>) or iRTX (<b>J</b>). M, mitochondria; ER, endoplasmic reticulum. Data are from at least three independent experiments. Statistics: * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 5
<p>Effects of TRPV1 conditioning on in vitro hypoxia/reoxygenation-induced cell death. (<b>A</b>) The experimental design representing hypoxia/reoxygenation (H/R) protocols achieved in control cells and cells treated with RTX (10 µM) or iRTX (10 µM): preconditioning (pre-C), per-conditioning (per-C), or postconditioning (post-C). (<b>B</b>) Dot plot showing mortality of SV40-transformed H9c2 cells to H/R (Ctrl) or concomitantly subjected to H/R and RTX or iRTX treatment. Evaluation of SV40-transformed H9c2 cell mortality was assessed via propidium iodide (PI) staining by flow cytometry. Sample size appears as follows: N = number of independent experiments; each symbol represents the mean of a triplicate, where each triplicate value corresponds to 10,000 events. Statistics: **** <span class="html-italic">p</span> &lt; 0.0001,* <span class="html-italic">p</span> &lt; 0.05 vs. Ctrl.</p>
Full article ">Figure 6
<p>Schematic summary of the main results. (<b>a</b>) Acute TRPV1 activation induced a slow decrease in the reticular Ca<sup>2+</sup> content. It led to a Ca<sup>2+</sup> increase at the mitochondrial surface and within the mitochondrial matrix, with an almost imperceptible cytosolic Ca<sup>2+</sup> change at the whole-cell level. (<b>b</b>) When TRPV1 activation was prolonged for 30 min, [Ca<sup>2+</sup>]<sub>r</sub> dropped, reducing the MAM interface and decreasing mitochondrial Ca<sup>2+</sup> content (matrix and surface) in favor of an increase in the cytosol. (<b>c</b>) Acute TRPV1 inhibition had no apparent effect on reticulum, cytosol, and mitochondrial hot spot contents, probably due to a SERCA pumping adaptation, but slowly depleted the mitochondrial compartment of Ca<sup>2+</sup>. (<b>d</b>) Over 30 min, TRPV1 inhibition increased and brought the ER–mitochondria interactions closer. We could expect an increase in [Ca<sup>2+</sup>]<sub>m</sub> in this condition. It was not the case, as the amount of Ca<sup>2+</sup> released from the reticulum was slightly lowered. The figure was created with BioRender.com (agreement number: JH25TUV6AR).</p>
Full article ">
14 pages, 643 KiB  
Review
What Does the Future Hold for Biomarkers of Response to Extracorporeal Photopheresis for Mycosis Fungoides and Sézary Syndrome?
by Oleg E. Akilov
Cells 2023, 12(18), 2321; https://doi.org/10.3390/cells12182321 - 20 Sep 2023
Cited by 1 | Viewed by 1079
Abstract
Extracorporeal photopheresis (ECP) is an FDA-approved immunotherapy for cutaneous T-cell lymphoma, which can provide a complete response in some patients. However, it is still being determined who will respond well, and predictive biomarkers are urgently needed to target patients for timely treatment and [...] Read more.
Extracorporeal photopheresis (ECP) is an FDA-approved immunotherapy for cutaneous T-cell lymphoma, which can provide a complete response in some patients. However, it is still being determined who will respond well, and predictive biomarkers are urgently needed to target patients for timely treatment and to monitor their response over time. The aim of this review is to analyze the current state of the diagnostic, prognostic, and disease state-monitoring biomarkers of ECP, and outline the future direction of the ECP biomarker discovery. Specifically, we focus on biomarkers of response to ECP in mycosis fungoides and Sézary syndrome. The review summarizes the current knowledge of ECP biomarkers, including their limitations and potential applications, and identifies key challenges in ECP biomarker discovery. In addition, we discuss emerging technologies that could revolutionize ECP biomarker discovery and accelerate the translation of biomarker research into clinical practice. This review will interest researchers and clinicians seeking to optimize ECP therapy for cutaneous T-cell lymphoma. Full article
Show Figures

Figure 1

Figure 1
<p>ECP Molecular Mechanisms of Action and Biomarkers of Response Overview. Summary of mechanisms of action from current knowledge of ECP, based on published studies. Abbreviations: MOP: 8-methoxypsoralen; APC: antigen-presenting cells; ECP: extracorporeal photopheresis; IL: interleukin; IL-1Ra: IL-1 receptor antagonist; INF-γ: interferon gamma; moDC: monocyte-Derived Cells; TNF-α: tumor necrosis factor alpha; Tregs: regulatory T cells; SzS: Sézary syndrome; UVA: ultraviolet A.</p>
Full article ">
21 pages, 3403 KiB  
Article
Effect of Acute Enriched Environment Exposure on Brain Oscillations and Activation of the Translation Initiation Factor 4E-BPs at Synapses across Wakefulness and Sleep in Rats
by José Lucas Santos, Evlalia Petsidou, Pallavi Saraogi, Ullrich Bartsch, André P. Gerber and Julie Seibt
Cells 2023, 12(18), 2320; https://doi.org/10.3390/cells12182320 - 20 Sep 2023
Viewed by 1502
Abstract
Brain plasticity is induced by learning during wakefulness and is consolidated during sleep. But the molecular mechanisms involved are poorly understood and their relation to experience-dependent changes in brain activity remains to be clarified. Localised mRNA translation is important for the structural changes [...] Read more.
Brain plasticity is induced by learning during wakefulness and is consolidated during sleep. But the molecular mechanisms involved are poorly understood and their relation to experience-dependent changes in brain activity remains to be clarified. Localised mRNA translation is important for the structural changes at synapses supporting brain plasticity consolidation. The translation mTOR pathway, via phosphorylation of 4E-BPs, is known to be activate during sleep and contributes to brain plasticity, but whether this activation is specific to synapses is not known. We investigated this question using acute exposure of rats to an enriched environment (EE). We measured brain activity with EEGs and 4E-BP phosphorylation at cortical and cerebellar synapses with Western blot analyses. Sleep significantly increased the conversion of 4E-BPs to their hyperphosphorylated forms at synapses, especially after EE exposure. EE exposure increased oscillations in the alpha band during active exploration and in the theta-to-beta (4–30 Hz) range, as well as spindle density, during NREM sleep. Theta activity during exploration and NREM spindle frequency predicted changes in 4E-BP hyperphosphorylation at synapses. Hence, our results suggest a functional link between EEG and molecular markers of plasticity across wakefulness and sleep. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Experimental design. (<b>A</b>) Experimental groups. After a 24 h habituation period (with or without EEG baseline recording), rats were kept awake (yellow squares) for 3 h at the end of the next dark period either in their home cage (HC) or in the enriched-environment (EE) cage. Some animals were sacrificed for tissue harvest (black arrowhead) immediately after this awake period (EE and SD groups) or left undisturbed to sleep for 3 h in their HC (SDS and EES groups). In a final group, rats were maintained awake for 6 h (SDSD) in their HC and sacrificed at the same circadian time as that of the sleeping groups. EEGs were recorded only in the sleeping groups. (<b>B</b>) Representation of the areas of the cortex and cerebellum harvested. (<b>C</b>) Whole-cell (TOT) and synapse-specific (SN) protein extracts were obtained from each tissue sample and processed for Western blot analyses. (<b>D</b>) Representative fronto-parietal (FP) and fronto-cerebellar (FC) EEsG and activity traces for the 5 main brains states. Created with BioRender.com.</p>
Full article ">Figure 2
<p>Effect of EE on wakefulness architecture and EEG. (<b>A</b>) Mean (±SEM) EEG power spectra (normalised to the mean across all frequencies (see <a href="#sec2-cells-12-02320" class="html-sec">Section 2</a>)) for AW and QW and both EEGs (FP and FC) separately for all groups (bsl and SD/EE periods). (<b>B</b>) Change in mean (±SEM) power density for AW (upper graph) and QW (lower graph) during the 3 h awake period in EE or HC expressed as % of corresponding baseline values (see <a href="#sec2-cells-12-02320" class="html-sec">Section 2</a>). EEG from both parietal (FP) and frontal (FC) EEGs for each group are represented for AW and QW. Time course (30 min bins) of sigma band in the parietal (FP) EEG for the EE and SD groups are represented as insets. For both (<b>A</b>,<b>B</b>), only the presence of significant differences with <span class="html-italic">p</span> &gt; 0.001 and <span class="html-italic">p</span> &gt; 0.0001 are reported underneath the traces for each group and EEG. Detailed statistics (Two-Way RM ANOVAs and level of significance for each comparison) can be found in the <a href="#app1-cells-12-02320" class="html-app">Data S1</a>.</p>
Full article ">Figure 3
<p>Effect of EE on sleep architecture and EEG. (<b>A</b>) Vigilance states (5–95% confidence interval, CI) and bout duration (5–95% CI, N = 8/group) during the 3 h rest following the awake period in either the HC (SD) or EE cage (EE). Two-way RM ANOVA did not reveal any effect of group (SD vs. EE) or condition (bsl vs. post-SD/EE) on vigilance states. (<b>B</b>) Change in mean (±SEM) power density for NREM and REM over the 3 h period after SD in EE or HC expressed as % of corresponding baseline values (see <a href="#sec2-cells-12-02320" class="html-sec">Section 2</a>). The parietal and frontal EEGs are shown separately. For clarity, only the presence of significant differences from baseline values are shown underneath the traces. Detailed statistics (Two-Way RM ANOVAs and level of significance for each comparison) can be found in <a href="#app1-cells-12-02320" class="html-app">Data S1</a>. Post-EE vs. post-SD, *: <span class="html-italic">p</span> &lt; 0.05, Sidak test. (<b>C</b>) Comparison of the time course of changes in NREM Delta, theta, sigma and beta frequency bands (mean ± SEM) in 30 min bins during the 3-h rest period after the awake period in an EE or HC (Two-way ANOVA Group effect, *: <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, Sidak test). Comparison between frontal and parietal EEGs are shown in insets for each frequency band. There were also no differences between EEGs for the SD group (Two-way ANOVA, <span class="html-italic">p</span> &gt; 0.05 for factors EEG and bins for each frequency band). SO and slow γ frequency bands (see <a href="#app1-cells-12-02320" class="html-app">Figure S3A</a>) did not show any significant differences between groups. (<b>D</b>) Changes in spindle characteristics (density, duration, frequency, amplitude) in rats from the EE group recorded from the parietal EEG (Paired <span class="html-italic">t</span>-test, **: <span class="html-italic">p</span> &lt; 0.01). (<b>E</b>) Correlations between spindle density and changes (as % of baseline) in sigma and beta frequency bands during NREM post-SD/EE in the parietal EEG. (<b>F</b>) Correlations between changes in frequency band power during the 3 h of wakefulness in the HC and EE and changes in parietal EEG power during NREM sleep during the following 3 h rest period. Data points for the EE and SD groups were grouped but are shown separately in the scatter plot. (Right graphs) Scatter plots showing the correlation between power in the sigma band in AW and NREM beta power and spindle duration. Results for the SD (open circles) and EE (filled circles) groups are shown. *: <span class="html-italic">p</span> &lt; 0.05, ***: <span class="html-italic">p</span> &lt; 0.001, Pearson’s.</p>
Full article ">Figure 4
<p>Sleep and experience-dependent changes in 4E-BP phosphorylation. (<b>A</b>) Representative Western Blot analysis of SN extracts obtained from the cerebellum of four different animals across the five groups indicated at the top. Phospo-4E-BPs (Thr37/46) are shown in green, 4E-BP2 protein in red. The individual channels are shown separately below. (<b>B</b>) Normalised mean (±SEM) phospo-4E-BPs (Thr37/46)/4E-BP2 (see <a href="#sec2-cells-12-02320" class="html-sec">Section 2</a>) for all brain regions pooled (N = 32) in SN (left graph) and TOT (right graph) extracts. Values are represented for all 4E-BP forms (α, β, γ) separately and grouped (α + β + γ; “ALL”). Awake groups are represented in grey and black, sleeping groups in light and dark blue and the circadian control group in orange. **** <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, Sidak test. (<b>C</b>) Normalised mean (±SEM) signals from antibodies detecting P-4E-BPs/4E-BP2 across groups. Note the specific decrease in the hypophosphorylated 4E-BP (α) form in the EES group. (<b>D</b>) Average (±SEM) 4E-BP conversion index (i.e., γ/α ratio) across groups. (<b>E</b>) Distribution of changes of the γ 4E-BP form in SN in all 4 brain regions (N = 8/region). A two-way RM ANOVA revealed an interaction between Group × brain region (<span class="html-italic">p</span> = 0.017) with significantly lower and higher levels of γ 4E-BP form in the cerebellum (Cb) in the control sleeping group (SDS) and circadian group (SDSD), respectively, compared to more frontal areas (SDS and SDSD: Cb vs. SM: * <span class="html-italic">p</span> = 0.048; SDSD: Cb vs. FR: ** <span class="html-italic">p</span> = 0.008).</p>
Full article ">Figure 5
<p>Relation between wakefulness and sleep EEG changes and synaptic 4E-BP measures. (<b>A</b>) Correlation matrix between changes in frequency band power in AW during the 3 h of wakefulness (in HC and EE) or sleep (NREM, IS and REM) during the rest period and changes in 4E-BP forms (α, β, γ) and conversion index (γ/α ratio). Results are shown for the SN fraction and separately for each EEG. Correlation coefficients were computed with datapoints from EE and SD groups combined. *: <span class="html-italic">p</span> &lt; 0.05, **: <span class="html-italic">p</span> &lt; 0.01, Pearson’s. (<b>B</b>) Scatter plots showing the correlation between theta power in AW and 4E-BP conversion index (γ/α ratio). Results for each EEG (FP = filled circles, FC = empty circle) and each group (EE = red, SD = black) are represented. Significance of correlations are shown for each EEG in the legend. (<b>C</b>) Scatter plots showing the correlation between Beta power in frontal EEG in REM sleep and 4E-BP conversion index (γ/α ratio) levels in SN. Results for each group (EE = red, SD = black) are represented. Significance of the correlation is shown next to the trendline. (<b>D</b>) Correlation matrix between changes in spindle characteristics (density, duration, frequency and amplitude) obtained from frontal (FC) and parietal (FP) EEGs and changes in 4E-BP forms (α, β, γ) and conversion index (γ/α ratio) at synapses. *: <span class="html-italic">p</span> &lt; 0.05, Pearson’s. The scatter plot showing the correlation between spindle frequency obtained from the parietal EEG and 4E-BP conversion index (γ/α ratio) is shown on the right. Results for each group (EE = red, SD = black) and significance of correlations are shown.</p>
Full article ">
14 pages, 17035 KiB  
Article
Modeling of FAN1-Deficient Kidney Disease Using a Human Induced Pluripotent Stem Cell-Derived Kidney Organoid System
by Sun Woo Lim, Dohyun Na, Hanbi Lee, Xianying Fang, Sheng Cui, Yoo Jin Shin, Kang In Lee, Jae Young Lee, Chul Woo Yang and Byung Ha Chung
Cells 2023, 12(18), 2319; https://doi.org/10.3390/cells12182319 - 20 Sep 2023
Cited by 2 | Viewed by 1543
Abstract
Karyomegalic interstitial nephritis (KIN) is a genetic kidney disease caused by mutations in the FANCD2/FANCI-Associated Nuclease 1 (FAN1) gene on 15q13.3, which results in karyomegaly and fibrosis of kidney cells through the incomplete repair of DNA damage. The aim of this [...] Read more.
Karyomegalic interstitial nephritis (KIN) is a genetic kidney disease caused by mutations in the FANCD2/FANCI-Associated Nuclease 1 (FAN1) gene on 15q13.3, which results in karyomegaly and fibrosis of kidney cells through the incomplete repair of DNA damage. The aim of this study was to explore the possibility of using a human induced pluripotent stem cell (hiPSC)-derived kidney organoid system for modeling FAN1-deficient kidney disease, also known as KIN. We generated kidney organoids using WTC-11 (wild-type) hiPSCs and FAN1-mutant hiPSCs which include KIN patient-derived hiPSCs and FAN1-edited hiPSCs (WTC-11 FAN1+/−), created using the CRISPR/Cas9 system in WTC-11-hiPSCs. Kidney organoids from each group were treated with 20 nM of mitomycin C (MMC) for 24 or 48 h, and the expression levels of Ki67 and H2A histone family member X (H2A.X) were analyzed to detect DNA damage and assess the viability of cells within the kidney organoids. Both WTC-11-hiPSCs and FAN1-mutant hiPSCs were successfully differentiated into kidney organoids without structural deformities. MMC treatment for 48 h significantly increased the expression of DNA damage markers, while cell viability in both FAN1-mutant kidney organoids was decreased. However, these findings were observed in WTC-11-kidney organoids. These results suggest that FAN1-mutant kidney organoids can recapitulate the phenotype of FAN1-deficient kidney disease. Full article
Show Figures

Figure 1

Figure 1
<p><b><span class="html-italic">FAN1</span> gene mutation in a patient with karyomegalic interstitial nephritis (KIN).</b> (<b>A</b>) Hematoxylin and eosin (H&amp;E) staining of kidney tissues of a patient with karyomegalic interstitial nephritis. Scale bar = 100 μm. White arrows in A point to karyomegaly. (<b>B</b>) Patient PBMC with KIN showing <span class="html-italic">FAN1</span> gene mutation on deletion c.1985-1944delTTGGGTGGAT on 15q13.3. (<b>C</b>) Pedigree of a family showing individuals affected by karyomegalic interstitial nephritis resulting in the appearance of p.Gly663llefs*54.</p>
Full article ">Figure 2
<p><b>Establishment of CRISPR/Cas9 ribonucleoproteins (RNP)-mediated <span class="html-italic">FAN1</span> gene editing in WTC-11 hiPSCs.</b> (<b>A</b>) Target site for guide RNA (gRNA) targeting <span class="html-italic">FAN1</span>. Target sites are indicated by a green color in exon 2 of the <span class="html-italic">FAN1</span> gene. The downward-pointing arrowhead indicates the position of the canonical cut site and predicted specificity based on the number and distribution of homoeologous SNPs at the corresponding target site/PAM. (<b>B</b>) PAM sequences (5′-NGG-3′) in target site are indicated by red letters. The table shows no mismatched number with gRNA. The asterisk indicates on-target gRNA. (<b>C</b>) Indel sequences after transfecting WTC-11 hiPSCs with gRNA using the CRISPR/Cas9 RNP method. Indel sequences are indicated by red letters (a green color indicates a target site). (<b>D</b>) Read number of In-del frequency. Indel of 1bp (A ins) in the target sequence read of about 50%. (<b>E</b>) Morphology of <span class="html-italic">FAN1</span>-gene-edited WTC-11 iPSC (WTC-11<span class="html-italic"><sup>FAN1+/−</sup></span>). (<b>F</b>,<b>G</b>) Flow cytometry analysis and immunofluorescence image of cells expressing NANOG, SSEA-4, and TRA-1-81 in WTC-11<span class="html-italic"><sup>FAN1+/−</sup></span> hiPSCs. Scale bar = 50 μm. (<b>H</b>) Immunofluorescence staining of three germ layer markers. Ectoderm, mesoderm, and entoderm differentiation were detected using PAX4, SM22a, and FOX2A, respectively. Scale bar = 50 μm. (<b>I</b>) Expression analysis via RT-PCR of <span class="html-italic">FAN1</span> and <span class="html-italic">GAPDH</span> in WTC-11 and WTC<span class="html-italic"><sup>FAN1+/−</sup></span> hiPSCs. All expression levels were normalized against <span class="html-italic">GAPDH</span> expression level. (<b>J</b>) Expression analysis via immunoblot of FAN1 and <span class="html-italic">β-actin</span> in WTC-11 and WTC<span class="html-italic"><sup>FAN1+/−</sup></span> hiPSCs. All expression levels were normalized against the <span class="html-italic">β-actin</span> expression level. Data are presented as mean ± standard error. *, <span class="html-italic">p</span> &lt; 0.05 vs. WTC-11 iPSC.</p>
Full article ">Figure 3
<p><b>Differentiation of kidney organoids from WTC-11, WTC-11<span class="html-italic"><sup>FAN1+/−</sup></span>, and KIN patient hiPSCs.</b> (<b>A</b>) Schematic timeline of the hiPSC differentiation protocol. (<b>B</b>) Representative immunofluorescence images of podocalyxin (PODXL), lotus tetragonolobus lectin (LTL), and e-cadherin (ECAD) kidney organoids from WTC-11, WTC-11<span class="html-italic"><sup>FAN1+/−</sup></span>, and KIN patient, respectively. (<b>C</b>) Quantification of PODXL, LTL or ECAD-positive cells per organoid in each group. Scale bar = 50 or 100 μm.</p>
Full article ">Figure 4
<p><b>Effect of mitomycin C (MMC) treatment in kidney organoids from WTC-11, WTC-11<span class="html-italic"><sup>FAN1+/−</sup></span>, and KIN patient hiPSCs.</b> (<b>A</b>) Each kidney organoid was treated with 20 nM MMC. After 24 h or 48 h of incubation, each kidney organoid was stained with Ki67 antibody. Images with DAPI show nuclei positive for the Ki67 antibody. Scale bar = 50 or 100 μm. (<b>B</b>) Quantification of Ki67-positive cells in kidney organoids from WTC-11, WTC-11<span class="html-italic"><sup>FAN1+/−</sup></span>, and KIN patient hiPSCs. (<b>C</b>) Flow cytometry gating strategy illustrating a viable cell population being subgated to the level of podocalyxin+ (PODXL), lotus tetragonolobus lectin+ (LTL), or e-cadherin+ (ECAD) in kidney organoid cells from WTC-11, WTC-11<span class="html-italic"><sup>FAN1+/−</sup></span>, and the KIN patient hiPSCs, respectively, after treatment with MMC for 24 h or 48 h. (<b>D</b>–<b>F</b>) Quantification of percentage of viable cells in PODXL+, LTL+, or ECAD+ in cells from each kidney organoid. (<b>G</b>,<b>H</b>) Immunoblot analysis and its quantification of H2A.X in kidney organoids from WTC-11, WTC-11<span class="html-italic"><sup>FAN1+/−</sup></span>, and KIN patient hiPSCs after treatment with MMC for 24 h or 48 h. Data were normalized against the β-actin expression level. All data are presented as mean ± standard error. *, <span class="html-italic">p</span> &lt; 0.05 vs. Nil group or 24 h MMC group.</p>
Full article ">
20 pages, 3016 KiB  
Review
Biochemical and Molecular Pathways in Neurodegenerative Diseases: An Integrated View
by Nitesh Sanghai and Geoffrey K. Tranmer
Cells 2023, 12(18), 2318; https://doi.org/10.3390/cells12182318 - 20 Sep 2023
Cited by 6 | Viewed by 5445
Abstract
Neurodegenerative diseases (NDDs) like Alzheimer’s disease (AD), Parkinson’s disease (PD), and amyotrophic lateral sclerosis (ALS) are defined by a myriad of complex aetiologies. Understanding the common biochemical molecular pathologies among NDDs gives an opportunity to decipher the overlapping and numerous cross-talk mechanisms of [...] Read more.
Neurodegenerative diseases (NDDs) like Alzheimer’s disease (AD), Parkinson’s disease (PD), and amyotrophic lateral sclerosis (ALS) are defined by a myriad of complex aetiologies. Understanding the common biochemical molecular pathologies among NDDs gives an opportunity to decipher the overlapping and numerous cross-talk mechanisms of neurodegeneration. Numerous interrelated pathways lead to the progression of neurodegeneration. We present evidence from the past pieces of literature for the most usual global convergent hallmarks like ageing, oxidative stress, excitotoxicity-induced calcium butterfly effect, defective proteostasis including chaperones, autophagy, mitophagy, and proteosome networks, and neuroinflammation. Herein, we applied a holistic approach to identify and represent the shared mechanism across NDDs. Further, we believe that this approach could be helpful in identifying key modulators across NDDs, with a particular focus on AD, PD, and ALS. Moreover, these concepts could be applied to the development and diagnosis of novel strategies for diverse NDDs. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic presentation of various biochemical cross-talks and their detrimental manifestations (<b>A</b>–<b>I</b>) in the brain provoked by oxidative stress and their implications in the progress of neurodegenerative diseases like Alzheimer’s disease (AD), Parkinson’s disease (PD), and amyotrophic lateral sclerosis (ALS). Brain is highly vulnerable to oxidative stress due to low regenerative capacity, enrichment of polyunsaturated fatty acids, high dependency on mitochondria for adenosine triphosphate (ATP) generation, elevated glucose demand, high concentration of metals like ferrous ion (Fe<sup>+2</sup>), cuprous ion (Cu<sup>+</sup>), zinc ion (Zn<sup>+2</sup>) and calcium ion (Ca<sup>+2</sup>), glutamate-induced excitotoxicity, high oxygen (O<sub>2</sub>) consumption, and relatively low antioxidant system. These multiple factors initiate various reaction pathways to create redox disbalance called oxidative and nitrosative stress in the brain, implicated in various NDDs. (<b>A</b>)<b>.</b> The triplet unstable O<sub>2</sub> undergoes reduction to produce the precursor of all radicals called superoxide anion radical (O<sub>2</sub><sup>•−</sup>) via NAD(P)H oxidases (NOXs) pathway, i.e., one-electron trans-membrane transfer to (O<sub>2</sub>) [<a href="#B95-cells-12-02318" class="html-bibr">95</a>]. (<b>B</b>)<b>.</b> Antioxidant superoxide dismutase (SOD1) undergoes dismutation to scavenge (O<sub>2</sub><sup>•−</sup>) to produce hydrogen peroxide (H<sub>2</sub>O<sub>2</sub>). (<b>C</b>)<b>.</b> The weakly liganded (Fe<sup>+2</sup>) and (Cu<sup>+</sup>) undergo reduction to produce nature’s most vulnerable oxidant hydroxyl radical (HO<sup>•</sup>) through Fenton’s reaction and Haber-Weiss reaction. (<b>D</b>)<b>.</b> The final 4th electron reduction of H<sub>2</sub>O<sub>2</sub> in the presence of antioxidants, like glutathione peroxidase (Gpx), catalase (cat), and peroxiredoxin system (Prx), forms water (H<sub>2</sub>O). (<b>E</b>)<b>.</b> Overactivation of neuronal nitric oxide synthase (nNOS) produces nitric oxide (NO<sup>•</sup>) radicals from L-arginine, which create nitrosative stress by modification of thiol group (SH) containing proteins. (<b>F</b>)<b>.</b> Excessive superoxide anion radicals lead to inactivation of nitric oxide production and switch the biology to production of highly potent oxidant peroxynitrite anion (ONOO<sup>−</sup>), which leads to the nitrosative stress by (SH) modification of free tyrosine (Tyr) residues to form 3-nitrotyrosine (3-NO2Tyr) (<b>G</b>), which act as a versatile biomarker of nitrosative stress and NDDs. (<b>H</b>)<b>.</b> Highly reactive and mutagenic oxidant (HO<sup>•</sup>) damages the nucleic acid deoxyribonucleic acid/ribonucleic acid (DNA/RNA) to form oxidative products 8-hydroxy-2′-deoxyguanosine(8-OHdG) and (8-OxoG), and acts as a universal biomarker for oxidative stress and NDDs (important to note that guanine is the most oxidation prone nucleobase because of low reduction potential [<a href="#B96-cells-12-02318" class="html-bibr">96</a>]). Further, HO<sup>•</sup> radical causes lipid peroxidation of lipid-rich neuronal membranes, resulting in the death of neurons. Lipid peroxides (ROO<b><sup>.</sup></b>) act as a biomarker of oxidative stress and NDDs. Created with BioRender.com (accessed on 19 September 2023).</p>
Full article ">Figure 2
<p>Schematic presentation of various biochemical cross-talks, involving calcium ion (Ca<sup>+2</sup>), ferrous ion (Fe<sup>+2</sup>), and Zinc ion (Zn<sup>+2</sup>) implicated in the progress of neurodegenerative diseases like Alzheimer’s disease (AD), Parkinson’s disease (PD), and amyotrophic lateral sclerosis (ALS). (<b>A</b>). Excitotoxicity (neuronal death) is triggered by the excessive release of excitatory neurotransmitter glutamate (neurotoxic) from the presynaptic neuron and leads to activation of various biochemical cascades leading to neurotoxicity and hence, neuronal death. This process is initiated by the activation of the <span class="html-italic">N</span>-methyl-D-aspartic acid receptors (NMDAR) by excessive glutamate at postsynaptic neurons and thereby the release and accumulation of toxic intraneuronal Ca<sup>2+</sup>. (<b>B</b>). Glutamate-mediated excitotoxicity is increased because of the astrocyte-mediated downregulation of excitatory amino acid transporters 2 (EAAT2), which slows down the uptake of glutamate from the synaptic cleft and incites the excitotoxicity cascade. (<b>C</b>). Ca<sup>2+</sup> overload initiates most of the deleterious downstream mechanisms of the cascade, through increasing Ca<sup>2+</sup> overload in mitochondria, induction of proteases (calpains and caspases), decreasing the proton gradient (ΔpH), mitochondrial membrane potential (ΔΨm) and adenosine triphosphate (ATP), activation of phospholipase A2 (PLA2) pathway initiating downstream activation of arachidonic acid and prostaglandin E2 (PGE2), aggravation of mitochondrial and endoplasmic reticulum stress leading to superoxide dismutase (SOD1) and TAR DNA-binding protein (TDP-43) aggregation. (<b>D</b>). Surge of reactive oxygen species (ROS) like hydrogen peroxide (H<sub>2</sub>O<sub>2</sub>) and hydroxyl radical (HO<sup>•</sup>) and reactive nitrogen species (RNS) like nitric oxide (NO<sup>•</sup>) radical, formation of peroxynitrite anion (ONOO<sup>−</sup>) increases the intraneuronal Zn<sup>2+</sup> mobilization, which targets mitochondria and further exacerbates Ca<sup>2+</sup> dysregulation and ROS production. (<b>E</b>). Ca<sup>+2</sup> and Fe<sup>+2</sup> dysregulation participates in the ferroptosis death of neurons. Iron dysregulation leads to Ca<sup>2+</sup> dysregulation and vice versa. Excessive glutamate increases the Fe<sup>+2</sup> intake inside the neurons, thereby leading to excitotoxicity and lipid peroxidation via Fenton’s reaction called Ferroptosis. Created with BioRender.com.</p>
Full article ">Figure 3
<p>Schematic presentation of various common neuronal biochemical pathways perturbed or compromised in multiple neurodegenerative diseases, such as AD, PD, and ALS. The key points in the pathway and the selected disease-associated proteins are demonstrated in this picture. 1. Protein Quality Control (PQC) proteostasis network: molecular chaperones, including heat shock proteins (Hsp90, Hsp70, and Hsp40), regulate protein folding and maturation. Ubiquitin-proteosome system (UPS) is a crucial protein degradation pathway and is important for PQC and homeostasis. Any defect in the PQC leads to neurodegeneration (AD, PD, ALS). Decline of proteostasis is the hallmark of ageing and it decreases with age, leading to the accumulation of toxic and non-functional aggregates. 2. Autophagy-Lysosome Pathway (a,b,c,d,e.): Perturbations throughout the pathway, from initiation of autophagosome formation to degradation in the autolysosomes, have been suggested to be involved in neurodegenerative diseases like AD, PD, and ALS and further, could build an accumulation of pathogenic and toxic protein aggregates and defective mitochondria. a. Autophagy initiation defects due to decreased expression of protein Beclin1 in case of AD. b. Loss of sequestration into autophagosomes due to mutations in the gene-encoding p62/optineurin in the case of ALS, and mitophagy defects due to mutations in the gene-encoding protein PINK1/Parkin in the case of PD c. Defects in the maturation of autophagosome are due to decreasing expression of PICLAM protein in the case of AD, whereas mutation in SIGMAR1 gene in the case of ALS. c. Defects in vesicle trafficking (lysosome to membrane) are due to the mutations in the gene-encoding protein dynactin/profilin in case of ALS. 3. Dysregulation of mitochondrial quality control (MQC): including a (mitochondrial damage), b (mitochondrial fusion and fission dynamics), c (selective autophagy of mitochondria called mitophagy) results in decreased ATP production and dysfunctional proteostasis network. 4. Axonal transport defects in AD, PD, and ALS and underlying mechanisms: Defective axonal transport is due to perturbed anterograde and retrograde transport mechanisms involving mitochondrial kinesin and endosomal transport protein dynein. Further, disrupted neurofilament (NF) in forms of phosphorylated NF in the case of AD, PD, and ALS and microtubules (including α-Tubulin and β-Tubulin) are involved in the impairment of transport across neurons. 5. Protein Seeding and Propagation: Dysfunction of Intracellular propagation and seeding of toxic protein aggregates involved in the disease progression in case of AD, PD, and ALS. Created with BioRender.com.</p>
Full article ">
27 pages, 4896 KiB  
Article
Differential Effects of Regulatory T Cells in the Meninges and Spinal Cord of Male and Female Mice with Neuropathic Pain
by Nathan T. Fiore, Brooke A. Keating, Yuting Chen, Sarah I. Williams and Gila Moalem-Taylor
Cells 2023, 12(18), 2317; https://doi.org/10.3390/cells12182317 - 20 Sep 2023
Cited by 7 | Viewed by 1832
Abstract
Immune cells play a critical role in promoting neuroinflammation and the development of neuropathic pain. However, some subsets of immune cells are essential for pain resolution. Among them are regulatory T cells (Tregs), a specialised subpopulation of T cells that limit excessive immune [...] Read more.
Immune cells play a critical role in promoting neuroinflammation and the development of neuropathic pain. However, some subsets of immune cells are essential for pain resolution. Among them are regulatory T cells (Tregs), a specialised subpopulation of T cells that limit excessive immune responses and preserve immune homeostasis. In this study, we utilised intrathecal adoptive transfer of activated Tregs in male and female mice after peripheral nerve injury to investigate Treg migration and whether Treg-mediated suppression of pain behaviours is associated with changes in peripheral immune cell populations in lymphoid and meningeal tissues and spinal microglial and astrocyte reactivity and phenotypes. Treatment with Tregs suppressed mechanical pain hypersensitivity and improved changes in exploratory behaviours after chronic constriction injury (CCI) of the sciatic nerve in both male and female mice. The injected Treg cells were detected in the choroid plexus and the pia mater and in peripheral lymphoid organs in both male and female recipient mice. Nonetheless, Treg treatment resulted in differential changes in meningeal and lymph node immune cell profiles in male and female mice. Moreover, in male mice, adoptive transfer of Tregs ameliorated the CCI-induced increase in microglia reactivity and inflammatory phenotypic shift, increasing M2-like phenotypic markers and attenuating astrocyte reactivity and neurotoxic astrocytes. Contrastingly, in CCI female mice, Treg injection increased astrocyte reactivity and neuroprotective astrocytes. These findings show that the adoptive transfer of Tregs modulates meningeal and peripheral immunity, as well as spinal glial populations, and alleviates neuropathic pain, potentially through different mechanisms in males and females. Full article
(This article belongs to the Special Issue Role of Glial Cells in Neuropathic Pain)
Show Figures

Figure 1

Figure 1
<p>Adoptively transferred Tregs accumulate in the pia 48 h after intrathecal injection in naïve mice. (<b>A</b>) GFP+ Tregs, or (<b>B</b>) FoxP3+ Tregs in the meninges and choroid plexus and (<b>C</b>) GFP+ Tregs, or (<b>D</b>) FoxP3+ Tregs in the peripheral lymphatic system of naïve female mice 48 h post-intrathecal delivery of activated Tregs, or saline-based vehicle (control). Raw data was analysed using estimation statistics (* = large effect size, 0.8 &lt; g &lt; 1.2; ** = very large effect size, g &gt; 1.2 for Treg-recipient mice compared to saline group, <span class="html-italic">n</span> = 5–6 per group). D/A = dura/arachnoid, CP = choroid plexus. Data expressed as mean ± SEM.</p>
Full article ">Figure 2
<p>Adoptive transfer using intrathecal administration of Tregs reduces pain behaviours following CCI. (<b>A</b>) Schematic of experimental timeline for adoptive transfer of Tregs following CCI, behavioural assays, and tissue dissection. Nerve-injured and sham-operated mice were injected with activated Tregs or saline-based vehicle (control). Mechanical allodynia in CCI is reduced in males (<b>B</b>) and females (<b>C</b>) following intrathecal delivery of activated Tregs. Mice developed reduced paw withdrawal thresholds following CCI (* 0.8 &lt; g &lt; 1.2 and ** g &gt; 1.2 relative to pre-injury), which transiently recovered following spinal delivery of Tregs (## = relative to day 3, pre-Treg injection, <span class="html-italic">n</span> = 7–10 per group).</p>
Full article ">Figure 3
<p>Intrathecal administration of activated Tregs improves exploratory behaviours in male and female mice following CCI. Nerve-injured and sham-operated mice were injected with activated Tregs or saline-based vehicle (control). The open field hole-board test was utilised to measure the number of rears (<b>A</b>), nose pokes (<b>B</b>), distance covered (<b>C</b>), average speed (<b>D</b>), as well as time spent in the centre of the open field (<b>E</b>). Testing was conducted on the 9th day after CCI. Data are shown as mean ± SEM (* 0.8 &lt; g &lt; 1.2 and ** g &gt; 1.2 for CCI relative to sham saline and # 0.8 &lt; g &lt; 1.2 and ## g &gt; 1.2 for Tregs relative to CCI saline control, <span class="html-italic">n</span> = 7–10 per group).</p>
Full article ">Figure 4
<p>Tracking intrathecally injected GFP+ Tregs following CCI. GFP+ cells in the meninges and choroid plexus of male (<b>A</b>) and female (<b>B</b>) mice, as well as in the peripheral lymphatics of male (<b>C</b>) and female (<b>D</b>) mice following intrathecal delivery of activated Tregs or saline-based vehicle (control) in naïve and nerve-injured mice. (* 0.8 &lt; g &lt; 1.2 for Tregs relative to saline control, <span class="html-italic">n</span> = 6 per group). D/A = dura/arachnoid, CP = choroid plexus, ILN = inguinal lymph nodes, CLN = cervical lymph nodes, SPLN = sciatic/popliteal lymph nodes. (<b>E</b>) Representative images taken from the lumbar spinal cord stained for eGFP+ cells (green) counterstained with DAPI (blue/white specks) in nerve-injured male and female mice following vehicle and GFP+ Treg injection. No eGFP+ cells were identified in the spinal cord, with total eGFP+ cell numbers shown in male (<b>F</b>) and female (<b>G</b>) mice following intrathecal delivery of activated Tregs or vehicle in naïve and nerve-injured mice (<span class="html-italic">n</span> = 4 per group).</p>
Full article ">Figure 5
<p>Immunological changes in the meninges and choroid plexus following intrathecal Treg injection in nerve-injured male and female mice. Heatmaps summarising immunological changes in the (<b>A</b>) dura/arachnoid, (<b>B</b>) pia, and (<b>C</b>) choroid plexus of male mice and (<b>D</b>) dura/arachnoid, (<b>E</b>) pia, and (<b>F</b>) choroid plexus of female mice following peripheral nerve injury or sham surgery and intrathecal delivery of activated Tregs, or saline-based vehicle (control). Data normalised for the heatmaps. Raw data were analysed using estimation statistics. (* 0.8 &lt; g &lt; 1.2 and ** g &gt; 1.2 for differences relative to sham saline and # 0.8 &lt; g &lt; 1.2 and ## g &gt; 1.2 relative to CCI saline, <span class="html-italic">n</span> = 7–10 per group).</p>
Full article ">Figure 6
<p>Immunological changes in peripheral lymph nodes following intrathecal Treg injection in nerve-injured male and female mice. Heatmaps summarising immunological changes in the (<b>A</b>) cervical LNs and (<b>B</b>) sciatic/popliteal LNs of male mice, and cervical LNs (<b>C</b>) and sciatic/popliteal LNs (<b>D</b>) of female mice following peripheral nerve injury or sham surgery and intrathecal delivery of activated Tregs, or saline-based vehicle (control). Data normalised for the heatmaps. Raw data were analysed using estimation statistics. (* 0.8 &lt; g &lt; 1.2 and ** g &gt; 1.2 for differences relative to sham saline and # 0.8 &lt; g &lt; 1.2 and ## g &gt; 1.2 relative to CCI saline, <span class="html-italic">n</span> = 7–10 per group).</p>
Full article ">Figure 7
<p>Changes in microglial phenotype in the ipsilateral spinal cord following intrathecal administration of Tregs in male and female mice. Nerve-injured and sham-operated mice were injected with activated Tregs or saline-based vehicles (control). Opal multiplex immunohistochemistry and immunofluorescent staining were used to assess microglial markers (<b>A</b>) IBA-1, (<b>B</b>) P2RY12, (<b>C</b>) LYZ:IBA-1 ratio in males and (<b>D</b>) IBA-1, (<b>E</b>) P2RY12, (<b>F</b>) LYZ:IBA-1 ratio in females in the ipsilateral dorsal and ventral horns. Data are shown as mean ± SEM (* 0.8 &lt; g &lt; 1.2 and ** g &gt; 1.2 for changes relative to sham saline and # 0.8 &lt; g &lt; 1.2 and ## g &gt; 1.2 relative to CCI saline, <span class="html-italic">n</span> = 4–6 per group).</p>
Full article ">Figure 8
<p>Changes in macrophage/microglial phenotype in the ipsilateral spinal cord following intrathecal administration of Tregs in male and female mice. Nerve-injured and sham-operated mice were injected with activated Tregs or saline-based vehicles (control). Opal multiplex immunohistochemistry and immunofluorescent staining were used to assess macrophage/microglial markers (<b>A</b>) CD206: P2RY12 ratio, (<b>B</b>) CD11c+ IGF1+ cell density and (<b>C</b>) CD163+ cell density in males and (<b>D</b>) CD206: P2RY12 ratio, (<b>E</b>) CD11c+ IGF1+ cell density and (<b>F</b>) CD163+ cell density in females in the ipsilateral dorsal and ventral horns. Data are shown as mean ± SEM (* 0.8 &lt; g &lt; 1.2 and ** g &gt; 1.2 for changes relative to sham saline and # 0.8 &lt; g &lt; 1.2 and ## g &gt; 1.2 relative to CCI saline, <span class="html-italic">n</span> = 4–6 per group).</p>
Full article ">Figure 9
<p>Changes in astrocyte phenotype in the ipsilateral spinal cord following intrathecal administration of Tregs in male and female mice. Nerve-injured and sham-operated mice were injected with activated Tregs or saline-based vehicles (control). Opal multiplex immunohistochemistry and immunofluorescent staining were used to assess astrocyte markers (<b>A</b>) GFAP, (<b>B</b>) C3+ GFAP+ cell density and (<b>C</b>) S100A10/GFAP colocalisation in males, and (<b>D</b>) GFAP, (<b>E</b>) C3+ GFAP+ cell density and (<b>F</b>) S100A10/GFAP colocalisation in females in the ipsilateral dorsal and ventral horns. Data are shown as mean ± SEM (* 0.8 &lt; g &lt; 1.2 and ** g &gt; 1.2 for CCI relative to sham saline and # 0.8 &lt; g &lt; 1.2 and ## g &gt; 1.2 relative to CCI saline, <span class="html-italic">n</span> = 4–6 per group).</p>
Full article ">
19 pages, 2940 KiB  
Article
Mutant Cytochrome C as a Potential Detector of Superoxide Generation: Effect of Mutations on the Function and Properties
by Rita V. Chertkova, Ilya P. Oleynikov, Alexey A. Pakhomov, Roman V. Sudakov, Victor N. Orlov, Marina A. Semenova, Alexander M. Arutyunyan, Vasily V. Ptushenko, Mikhail P. Kirpichnikov, Dmitry A. Dolgikh and Tatiana V. Vygodina
Cells 2023, 12(18), 2316; https://doi.org/10.3390/cells12182316 - 19 Sep 2023
Cited by 2 | Viewed by 1400
Abstract
Cytochrome c (CytC) is a single-electron carrier between complex bc1 and cytochrome c-oxidase (CcO) in the electron transport chain (ETC). It is also known as a good radical scavenger but its participation in electron flow through the ETC makes it impossible to use [...] Read more.
Cytochrome c (CytC) is a single-electron carrier between complex bc1 and cytochrome c-oxidase (CcO) in the electron transport chain (ETC). It is also known as a good radical scavenger but its participation in electron flow through the ETC makes it impossible to use CytC as a radical sensor. To solve this problem, a series of mutants were constructed with substitutions of Lys residues in the universal binding site (UBS) which interact electrostatically with negatively charged Asp and Glu residues at the binding sites of CytC partners, bc1 complex and CcO. The aim of this study was to select a mutant that had lost its function as an electron carrier in the ETC, retaining the structure and ability to quench radicals. It was shown that a mutant CytC with substitutions of five (8Mut) and four (5Mut) Lys residues in the UBS was almost inactive toward CcO. However, all mutant proteins kept their antioxidant activity sufficiently with respect to the superoxide radical. Mutations shifted the dipole moment of the CytC molecule due to seriously changed electrostatics on the surface of the protein. In addition, a decrease in the redox potential of the protein as revealed by the redox titrations of 8Mut was detected. Nevertheless, the CD spectrum and dynamic light scattering suggested no significant changes in the secondary structure or aggregation of the molecules of CytC 8Mut. Thus, a variant 8Mut with multiple mutations in the UBS which lost its ability to electron transfer and saved most of its physico-chemical properties can be effectively used as a detector of superoxide generation both in mitochondria and in other systems. Full article
Show Figures

Figure 1

Figure 1
<p>Locations of mutated residues in the 3D structure of CytC. Qualitative electrostatic potential surface of horse CytC WT (<b>a</b>) and 8Mut (<b>b</b>) are shown from the side of the heme cavity and when rotated by 180° around the vertical axis. ε-amino groups of Lys residues and carboxyl groups of Glu residues subjected to mutagenesis are marked with blue and red balls/numbers, respectively. The positions of unaltered Lys13 and Lys79 residues near heme cavity are marked with grey numbers. The images are obtained using the PyMOL program (pdb code 1HRC).</p>
Full article ">Figure 2
<p>Isoelectric surfaces of CytC WT (<b>a</b>) and 8Mut (<b>b</b>). Blue and red surfaces correspond to charge +0.1 mV and −0.1 mV, respectively. Orientation of the molecules is the same as in <a href="#cells-12-02316-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure 3
<p>Typical oxygen consumption kinetics of CcO oxidizing CytC in the presence of TMPD and ascorbate: WT (black line) and CytC mutants: 2Mut (red line), 5Mut (green line), and 8Mut (blue line). The final concentrations in the measurement medium were CcO, 24 nM (black line) and 120 nM (red, green, and blue lines); ascorbate (5 mM); TMPD (0.1 mM). The addition of cytochromes (1 µM each) is marked with arrows with the letter c. The inset shows the dependences of the oxygen consumption rate on the concentration of CytC 2Mut, obtained in the presence of TMPD (1, red dots) and without it (2, black dots).</p>
Full article ">Figure 4
<p>Typical kinetics of superoxide anion radical reduction during hypoxanthine oxidation by xanthine oxidase, measured by the reduction of CytC WT (black line) and mutant forms of CytC 2Mut (red line), CytC 5Mut (green line), and CytC 8Mut (blue line). CytC reduction was recorded spectrally in the two-wavelength mode by the difference in absorption at 550 nm and 535 nm (comparison wave). The concentration of CytC in the measurement medium in all experiments was 4 µM, hypoxanthine, 50 µM, and xanthine oxidase, 0.015 U/mL. The measurements were carried out in 50 mM K-phosphate buffer, pH 7.5, 0.1 mM EDTA.</p>
Full article ">Figure 5
<p>The redox titration curves of CytC, wild-type (WT—transparent circles), and mutated (8Mut—blue circles) forms.</p>
Full article ">Figure 6
<p>The longwave region of the absorption spectrum of ferric CytC: CytC (S) (magenta line), horse heart CytC (Sigma, Burlington, MA, USA), recombinant CytC WT (black line), and CytC 8Mut (blue line). For comparison, the spectrum of ferrous CytC 8Mut is also shown (red line, dotted line). Absorbance is normalized to the concentration of cytochromes in measurement samples that have been pre-oxidized with substoichiometric amounts of ferricyanide. The measurements were carried out in basic buffer, pH 7.6.</p>
Full article ">Figure 7
<p>CD spectra of CytC WT (black line) and mutant variants 2Mut (red line), 5Mut (green line), and 8Mut (blue line). The spectra were normalized to the concentration of CytC in the studied samples, which was about 0.6 mM (measured from the band at 550 nm). The inset shows what the same CD spectra look like normalized to protein concentration based on far UV molar extinction at 205 nm, 210 nm, and 215 nm (<a href="#cells-12-02316-t004" class="html-table">Table 4</a> in [<a href="#B38-cells-12-02316" class="html-bibr">38</a>]). Three values are averaged.</p>
Full article ">Figure 8
<p>Particle size of CytC samples: CytC (S) (black squares), horse heart CytC from Sigma, and mutant forms 2Mut (red circles), 5Mut (green triangles), and 8Mut (blue triangles) measured by dynamic light scattering.</p>
Full article ">
21 pages, 4862 KiB  
Article
Characterisation of Lipoma-Preferred Partner as a Novel Mechanotransducer in Vascular Smooth Muscle Cells
by Alexandra Sporkova, Taslima Nahar, Mingsi Cao, Subhajit Ghosh, Carla Sens-Albert, Prisca Amayi Patricia Friede, Anika Nagel, Jaafar Al-Hasani and Markus Hecker
Cells 2023, 12(18), 2315; https://doi.org/10.3390/cells12182315 - 19 Sep 2023
Viewed by 5113
Abstract
In arteries and arterioles, a chronic increase in blood pressure raises wall tension. This continuous biomechanical strain causes a change in gene expression in vascular smooth muscle cells (VSMCs) that may lead to pathological changes. Here we have characterised the functional properties of [...] Read more.
In arteries and arterioles, a chronic increase in blood pressure raises wall tension. This continuous biomechanical strain causes a change in gene expression in vascular smooth muscle cells (VSMCs) that may lead to pathological changes. Here we have characterised the functional properties of lipoma-preferred partner (LPP), a Lin11–Isl1–Mec3 (LIM)-domain protein, which is most closely related to the mechanotransducer zyxin but selectively expressed by smooth muscle cells, including VSMCs in adult mice. VSMCs isolated from the aorta of LPP knockout (LPP-KO) mice displayed a higher rate of proliferation than their wildtype (WT) counterparts, and when cultured as three-dimensional spheroids, they revealed a higher expression of the proliferation marker Ki 67 and showed greater invasion into a collagen gel. Accordingly, the gelatinase activity was increased in LPP-KO but not WT spheroids. The LPP-KO spheroids adhering to the collagen gel responded with decreased contraction to potassium chloride. The relaxation response to caffeine and norepinephrine was also smaller in the LPP-KO spheroids than in their WT counterparts. The overexpression of zyxin in LPP-KO VSMCs resulted in a reversal to a more quiescent differentiated phenotype. In native VSMCs, i.e., in isolated perfused segments of the mesenteric artery (MA), the contractile responses of LPP-KO segments to potassium chloride, phenylephrine or endothelin-1 did not vary from those in isolated perfused WT segments. In contrast, the myogenic response of LPP-KO MA segments was significantly attenuated while zyxin-deficient MA segments displayed a normal myogenic response. We propose that LPP, which we found to be expressed solely in the medial layer of different arteries from adult mice, may play an important role in controlling the quiescent contractile phenotype of VSMCs. Full article
(This article belongs to the Special Issue Role of Vascular Smooth Muscle Cells in Cardiovascular Disease)
Show Figures

Figure 1

Figure 1
<p>Distribution of (<b>a</b>) LPP and (<b>b</b>) zyxin in cultured VSMCs isolated from the aorta of 3-month-old C57BL/6 mice. LPP and zyxin immunoreactivity is shown in red, and stress fibres stained with an anti-α-SMA antibody are shown in green. Nuclei were counterstained with DAPI (blue). (<b>c</b>) LPP staining in LPP-KO VSMCs to control for the specificity of the antibody, and (<b>d</b>) zyxin staining in the LPP-KO VSMCs (<b>d</b>). Representative images; the scale bar shown in d represents 20 µm.</p>
Full article ">Figure 2
<p>Subcellular distribution of zyxin and LPP in cultured VSMCs (WT) in vitro. (<b>a</b>) Indirect immunofluorescence confocal laser scanning images of VSCM cultured for 48 h under static conditions (static), followed by 1 h or 8 h of exposure to cyclic stretching (13% elongation, 0.5 Hz). Inserts: Overlay of zyxin- or LPP-specific (purple) and α-smooth muscle actin (SMA)-specific (green) antibody staining; nuclei are additionally stained with DAPI (blue). Narrow arrow heads: focal adhesions, arrows: (position of) SMA-containing fibres; scale bar: 20 μm. (<b>b</b>) Representative Western blot analysis of cytosolic and nuclear fractions of VSMCs from 3-month-old WT mice pre-cultured for 48 h followed by 24 h incubation under static conditions or 8 and 24 h of exposure to cyclic stretching (13% elongation, 0.5 Hz), respectively. Alpha-tubulin and histone H3 (HH3) served as indicators for the purity of the subcellular fractions and as loading controls. (<b>c</b>) Statistical summary of 3 independent experiments with individual preparations of VSMCs isolated from the aorta of the 3-month-old WT mice. * <span class="html-italic">p</span> ˂ 0.05 as indicated.</p>
Full article ">Figure 3
<p>Proliferation of WT and LPP-KO VSMCs. (<b>a</b>) The rate of proliferation was determined by counting the number of cells per optical field of view at 0 and 72 h. Each data point represents an individual VSMC’s preparation with &gt;6 separate optical fields of view analysed. n = 4, * <span class="html-italic">p</span> ˂ 0.05 as indicated. (<b>b</b>) Percentage of WT and LPP-KO VSMCs grown in 3D spheroids expressing the proliferation marker Ki67. n = 4, * <span class="html-italic">p</span> &lt; 0.05 as indicated. (<b>c</b>) Transient overexpression of zyxin and eGFP in LPP-KO VSMCs and their effect on the rate of proliferation. Each data point represents an individual VSMC’s preparation with &gt;6 fields of view analysed. n = 5, * <span class="html-italic">p</span> &lt; 0.05 as indicated.</p>
Full article ">Figure 4
<p>LPP-KO VSMCs and WT VSMCs grown in 3D spheroids embedded in collagen gels form sprouts and invade the collagen gel. (<b>a</b>) Representative images, the scale bar represents 50 µm. (<b>b</b>) Summary of the average number of sprouts produced by individual LPP KO VSMC (n = 6, in red) and WT VSMC preparations (n = 4) with &gt;10 spheroids per VSMC preparation analysed using the cellSens software. (<b>c</b>) Summary of the cumulative distance travelled by individual sprouts of a VSMC spheroid. n = 6 and n = 4 as in (<b>b</b>), * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 as indicated in (<b>b</b>,<b>c</b>) Transient overexpression of zyxin and eGFP in LPP-KO VSMCs and their effect on (<b>d</b>) the average number of sprouts and (<b>e</b>) cumulative sprout length. Each data point represents an individual VSMC preparation with &gt;10 spheroids per VSMC preparation analysed. n = 3, * <span class="html-italic">p</span> ˂ 0.05 as indicated.</p>
Full article ">Figure 5
<p>Comparison of the contractile properties of the VSMCs grown as 3D spheroids attached to a collagen gel. (<b>a</b>) The contractile response of the VSMC spheroids to potassium chloride (KCl, 60 mM) is calculated as reduction in the area of the collagen gel covered by the spheroids. (<b>b</b>) Relaxation response of the VSMC spheroids to caffeine (10 mM) and to norepinephrine (NE, 10 μM) (<b>c</b>). Each column represents the average responses of VSMC spheroids with &gt;5 spheroids analysed per preparation and a total of 6 VSMCs preparations from individual animals per group (LPP KO or WT). * <span class="html-italic">p</span> ˂ 0.05 as indicated.</p>
Full article ">Figure 6
<p>Immunofluorescence analysis of LPP and zyxin in 3rd-order mesenteric artery (MA) and femoral artery (FA) segments isolated from WT, LPP-KO or ZXY-KO mice. (<b>a</b>–<b>d</b>) Representative images for LPP ((<b>a</b>,<b>c</b>); red) and zyxin ((<b>b</b>,<b>d</b>); red) in MA and FA segments isolated from WT mice. To properly localize both LIM domain proteins, an anti-α-SMA antibody was used to stain the VSMCs in the media ((<b>e</b>–<b>h</b>), green) and an anti-CD31 antibody was used to visualize the endothelial cells lining the inner lumen of these segments ((<b>a</b>–<b>d</b>), green). While there is a clear colocalization of zyxin with the endothelium (<b>b</b>,<b>d</b>), LPP solely colocalizes with the medial VSMCs (<b>e</b>,<b>f</b>) with which zyxin also colocalizes but to a smaller extent. Staining of both LIM domain proteins in MA segments isolated from the respective knockout mice confirmed the specificity of the antibodies used (<b>i</b>,<b>j</b>). LPP ((<b>k</b>), red) and zyxin ((<b>m</b>), red) were also stained in segments of the aorta isolated from WT mice. Here, co-staining of LPP and CD31 ((<b>l</b>), green) confirmed their mutually exclusive localization, whereas co-staining of zyxin and α-SMA ((<b>n</b>), green) revealed that zyxin is highly abundant in the media of this large conduit artery (yellow colour indicates prominent colocalization with the VSMCs in the media) as well as a distinct staining of the endothelium. Scale bars represent 10 μm in (<b>a</b>–<b>c</b>,<b>f</b>) and 20 μm in (<b>d</b>,<b>e</b>,<b>g</b>–<b>n</b>).</p>
Full article ">Figure 7
<p>Comparison of the contractile response of 3rd-order mesenteric artery segments isolated from WT, LPP-KO or ZYX-KO mice to different vasoactive stimuli. The number of isolated perfused segments derived from individual mice is indicated on the graphs. Cumulative concentration–response curves for (<b>a</b>,<b>b</b>) phenylephrine (PE), (<b>c</b>,<b>d</b>) potassium chloride (KCl) and (<b>e</b>,<b>f</b>) endothelin-1 (ET-1).</p>
Full article ">Figure 8
<p>Comparison of the myogenic responses of 3rd-order mesenteric artery segments isolated from WT, LPP-KO or ZYX-KO mice. The number of isolated perfused segments derived from individual mice is indicated on the graphs. (<b>a</b>) Pressure–response curves for segments of (<b>a</b>) 3-month-old, (<b>c</b>) 6-month-old and (<b>e</b>) 12-month-old WT and LPP-KO mice; (<b>b</b>) 6-month-old and (<b>d</b>) 12-month-old WT and ZYX-KO mice; and (<b>f</b>) 6-month-old WT and LPP-KO mice made hypertensive using the DOCA-salt model of experimental hypertension. The area under the curve (AUC) was calculated for each individual group. * <span class="html-italic">p</span> ˂ 0.05, ** <span class="html-italic">p</span> ˂ 0.01 as indicated.</p>
Full article ">Figure 9
<p>Susceptibility of the myogenic response of 3rd-order mesenteric artery segments isolated from WT or LPP-KO mice made hypertensive by employing the DOCA-salt model of experimental hypertension. The number of isolated perfused segments derived from individual mice is indicated on the graphs. Effect of the general gelatinase inhibitor GM6001 (0.5 µM) on (<b>a</b>) MA segments derived from (<b>a</b>) hypertensive WT and (<b>b</b>) LPP-KO mice and of the TRPC3 channel blocker Pyr 3 (3 µM) on MA segments isolated from (<b>c</b>) hypertensive WT and (<b>d</b>) LPP-KO mice. The area under the curve (AUC) was calculated for each individual group. * <span class="html-italic">p</span> ˂ 0.05, ** <span class="html-italic">p</span> ˂ 0.01 as indicated.</p>
Full article ">Figure 10
<p>Passive pressure–diameter curves of 3rd-order MA segments isolated from WT, LPP-KO or ZYX-KO mice. The curves were generated in Ca<sup>2+</sup>-free physiological saline solution in the presence of EGTA. The number of isolated perfused segments derived from individual mice is indicated on the graph. ** <span class="html-italic">p</span> ˂ 0.01 as indicated.</p>
Full article ">Figure 11
<p>Passive wall thickness of 3rd-order MA segments isolated from (<b>a</b>) 6- and (<b>b</b>) 12-month-old WT, LPP-KO or ZYX-KO mice. Wall thickness was determined at 80 mm Hg in Ca<sup>2+</sup>-free physiological saline solution in the presence of EGTA (passive distention). The number of isolated perfused segments derived from individual mice is indicated on the graph. * <span class="html-italic">p</span> ˂ 0.05 as indicated.</p>
Full article ">
44 pages, 552 KiB  
Review
Inborn Errors of Metabolism with Ataxia: Current and Future Treatment Options
by Tatiana Bremova-Ertl, Jan Hofmann, Janine Stucki, Anja Vossenkaul and Matthias Gautschi
Cells 2023, 12(18), 2314; https://doi.org/10.3390/cells12182314 - 19 Sep 2023
Viewed by 2012
Abstract
A number of hereditary ataxias are caused by inborn errors of metabolism (IEM), most of which are highly heterogeneous in their clinical presentation. Prompt diagnosis is important because disease-specific therapies may be available. In this review, we offer a comprehensive overview of metabolic [...] Read more.
A number of hereditary ataxias are caused by inborn errors of metabolism (IEM), most of which are highly heterogeneous in their clinical presentation. Prompt diagnosis is important because disease-specific therapies may be available. In this review, we offer a comprehensive overview of metabolic ataxias summarized by disease, highlighting novel clinical trials and emerging therapies with a particular emphasis on first-in-human gene therapies. We present disease-specific treatments if they exist and review the current evidence for symptomatic treatments of these highly heterogeneous diseases (where cerebellar ataxia is part of their phenotype) that aim to improve the disease burden and enhance quality of life. In general, a multimodal and holistic approach to the treatment of cerebellar ataxia, irrespective of etiology, is necessary to offer the best medical care. Physical therapy and speech and occupational therapy are obligatory. Genetic counseling is essential for making informed decisions about family planning. Full article
(This article belongs to the Special Issue Emerging Therapies for Hereditary Ataxia)
26 pages, 4596 KiB  
Article
Let-7g Upregulation Attenuated the KRAS–PI3K–Rac1–Akt Axis-Mediated Bioenergetic Functions
by Kuang-Chen Hung, Ni Tien, Da-Tian Bau, Chun-Hsu Yao, Chan-Hung Chen, Jiun-Long Yang, Meng-Liang Lin and Shih-Shun Chen
Cells 2023, 12(18), 2313; https://doi.org/10.3390/cells12182313 - 19 Sep 2023
Cited by 1 | Viewed by 1673
Abstract
The aberrant activation of signaling pathways contributes to cancer cells with metabolic reprogramming. Thus, targeting signaling modulators is considered a potential therapeutic strategy for cancer. Subcellular fractionation, coimmunoprecipitation, biochemical analysis, and gene manipulation experiments revealed that decreasing the interaction of kirsten rat sarcoma [...] Read more.
The aberrant activation of signaling pathways contributes to cancer cells with metabolic reprogramming. Thus, targeting signaling modulators is considered a potential therapeutic strategy for cancer. Subcellular fractionation, coimmunoprecipitation, biochemical analysis, and gene manipulation experiments revealed that decreasing the interaction of kirsten rat sarcoma viral oncogene homolog (KRAS) with p110α in lipid rafts with the use of naringenin (NGN), a citrus flavonoid, causes lipid raft-associated phosphatidylinositol 3-kinase (PI3K)−GTP-ras-related C3 botulinum toxin substrate 1 (Rac1)−protein kinase B (Akt)-regulated metabolic dysfunction of glycolysis and mitochondrial oxidative phosphorylation (OXPHOS), leading to apoptosis in human nasopharyngeal carcinoma (NPC) cells. The use of lethal-7g (let-7g) mimic and let-7g inhibitor confirmed that elevated let-7g resulted in a decrease in KRAS expression, which attenuated the PI3K−Rac1−Akt−BCL-2/BCL-xL-modulated mitochondrial energy metabolic functions. Increased let-7g depends on the suppression of the RNA-specificity of monocyte chemoattractant protein-induced protein-1 (MCPIP1) ribonuclease since NGN specifically blocks the degradation of pre-let-7g by NPC cell-derived immunoprecipitated MCPIP1. Converging lines of evidence indicate that the inhibition of MCPIP1 by NGN leads to let-7g upregulation, suppressing oncogenic KRAS-modulated PI3K–Rac1–Akt signaling and thereby impeding the metabolic activities of aerobic glycolysis and mitochondrial OXPHOS. Full article
(This article belongs to the Section Intracellular and Plasma Membranes)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Dysfunction of glycolysis and mitochondrial oxidative phosphorylation (OXPHOS) by naringenin (NGN) confers apoptosis in nasopharyngeal carcinoma (NPC) cells. (<b>A</b>) The effect of NGN on NPC and Smulow–Glickman (S-G) cell growth. Cells treated with vehicle (−) or the indicated concentrations of NGN for 36 h. The 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide (MTT) assay determined cell growth. (<b>B</b>–<b>F</b>) The effects of NGN on the induction of NPC cell growth inhibition and apoptosis. After 36 h treatment with (−), NGN (160 μM), Z-VAD-FMK (8 μM) or NGN (160 μM) and Z-VAD-FMK (8 μM), cell growth and viability were determined by the MTT and flow cytometric analysis of PI uptake, respectively. DNA fragmentation was determined using a Cell Death Detection enzyme-linked immunosorbent assay (ELISA) kit. Annexin V-biotinylated vehicle- or NGN-treated cells were fractionated by subcellular fractionation centrifugation to isolate the plasma membrane (M) fraction. The levels of the indicated proteins in the lysates of (−)-, NGN-, -Z-VAD-FMK, and the NGN plus Z-VAD-FMK co-treated M fraction were determined by Western blot analysis using streptavidin-horseradish peroxidase (HRP) and specific antibody to Annexin V or cadherin. Antibody against cadherin was used as an internal control for the plasma membrane. The levels of phosphorylated histone H2A.X (Ser 139) (p-γ-H2AX (Ser 139)), H2AX, and caspase-3 in the total cell (T) lysates were determined by Western blot analysis with specific antibodies. β-Actin was used as an internal control for sample loading. (<b>G</b>,<b>H</b>) The effects of NGN on glycolysis and mitochondrial oxidative phosphorylation (OXPHOS). Cells treated with vehicle or NGN (160 μM) for the indicated periods. The oxygen consumption rate (OCR) was measured in the presence of oligomycin-A (Oligo-A) (1 μM), carbonylcyanide-p-trifluoromethoxyphenylhydrazone (FCCP) (0.5 μM), and rotenone (Rot) (30 μM) plus myxothiazol (Myx) (10 μM) at the indicated time points. The extracellular acidification rate (ECAR) was examined under the sequential addition of glucose, Oligo-A (1 μM), and 2-deoxy-D-glucose (2-DG) (25 mM) at the indicated time points. The OCR and ECAR were measured using a Seahorse Bioscience XF24 Analyzer. (<b>I</b>–<b>M</b>) The effects of NGN on the levels of glucose uptake, pyruvate, lactate, ATP, and mitochondrial DNA (mtDNA) copy number. After 36 h of treatment with (−) or NGN (160 μM), the pyruvate, lactate, and ATP values were analyzed using the Pyruvate Assay kit, Lactate Assay kit, and ATP-based CellTiter-Glo Luminescent Cell Viability kit, respectively. Glucose uptake was measured using the Glucose Uptake Colorimetric assay kit. The level of mtDNA was analyzed using quantitative real-time PCR (qRT-PCR). The mtDNA expression was determined relative to that of β-actin. The values are presented as three independent experiments’ mean ± standard error. * <span class="html-italic">p</span> &lt; 0.05: significantly different from vehicle- or NGN-treated cells.</p>
Full article ">Figure 2
<p>Decreased level of kirsten rat sarcoma viral oncogene homolog (KRAS) in the lipid rafts causes attenuation in the formation of lipid raft-associated KRAS–phosphatidylinositol 3-kinase (PI3K)–active GTP-binding Ras-related C3 botulinum toxin substrate (GTP-Rac1) complexes. (<b>A</b>) Nasopharyngeal carcinoma (NPC) cells were treated with vehicle (−) or naringenin (NGN) for 36 h. Detergent-resistant membrane (DRM) and detergent-soluble (DS) fractions were prepared by flotation along a sucrose density gradient. The levels of the indicated proteins in the lysates of (−)- or NGN-treated DRM and DS fractions and total cell (T) lysates were determined by Western blot analysis using specific antibodies. Antibodies against caveolin-1/CD55 and CD71 were used as internal controls for DRM and DS fractions. (<b>B</b>,<b>C</b>) After 36 h of treatment with (−) or NGN, DRM fractions were prepared by flotation along a sucrose density gradient. The antibody used for coimmunoprecipitation is indicated at the top. The proteins from the immunoprecipitated complexes were detected using Western blotting with specific antibodies. Normal IgG was used as a control for antibody specificity. Total and DRM lysates from (−)- or NGN-treated cells were used to monitor the indicated protein levels and were determined using Western blot analysis with specific antibodies.</p>
Full article ">Figure 3
<p>Disruption of kirsten rat sarcoma viral oncogene homolog (KRAS)–phosphatidylinositol 3-kinase (PI3K)–active GTP-binding Ras-related C3 botulinum toxin substrate (GTP-Rac1) complex formation and impaired energetic synthesis of the glycolysis and mitochondrial respiration pathways. (<b>A</b>–<b>C</b>) At 12 h after transfection with hemagglutinin (HA)-KRAS, GFP short hairpin RNA (shRNA), or KRAS shRNA, cells were treated with vehicle (−) or naringenin (NGN) for 36 h. The levels of the indicated proteins in the lysates of the total cell (T), detergent-resistant membrane (DRM), and detergent-soluble (DS) fractions were determined by Western blot analysis using specific antibodies. Co-immunoprecipitation of KRAS, p85α, p110α, and GTP-Rac1 was performed using the DRM fractions prepared from the cells treated as described above. The KRAS antibody used for co-immunoprecipitation is indicated at the top. The proteins from the immunoprecipitated complexes were detected using Western blotting with specific antibodies. Normal IgG was used as a control for antibody specificity. (<b>D</b>,<b>E</b>) Oxygen consumption rate (OCR) was measured in the presence of oligomycin-A (Oligo-A), carbonylcyanide-p-trifluoromethoxyphenylhydrazone (FCCP), and rotenone (Rot) plus myxothiazol (Myx) at the indicated time points. Extracellular acidification rate (ECAR) was examined in the sequential addition of glucose, Oligo-A, and 2-deoxy-D-glucose (2-DG) at the indicated time points. The OCR and ECAR were measured using a Seahorse Bioscience XF24 Analyzer. (<b>F</b>–<b>I</b>) The pyruvate, lactate, and ATP values were analyzed using the Pyruvate Assay kit, Lactate Assay kit, and ATP-based CellTiter-Glo Luminescent Cell Viability kit, respectively. In addition, glucose uptake was measured using the Glucose Uptake Colorimetric assay kit. The values are presented as three independent experiments’ mean ± standard error. * <span class="html-italic">p</span> &lt; 0.05: significantly different from vehicle-treated empty vector-transfected, vehicle-treated GFP-transfected, or NGN-treated empty vector-transfected cells.</p>
Full article ">Figure 4
<p>Metabolic dysfunction of glycolysis and mitochondrial oxidative phosphorylation associated with <span class="html-italic">lethal-7g</span> (<span class="html-italic">let-7g</span>)-attenuated kirsten rat sarcoma viral oncogene homolog (KRAS)–phosphatidylinositol 3-kinase (PI3K)–Ras-related C3 botulinum toxin substrate (Rac1)–protein kinase B (Akt) signaling. (<b>A</b>) Cells were treated with vehicle (−) or naringenin (NGN) for 36 h. The expression of <span class="html-italic">let-7</span> was determined using quantitative real-time polymerase chain reaction (qRT-PCR). The <span class="html-italic">let-7</span> value was normalized to the U6 level. The Y-axis shows the denary logarithm of the normalized <span class="html-italic">let-7</span> copy number. (<b>B</b>,<b>C</b>) The effects of the protein kinase C (PKC) inhibitor chelerythrine (CHE) on <span class="html-italic">let-7g</span> and <span class="html-italic">OCT-1</span> expression. After treating with CHE (0.5 μM) for 36 h, the relative expression levels of <span class="html-italic">let-7g</span> and <span class="html-italic">OCT-1</span> were determined by qRT-PCR. The <span class="html-italic">let-7g</span> and <span class="html-italic">OCT-1</span> values were normalized to the U6 level and β-actin, respectively. The Y-axis shows the denary logarithm of the normalized <span class="html-italic">let-7g</span> or <span class="html-italic">OCT-1</span> copy number. (<b>D</b>–<b>F</b>) At 12 h after transfection with the negative (N) mimic control, <span class="html-italic">let-7g</span> mimic, (N) mimic control inhibitor, or <span class="html-italic">let-7g</span> inhibitor, cells were treated with (−) or NGN for 36 h. Five mM bismaleimidohexane (BMH)-treated cells were subjected to subcellular fractionation to obtain the mitochondrial (Mt) and endoplasmic reticulum (ER)/microsomal (Ms) fractions. In total, 20 μg of total protein from the recovered fractions was analyzed by 10% sodium dodecyl sulfate polyacrylamide gel electrophoresis (SDS-PAGE) and probed with specific antibodies, as indicated. Cytochrome <span class="html-italic">c</span> oxidase subunit II (Cox-2), calnexin, and α-tubulin were used as internal controls for the mitochondria, ER, and cytosol, respectively. The levels of the indicated proteins in the total cell (T) lysates were determined by Western blot analysis using specific antibodies. (<b>G</b>,<b>H</b>) At 12 h after transfection with the (N) mimic control, <span class="html-italic">let-7g</span> mimic, (N) mimic control inhibitor, or <span class="html-italic">let-7g</span> inhibitor, cells were treated with (−) or NGN in the presence of 1 μM oligomycin-A (Oligo-A), carbonylcyanide-p-trifluoromethoxyphenylhydrazone (FCCP), and rotenone (Rot) plus myxothiazol (Myx) or the sequential addition of glucose, Oligo-A, and 2-DG at the indicated time points. The oxygen consumption rate (OCR) and extracellular acidification rate (ECAR) were measured using a Seahorse Bioscience XF24 Analyzer. (<b>I</b>,<b>J</b>) At 12 h after transfection with the (N) mimic control, <span class="html-italic">let-7g</span> mimic, (N) mimic control inhibitor, or <span class="html-italic">let-7g</span> inhibitor, cells were treated with vehicle (−), NGN, NGN plus Mitoquinone (MitoQ) (10 μM), decyltriphenylphosphonium bromide (DecylTPP) (1 μM), NGN plus dantrolene (25 μM), or NGN plus ruthenium red (1 μM) for 36 h. Flow cytometry determined the levels of mitochondrial reactive oxygen species (ROS) and cytosolic calcium (Ca<sup>++</sup>) by measuring the increased fluorescence. (<b>K</b>–<b>O</b>) Transfected cells were harvested 36 h after treatment with (−) or NGN. The glucose, pyruvate, lactate, and ATP values were analyzed using the Glucose assay, Pyruvate Assay kit, Lactate Assay kit, and ATP-based CellTiter-Glo Luminescent Cell Viability kit, respectively. In addition, glucose uptake was measured using the Glucose Uptake Colorimetric Assay kit. The values are presented as three independent experiments’ mean ± standard error. * <span class="html-italic">p</span> &lt; 0.05: significantly different from vehicle-treated empty vector-transfected, vehicle-treated (N) mimic control-transfected, or vehicle-treated (N) mimic control inhibitor-transfected cells.</p>
Full article ">Figure 4 Cont.
<p>Metabolic dysfunction of glycolysis and mitochondrial oxidative phosphorylation associated with <span class="html-italic">lethal-7g</span> (<span class="html-italic">let-7g</span>)-attenuated kirsten rat sarcoma viral oncogene homolog (KRAS)–phosphatidylinositol 3-kinase (PI3K)–Ras-related C3 botulinum toxin substrate (Rac1)–protein kinase B (Akt) signaling. (<b>A</b>) Cells were treated with vehicle (−) or naringenin (NGN) for 36 h. The expression of <span class="html-italic">let-7</span> was determined using quantitative real-time polymerase chain reaction (qRT-PCR). The <span class="html-italic">let-7</span> value was normalized to the U6 level. The Y-axis shows the denary logarithm of the normalized <span class="html-italic">let-7</span> copy number. (<b>B</b>,<b>C</b>) The effects of the protein kinase C (PKC) inhibitor chelerythrine (CHE) on <span class="html-italic">let-7g</span> and <span class="html-italic">OCT-1</span> expression. After treating with CHE (0.5 μM) for 36 h, the relative expression levels of <span class="html-italic">let-7g</span> and <span class="html-italic">OCT-1</span> were determined by qRT-PCR. The <span class="html-italic">let-7g</span> and <span class="html-italic">OCT-1</span> values were normalized to the U6 level and β-actin, respectively. The Y-axis shows the denary logarithm of the normalized <span class="html-italic">let-7g</span> or <span class="html-italic">OCT-1</span> copy number. (<b>D</b>–<b>F</b>) At 12 h after transfection with the negative (N) mimic control, <span class="html-italic">let-7g</span> mimic, (N) mimic control inhibitor, or <span class="html-italic">let-7g</span> inhibitor, cells were treated with (−) or NGN for 36 h. Five mM bismaleimidohexane (BMH)-treated cells were subjected to subcellular fractionation to obtain the mitochondrial (Mt) and endoplasmic reticulum (ER)/microsomal (Ms) fractions. In total, 20 μg of total protein from the recovered fractions was analyzed by 10% sodium dodecyl sulfate polyacrylamide gel electrophoresis (SDS-PAGE) and probed with specific antibodies, as indicated. Cytochrome <span class="html-italic">c</span> oxidase subunit II (Cox-2), calnexin, and α-tubulin were used as internal controls for the mitochondria, ER, and cytosol, respectively. The levels of the indicated proteins in the total cell (T) lysates were determined by Western blot analysis using specific antibodies. (<b>G</b>,<b>H</b>) At 12 h after transfection with the (N) mimic control, <span class="html-italic">let-7g</span> mimic, (N) mimic control inhibitor, or <span class="html-italic">let-7g</span> inhibitor, cells were treated with (−) or NGN in the presence of 1 μM oligomycin-A (Oligo-A), carbonylcyanide-p-trifluoromethoxyphenylhydrazone (FCCP), and rotenone (Rot) plus myxothiazol (Myx) or the sequential addition of glucose, Oligo-A, and 2-DG at the indicated time points. The oxygen consumption rate (OCR) and extracellular acidification rate (ECAR) were measured using a Seahorse Bioscience XF24 Analyzer. (<b>I</b>,<b>J</b>) At 12 h after transfection with the (N) mimic control, <span class="html-italic">let-7g</span> mimic, (N) mimic control inhibitor, or <span class="html-italic">let-7g</span> inhibitor, cells were treated with vehicle (−), NGN, NGN plus Mitoquinone (MitoQ) (10 μM), decyltriphenylphosphonium bromide (DecylTPP) (1 μM), NGN plus dantrolene (25 μM), or NGN plus ruthenium red (1 μM) for 36 h. Flow cytometry determined the levels of mitochondrial reactive oxygen species (ROS) and cytosolic calcium (Ca<sup>++</sup>) by measuring the increased fluorescence. (<b>K</b>–<b>O</b>) Transfected cells were harvested 36 h after treatment with (−) or NGN. The glucose, pyruvate, lactate, and ATP values were analyzed using the Glucose assay, Pyruvate Assay kit, Lactate Assay kit, and ATP-based CellTiter-Glo Luminescent Cell Viability kit, respectively. In addition, glucose uptake was measured using the Glucose Uptake Colorimetric Assay kit. The values are presented as three independent experiments’ mean ± standard error. * <span class="html-italic">p</span> &lt; 0.05: significantly different from vehicle-treated empty vector-transfected, vehicle-treated (N) mimic control-transfected, or vehicle-treated (N) mimic control inhibitor-transfected cells.</p>
Full article ">Figure 5
<p><span class="html-italic">Lethal-7g</span> (<span class="html-italic">let-7g</span>) upregulation attenuates glucose transporter-1 (GLUT-1) lipid raft membrane-targeting without affecting hypoxia-inducible factor 1α (HIF-1α)-mediated pyruvate kinase type M2 (PKM2), pyruvate dehydrogenase kinase 1 (PDK1), hexokinase II (HK-II), lactate dehydrogenase (LDH), and succinate dehydrogenase (SDH) activities. (<b>A</b>–<b>I</b>) At 12 h after transfection with the negative (N) mimic control, <span class="html-italic">let-7g</span> mimic, (N) mimic control inhibitor, or <span class="html-italic">let-7g</span> inhibitor, cells treated with vehicle (−) or naringenin (NGN) for 36 h. The levels of the indicated proteins in the total cell (T) lysates or detergent-resistant membrane (DRM) fractions were determined by Western blot analysis using specific antibodies. The FKM2, PDK1, HK-II, LDH, and SDH activities were analyzed using the Colorimetric-Based Pyruvate Kinase Activity Assay, ADP-GloTM Kinase Assay, Colorimetric Hexokinase Activity Assay, Lactate Dehydrogenase Assay, and Succinate Dehydrogenase Activity Colorimetric Assay, respectively. The Succinate Colorimetric Assay kit determined the succinate level. β-Actin was used as an internal control for sample loading. * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 6
<p>Suppression of monocyte chemoattractant protein-induced protein-1 (MCPIP1)-mediated degradation of <span class="html-italic">lethal-7g</span> (<span class="html-italic">let-7g</span>) by naringenin (NGN) involved the attenuation of glycolysis and mitochondrial mitochondrial oxidative phosphorylation (OXPHOS). (<b>A</b>,<b>B</b>) Immunoprecipitated MCPIP1 (IP:MCPIP1) from the nasopharyngeal carcinoma (NPC) or Smulow–Glickman (S-G) cells and recombinant purified MCPIP1 (rMCPIP1) used in the experiments. The inhibition of the <span class="html-italic">let-7g</span> degradation of MCPIP1-mediated by NGN, performed by in vitro RNA cleavage assay and analyzed by Northern blotting (NB). The levels of the IP:MCPIP1 or rMCPIP1 in reactions were determined by Western blot (WB) analysis using specific antibodies. (<b>C</b>–<b>E</b>) At 12 h after transfection with FLAG-MCPIP1, FLAG-MCPIP1 (D141N), or MCPIP1 short hairpin RNA (shRNA), cells were treated with vehicle (−) or NGN in the presence of 1 μM oligomycin-A (Oligo-A), carbonylcyanide-p-trifluoromethoxyphenylhydrazone (FCCP), and rotenone (Rot) plus myxothiazol (Myx) or the sequential addition of glucose, Oligo-A, and 2-deoxy-D-glucose (2-DG) at the indicated time points. The oxygen consumption rate (OCR) and extracellular acidification rate (ECAR) were measured using a Seahorse Bioscience XF24 Analyzer. Western blot analysis using specific antibodies determined the levels of the indicated proteins in the total cell lysates. Quantitative real-time polymerase chain reaction (qRT-PCR) determined the relative expression level of <span class="html-italic">let-7g.</span> The <span class="html-italic">let-7g</span> value was normalized to the U6 level. The Y-axis shows the denary logarithm of the normalized <span class="html-italic">let-7g</span> copy number. The values are presented as three independent experiments’ mean ± standard error. * <span class="html-italic">p</span> &lt; 0.05: significantly different from (−)-treated empty vector-transfected, NGN-treated empty vector-transfected, (−)-treated FLAG-MCPIP1-transfected, or (−)-treated MCPIP1 shRNA-transfected cells.</p>
Full article ">Figure 7
<p>A molecular model for the naringenin (NGN)-induced impairment of the mitochondria-mediated bioenergetics in nasopharyngeal carcinoma (NPC) cells. (<b>A</b>) The selective interaction of argonaute (AG) protein or an undefined factor with monocyte chemoattractant protein-induced protein-1 (MCPIP1) may accomplish the sequence-specific targeting of <span class="html-italic">pre-lethal-7g</span> (<span class="html-italic">pre-let-7g</span>) and promote MCPIP1-mediated <span class="html-italic">pre-let-7g</span> degradation to decrease the biogenesis of <span class="html-italic">let-7g</span>, resulting in the attenuation of the <span class="html-italic">let-7g</span>-mediated translational repression of kirsten rat sarcoma viral oncogene homolog (<span class="html-italic">KRAS</span>) mRNA. The resultant elevated KRAS increases the formation of clustered KRAS- phosphatidylinositol 3-kinase (PI3K)-active GTP-binding Ras-related C3 botulinum toxin substrate (GTP-Rac1)–protein kinase B (Akt) signaling molecules in the lipid raft membranes, constituting a central element in the initiation of the coordination of glycolysis with mitochondrial oxidative phosphorylation (OXPHOS) for ATP generation. (<b>B</b>) Under the condition of the cellular uptake of NGN, NGN may bind to the AG protein or an undefined cofactor to block the degradation of <span class="html-italic">pre-let-7g</span> by MCPIP1, increasing the level of <span class="html-italic">let-7g</span>, thus inducing the <span class="html-italic">let-7g</span>-mediated translational repression of <span class="html-italic">KRAS</span> mRNA and thereby dismissing the interaction between KRAS and p110α in the lipid raft membrane. The absence of KRAS in the p85α–p110α complexes causes the destabilization of p85α–p110α complexes in the lipid raft membrane. The resultant loss of the KRAS–p85α–p110α complexes in the lipid raft membrane leads to blocking PI3K-GTP-Rac1-mediated Akt activation. Attenuated Akt impaired the aerobic glycolysis and mitochondria-regulated bioenergetic functions in NPC cells.</p>
Full article ">
32 pages, 5221 KiB  
Review
The Chemical Inhibitors of Endocytosis: From Mechanisms to Potential Clinical Applications
by Olga Klaudia Szewczyk-Roszczenko, Piotr Roszczenko, Anna Shmakova, Nataliya Finiuk, Serhii Holota, Roman Lesyk, Anna Bielawska, Yegor Vassetzky and Krzysztof Bielawski
Cells 2023, 12(18), 2312; https://doi.org/10.3390/cells12182312 - 19 Sep 2023
Cited by 10 | Viewed by 5528
Abstract
Endocytosis is one of the major ways cells communicate with their environment. This process is frequently hijacked by pathogens. Endocytosis also participates in the oncogenic transformation. Here, we review the approaches to inhibit endocytosis, discuss chemical inhibitors of this process, and discuss potential [...] Read more.
Endocytosis is one of the major ways cells communicate with their environment. This process is frequently hijacked by pathogens. Endocytosis also participates in the oncogenic transformation. Here, we review the approaches to inhibit endocytosis, discuss chemical inhibitors of this process, and discuss potential clinical applications of the endocytosis inhibitors. Full article
(This article belongs to the Section Cellular Pathology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Clathrin-mediated endocytosis.</p>
Full article ">Figure 2
<p>Caveolae-dependent endocytosis.</p>
Full article ">Figure 3
<p>CLIC/GEEC endocytosis.</p>
Full article ">Figure 4
<p>IL2Rβ pathway.</p>
Full article ">Figure 5
<p>Arf6-dependent endocytosis.</p>
Full article ">Figure 6
<p>Flotillin-dependent endocytosis.</p>
Full article ">Figure 7
<p>Phagocytosis.</p>
Full article ">Figure 8
<p>Micropinocytosis.</p>
Full article ">Figure 9
<p>FEME endocytosis.</p>
Full article ">Figure 10
<p>Structures of clathrin inhibitors.</p>
Full article ">Figure 11
<p>Structures of dynamin inhibitors that affect the pleckstrin homology (PH) domain.</p>
Full article ">Figure 12
<p>Structures of dynamin inhibitors that inhibit the GTPase activity of dynamin, affect allosteric sites on the G domain, and those with unknown or complex mechanisms of action.</p>
Full article ">
23 pages, 7211 KiB  
Article
Unraveling the Role of Hepatic PGC1α in Breast Cancer Invasion: A New Target for Therapeutic Intervention?
by Kumar Ganesan, Cong Xu, Qingqing Liu, Yue Sui and Jianping Chen
Cells 2023, 12(18), 2311; https://doi.org/10.3390/cells12182311 - 19 Sep 2023
Cited by 3 | Viewed by 1159
Abstract
Breast cancer (BC) is the most common cancer among women worldwide and the main cause of cancer deaths in women. Metabolic components are key risk factors for the development of non-alcoholic fatty liver disease (NAFLD), which may promote BC. Studies have reported that [...] Read more.
Breast cancer (BC) is the most common cancer among women worldwide and the main cause of cancer deaths in women. Metabolic components are key risk factors for the development of non-alcoholic fatty liver disease (NAFLD), which may promote BC. Studies have reported that increasing PGC1α levels increases mitochondrial biogenesis, thereby increasing cell proliferation and metastasis. Moreover, the PGC1α/ERRα axis is a crucial regulator of cellular metabolism in various tissues, including BC. However, it remains unclear whether NAFLD is closely associated with the risk of BC. Therefore, the present study aimed to determine whether hepatic PGC1α promotes BC cell invasion via ERRα. Various assays, including ELISA, western blotting, and immunoprecipitation, have been employed to explore these mechanisms. According to the KM plot and TCGA data, elevated PGC1α expression was highly associated with a shorter overall survival time in patients with BC. High concentrations of palmitic acid (PA) promoted PGC1α expression, lipogenesis, and inflammatory processes in hepatocytes. Conditioned medium obtained from PA-treated hepatocytes significantly increased BC cell proliferation. Similarly, recombinant PGC1α in E0771 and MCF7 cells promoted cell proliferation, migration, and invasion in vitro. However, silencing PGC1α in both BC cell lines resulted in a decrease in this trend. As determined by immunoprecipitation assay, PCG1a interacted with ERRα, thereby facilitating the proliferation of BC cells. This outcome recognizes the importance of further investigations in exploring the full potential of hepatic PGC1α as a prognostic marker for BC development. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Overall survival of patients with BC correlates to the expression of PGC1α. (<b>B</b>) The survival curves from the KM plotter show that lower PPS levels are associated with reduced survival. (<b>C</b>) Cancer Genome Atlas (TCGA) database curves of menopausal status in BC patients with survival. According to the results, the postmenopausal group (n = 150, <span class="html-italic">p</span> = 0.022) exhibited increased expression of PGC1α. (<b>D</b>) clinical significance of PGC1α expression in patients with BC. * <span class="html-italic">p</span> &lt; 0.05. (<b>E</b>,<b>F</b>) PA (0.2, 0.4, and 0.8 mM) treatment promotes liver toxicity, which was tested for 12, 24, and 48 h using cell viability assay. The data were presented as means ± SD for three replicates. Data were analyzed with a one-way ANOVA followed by group comparisons using a post hoc Tukey’s multiple comparison test. (<b>G</b>) Exposure to PA results in intracellular lipid accumulation in hepatocytes. AML (<b>a</b>–<b>e</b>), and MIHA (<b>f</b>–<b>j</b>) were treated with PA (0.2, 04, 0.8 mM) for 24 h and were stained with Oil Red O, and nuclei were stained with hematoxylin, under microscopic observation at 20× magnification. Control (<b>a</b>,<b>f</b>), BSA (<b>b</b>,<b>g</b>), 0.2 mM PA (<b>c</b>,<b>h</b>), 0.4 mM PA (<b>d</b>,<b>i</b>), and 0.8 mM PA (<b>e</b>,<b>j</b>). Quantification of lipids in (<b>H</b>) AML12 and (<b>I</b>) MIHA using ORO staining. The data were presented as means ± SD for three replicates. Data were analyzed with a one-way ANOVA followed by group comparisons using a post hoc Tukey’s multiple comparison test. Abbreviation: OS, overall survival; PPS, post-progression survival; HR, hazard ratio; PA, palmitic acid.</p>
Full article ">Figure 2
<p>(<b>A</b>) Expression of PGC1α by Western Blot analysis. The respective hepatocytes were treated with PA (0.2, 0.4, and 0.8 mM) for 48 h, and the expression of PGC1α was dose-dependent. (<b>B</b>) The quantification of the target protein was calculated with a densitometer. The values are expressed as the fold of change (X basal), in mean ± SEM, where <span class="html-italic">n</span> = 3. * <span class="html-italic">p</span> &lt; 0.01 and #### <span class="html-italic">p</span> &lt; 0.0001 was considered a significant result compared to BSA control. (<b>C</b>) Representative western blot results showed PA-induced expression of lipogenic markers. AML12 cells were treated with PA (0.2, 0.4, and 0.8 mM) for 24 h, and the expression of lipogenic markers was dose-dependent. (<b>D</b>) The quantification of the target protein was calculated with a densitometer. The values are expressed as the fold of change (X basal), in Mean ± SEM, where <span class="html-italic">n</span> = 3. *** <span class="html-italic">p</span> &lt; 0.01 and #### <span class="html-italic">p</span> &lt; 0.0001 compared to the control. (<b>E</b>) Representative western blot results showing PA-induced expression of lipogenic markers. MIHA cells were treated with PA (0.2, 0.4, and 0.8 mM) for 24 h, and the expression of lipogenic markers was dose-dependent. (<b>F</b>) The quantification of the target protein was calculated with a densitometer. The values are expressed as the fold of change (X basal), in Mean ± SEM, where <span class="html-italic">n</span> = 3. *** <span class="html-italic">p</span> &lt; 0.01 and #### <span class="html-italic">p</span> &lt; 0.0001 compared to the control. Abbreviation: PGC1α, peroxisome proliferator-activated receptor gamma coactivator 1 alpha; PA, palmitic acid; GAPDH, Glyceraldehyde 3-phosphate dehydrogenase; ACC, acetyl-CoA carboxylase; FAS, fatty acid synthase; SCD1, stearoyl CoA desaturase-1; LPL, lipoprotein lipase; FABP-L, liver-type fatty acid-binding protein.</p>
Full article ">Figure 3
<p>(<b>A</b>) Representative western blot results showed PA-induced expression of inflammatory markers. AML12 cells were treated with PA (0.2, 0.4, 0.8 mM) for 24 h, and the expression of inflammatory markers was dose-dependent. (<b>B</b>) The quantification of the target protein was calculated with a densitometer. The values are expressed as the fold of change (X basal), in mean ± SEM, where <span class="html-italic">n</span> = 3. *** <span class="html-italic">p</span> &lt; 0.001 and #### <span class="html-italic">p</span> &lt; 0.0001 compared to the control. (<b>C</b>) Representative western blot results showed PA-induced expression of inflammatory markers. MIHA cells were treated with PA (0.2, 04, 0.8 mM) for 24 h, and the expression of inflammatory markers was dose-dependent. (<b>D</b>) The quantification of the target protein was calculated with a densitometer. The values are expressed as the fold of change (X basal), in Mean ± SEM, where <span class="html-italic">n</span> = 3. *** <span class="html-italic">p</span> &lt; 0.001 and #### <span class="html-italic">p</span> &lt; 0.0001 compared to the control. Abbreviation: PA, palmitic acid; NF-kB, nuclear factor-kappa B; COX-2, cyclooxygenase-2; TNF-α, tumor necrosis factor-alpha; IL-6, interleukin-6; GAPDH, Glyceraldehyde 3-phosphate dehydrogenase.</p>
Full article ">Figure 4
<p>CM derived from (<b>A</b>) AML12 cells and (<b>B</b>) MIHA cells promote the proliferation of BC. The respective BC cell lines were treated with conditional medium obtained from hepatic cells for 24 h, followed by the assay of MTT. The data are presented as means ± SD for three replicates. Data were analyzed with a student’s <span class="html-italic">t</span>-test. CM was prepared from hepatic cells that were previously treated with PA (0.8 mM) for 24 h, followed by 12 h treatment with fresh complete medium. In the next step, CM was filtered and diluted with 25% fresh medium. The values are expressed in mean ± SEM, where <span class="html-italic">n</span> = 3. (<b>C</b>,<b>D</b>) Treatment of rPGC1α promotes cell proliferation in BC (E0771 and MCF7) cell lines. The respective BC cell lines were treated with PGC1α (0–40 ng/mL) for 24 h, followed by an MTT assay. The data are presented as means ± SD for three replicates. Data were analyzed with a one-way ANOVA followed by group comparisons using a post hoc Tukey’s multiple comparison test. (<b>E</b>) Quantification of PGC1α in CM using ELISA. CM was prepared from hepatic cells that were previously treated with 0.8 mM PA for 6, 24, and 48 h, followed by a 12 h treatment with a fresh complete medium. Elevated levels of PGC1α were observed in the CM at all time points compared to those in the control, according to ELISA results. The values are given in mean SEM, with <span class="html-italic">n</span> = 3. (<b>F</b>) The results of the co-immunoprecipitation assay suggested an interaction between PGC1α and ERRα. (<b>G</b>) PGC1α promotes colony formation in BC cells. Clonogenic assay results for E0771 and MCF7 cells treated with rPGC1α at the indicated concentrations. The data are presented as means ± SD for three replicates. Data were analyzed with a one-way ANOVA followed by group comparisons using a post hoc Tukey’s multiple comparison test. (<b>H</b>) The clonogenic number was obtained from the treatment of rPGC1α with the indicated concentrations in E0771 and MCF7 cells. The colonies’ numbers were counted. The data are presented as means ± SD for three replicates. Data were analyzed with a one-way ANOVA followed by group comparisons using a post hoc Tukey’s multiple comparison test.</p>
Full article ">Figure 5
<p>(<b>A</b>) High levels of rPGC1α inhibit apoptosis-related factors. The respective BC (E0771 and MCF7) cell lines were treated with PGC1α (5 and 40 ng/mL) for 24 h followed by the assay of apoptosis markers using western blot. (<b>B</b>) The quantification of the target protein was calculated with a densitometer. The values are expressed as the fold of change (X basal), in mean ± SEM, where <span class="html-italic">n</span> = 3. #### <span class="html-italic">p</span> &lt; 0.0001 was considered a significant result when compared to the untreated rPGC1α group. (<b>C</b>) Protein expression of rPGC1α treated BC cell lines. Treatment of rPGC1α (0, 5, and 40 ng/mL) increased cell proliferation through the downstream expression of RAS-RAF-MAPK and PI3K-Akt-mTOR pathways on respective BC cell lines. (<b>D</b>) The quantification of the target protein was calculated with a densitometer. The values are expressed as the fold of change (X basal), in mean ± SEM, where <span class="html-italic">n</span> = 3. #### <span class="html-italic">p</span> &lt; 0.0001 was considered a significant result when compared to the untreated rPGC1α group. (<b>E</b>) Deletion of the PGC1 gene in BC cells by transfection using Lipofectamine RNAiMAX. Represents the two sets of PGC1α siRNA transfection and thereby the absence of PGC1α expression when compared to a blank and negative control, which were analyzed by western blot. (<b>F</b>) Deletion of the PGC1α gene in BC cells for 24 h decreased the growth rate, which was determined by the MTT assay. The data are presented as means ± SEM for three replicates. Data were analyzed with one-way ANOVA followed by group comparisons using a post hoc Tukey’s multiple comparison test. Abbreviation: Bcl2, B-cell lymphoma 2; BAX, Bcl-2 Associated X-protein; cyc, cytochrome; Casp, caspase; PARP1, Poly [ADP-ribose] polymerase 1; PI3K, Phosphoinositide 3-kinases; Akt, serine/threonine-protein kinase; mTOR, mammalian target of rapamycin; RAS, rat sarcoma; ERK1/2, extracellular signal-regulated kinase 2; GAPDH, Glyceraldehyde 3-phosphate dehydrogenase.</p>
Full article ">Figure 6
<p>(<b>A</b>) Silencing of PGC1α inhibits BC cell proliferation. Scrambled or siPGC1α of respective BC cells treated with the medium for 24 h followed by the assay of proliferation markers (Ki67 and PCNA) using a western blot. (<b>B</b>) The quantification of the target protein was calculated with a densitometer. The values are expressed as the fold of change (X basal), in mean ± SEM, where <span class="html-italic">n</span> = 3. #### <span class="html-italic">p</span> &lt; 0.0001 was considered a significant result when compared to the scrambled group. (<b>C</b>) Silencing of PGC1α promotes BC cell apoptosis-related factors. Scrambled or siPGC1α of respective BC cells treated with the medium for 24 h, followed by the assay of apoptosis markers using a western blot. (<b>D</b>) The quantification of the apoptosis target protein was calculated with a densitometer. The values are expressed as the fold of change (X basal), in mean ± SEM, where <span class="html-italic">n</span> = 3. *** <span class="html-italic">p</span> &lt; 0.01 was considered a significant result when compared to the control, and #### <span class="html-italic">p</span> &lt; 0.001 was considered a significant result when compared to the scrambled group. (<b>E</b>) siPGC1α inhibits colony formation in BC cells. Clonogenic assay results for E0771 and MCF7 cells treated with scrambled or siPGC1α. (<b>F</b>) Quantification of colonies treated with scrambled or siPGC1α and colonies counted relative to scrambled. The data are presented as means ± SD for three replicates. Data were analyzed with a one-way ANOVA followed by group comparisons using a post hoc Tukey’s multiple comparison test. Abbreviation: PCNA, Proliferating cell nuclear antigen; Bcl2, B-cell lymphoma 2; BAX, Bcl-2 Associated X-protein; cyc, cytochrome; Casp, caspase; PARP1, Poly [ADP-ribose] polymerase 1; GAPDH, Glyceraldehyde 3-phosphate dehydrogenase.</p>
Full article ">Figure 7
<p>(<b>A</b>) Expression of ERRα in the rPGC1α-treated BC cells by Western blot analysis. The respective BC cells were injected with rPGC1α (0, 5, 40 ng/mL) for 24 h and the expression of ERRα was dose-dependent. (<b>B</b>) The quantification of the target protein was calculated with a densitometer. The values are expressed as the fold of change (X basal), in mean ± SEM, where <span class="html-italic">n</span> = 3. ### <span class="html-italic">p</span> &lt; 0.01 and #### <span class="html-italic">p</span> &lt; 0.0001 was considered a significant result when compared to the untreated rPGC1α group. (<b>C</b>) ERRα antagonist (XCT-790) inhibits PGC1α-induced BC proliferation. The respective BC cells were treated with XCT-790 (5 μM) and rPGC1α (0, 5, 40 ng/mL) for 24 h, which reduced the expression of ERRα cell proliferation markers through the downstream expression of RAS-RAF-MAPK and PI3K-Akt-mTOR pathways on respective BC cell lines. (<b>D</b>) The quantification of the target protein was calculated with a densitometer. The values are expressed as the fold of change (X basal), in mean ± SEM, where <span class="html-italic">n</span> = 3. #### <span class="html-italic">p</span> &lt; 0.0001 was considered a significant result when compared to the untreated rPGC1α group. Abbreviation: ERRα, estrogen-related receptor alpha; rPGC1α, recombinant peroxisome proliferator-activated receptor gamma coactivator 1 alpha; PI3K, Phosphoinositide 3-kinases; Akt, serine/threonine-protein kinase; mTOR, mammalian target of rapamycin; RAS, rat sarcoma; ERK1/2, extracellular signal-regulated kinase 2; GAPDH, Glyceraldehyde 3-phosphate dehydrogenase.</p>
Full article ">Figure 8
<p>(<b>A</b>) A high level of PGC1α promotes migration. The effect of PGC1α on cell motility was analyzed by a wound-healing assay. (<b>B</b>) The data were analyzed with a one-way ANOVA followed by group comparisons using a post hoc Tukey’s multiple comparison test. (<b>C</b>) A high level of PGC1α promotes migration. The effect of rPGC1α on cell motility was examined by a transwell assay. The data are presented as means ± SD for three replicates. Data were analyzed with a one-way ANOVA followed by group comparisons using a post hoc Tukey’s multiple comparison test. (<b>D</b>) A high level of PGC1α promotes invasion. The effect of rPGC1α on cell motility was examined by a matrigel-coated transwell assay. The data are presented as means ± SD for three replicates. Data were analyzed with a one-way ANOVA followed by group comparisons using a post hoc Tukey’s multiple comparison test. (<b>E</b>) A high level of PGC1α promotes the EMT process. The markers of EMT were determined using high levels of PGC1α on respective breast cancer cells (EO771 and T47D), which were measured by western blot. (<b>F</b>) The quantification of the target protein was calculated with a densitometer. The values are expressed as the fold of change (X basal), in mean ± SEM, where <span class="html-italic">n</span> = 3. #### <span class="html-italic">p</span> &lt; 0.0001 was considered a significant result when compared to the untreated rPGC1α group. Abbreviation: rPGC1α, recombinant peroxisome proliferator-activated receptor gamma coactivator 1 alpha; E-cad; E-cadherin; N-cad, N-cadherin; α-SMA, alpha-smooth muscle actin; MMP, matrix metalloproteinases; GAPDH, Glyceraldehyde 3-phosphate dehydrogenase.</p>
Full article ">Figure 9
<p>(<b>A</b>) Silencing of PGC1α suppresses migration. The effect of scrambled or siPGC1α on cell motility was analyzed by wound healing assay. (<b>B</b>) The data were analyzed with a one-way ANOVA followed by group comparisons using a post hoc Tukey’s multiple comparison test. (<b>C</b>) Silencing of PGC1α suppresses migration. The effect of PGC1α on cell motility was examined by transwell assay. The data are presented as means ± SD for three replicates. Data were analyzed with a one-way ANOVA followed by group comparisons using a post hoc Tukey’s multiple comparison test. (<b>D</b>) The silencing of PGC1α suppresses invasion. The effect of PGC1α on cell motility was assessed by a matrigel-coated transwell assay. The data are presented as means ± SD for three replicates. Data were analyzed with one-way ANOVA followed by group comparisons using a post hoc Tukey’s multiple comparison test. (<b>E</b>) Silencing of PGC1α suppresses the EMT process. As a result of treatment with the medium for 24 h, scrambled or siPGC1α cells of BC cells (EO771 and T47D) were analyzed by western blotting. (<b>F</b>) The quantification of the target protein was calculated with a densitometer. The values are expressed as the fold of change (X basal), in mean ± SEM, where <span class="html-italic">n</span> = 3. **** <span class="html-italic">p</span> &lt; 0.0001 was considered a significant result when compared to blank, and #### <span class="html-italic">p &lt;</span> 0.0001 was considered a significant result when compared to scrambled. Abbreviation: rPGC1α, recombinant peroxisome proliferator-activated receptor gamma coactivator 1 alpha; E-cad; E-cadherin; N-cad, N-cadherin; α-SMA, alpha-smooth muscle actin; MMP, matrix metalloproteinases; GAPDH, Glyceraldehyde 3-phosphate dehydrogenase.</p>
Full article ">
26 pages, 8216 KiB  
Article
Syringin Prevents 6-Hydroxydopamine Neurotoxicity by Mediating the MiR-34a/SIRT1/Beclin-1 Pathway and Activating Autophagy in SH-SY5Y Cells and the Caenorhabditis elegans Model
by Ru-Huei Fu, Syuan-Yu Hong and Hui-Jye Chen
Cells 2023, 12(18), 2310; https://doi.org/10.3390/cells12182310 - 19 Sep 2023
Cited by 1 | Viewed by 1625
Abstract
Defective autophagy is one of the cellular hallmarks of Parkinson’s disease (PD). Therefore, a therapeutic strategy could be a modest enhancement of autophagic activity in dopamine (DA) neurons to deal with the clearance of damaged mitochondria and abnormal protein aggregates. Syringin (SRG) is [...] Read more.
Defective autophagy is one of the cellular hallmarks of Parkinson’s disease (PD). Therefore, a therapeutic strategy could be a modest enhancement of autophagic activity in dopamine (DA) neurons to deal with the clearance of damaged mitochondria and abnormal protein aggregates. Syringin (SRG) is a phenolic glycoside derived from the root of Acanthopanax senticosus. It has antioxidant, anti-apoptotic, and anti-inflammatory properties. However, whether it has a preventive effect on PD remains unclear. The present study found that SRG reversed the increase in intracellular ROS-caused apoptosis in SH-SY5Y cells induced by neurotoxin 6-OHDA exposure. Likewise, in C. elegans, degeneration of DA neurons, DA-related food-sensitive behaviors, longevity, and accumulation of α-synuclein were also improved. Studies of neuroprotective mechanisms have shown that SRG can reverse the suppressed expression of SIRT1, Beclin-1, and other autophagy markers in 6-OHDA-exposed cells. Thus, these enhanced the formation of autophagic vacuoles and autophagy activity. This protective effect can be blocked by pretreatment with wortmannin (an autophagosome formation blocker) and bafilomycin A1 (an autophagosome–lysosome fusion blocker). In addition, 6-OHDA increases the acetylation of Beclin-1, leading to its inactivation. SRG can induce the expression of SIRT1 and promote the deacetylation of Beclin-1. Finally, we found that SRG reduced the 6-OHDA-induced expression of miR-34a targeting SIRT1. The overexpression of miR-34a mimic abolishes the neuroprotective ability of SRG. In conclusion, SRG induces autophagy via partially regulating the miR-34a/SIRT1/Beclin-1 axis to prevent 6-OHDA-induced apoptosis and α-synuclein accumulation. SRG has the opportunity to be established as a candidate agent for the prevention and cure of PD. Full article
(This article belongs to the Special Issue Autophagy in Parkinson's Disease)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The molecular structure diagram of syringin (SRG) from <span class="html-italic">Acanthopanax senticosus</span>.</p>
Full article ">Figure 2
<p>SRG prevents the apoptosis of SH-SY5Y cells damaged via 6-OHDA. (<b>A</b>) SH-SY5Y cells were treated with serially diluted SRG. The percentage of viable cells after 24 h was determined using the CellTiter Blue cell viability assay. (<b>B</b>) Cell viability was analyzed following exposure of (<b>A</b>)-treated cells to 100 μM 6-OHDA for 18 h. (<b>C</b>) DiOC6 staining was used to analyze changes in the mitochondrial membrane potential (MMP) of cells treated with (<b>B</b>) under a fluorescence microscope (scale bar = 50 µm). The fluorescence intensity of the image was estimated via ImageJ software (version 1.53). (<b>D</b>) The proportion of cells with apoptosis-related DNA fragmentation was determined via fluorescence microscopy (scale bar = 100 µm) using TUNEL staining. (<b>E</b>) Cells treated with (<b>B</b>) were analyzed via flow cytometry to determine the number of apoptotic cell populations presented via Annexin V-FITC binding and propidium iodide (PI) staining. (<b>F</b>) Changes in levels of apoptosis-related proteins in cells treated with (<b>B</b>) were analyzed via Western blotting. β-tubulin was used to normalize the level of total protein in each group.</p>
Full article ">Figure 2 Cont.
<p>SRG prevents the apoptosis of SH-SY5Y cells damaged via 6-OHDA. (<b>A</b>) SH-SY5Y cells were treated with serially diluted SRG. The percentage of viable cells after 24 h was determined using the CellTiter Blue cell viability assay. (<b>B</b>) Cell viability was analyzed following exposure of (<b>A</b>)-treated cells to 100 μM 6-OHDA for 18 h. (<b>C</b>) DiOC6 staining was used to analyze changes in the mitochondrial membrane potential (MMP) of cells treated with (<b>B</b>) under a fluorescence microscope (scale bar = 50 µm). The fluorescence intensity of the image was estimated via ImageJ software (version 1.53). (<b>D</b>) The proportion of cells with apoptosis-related DNA fragmentation was determined via fluorescence microscopy (scale bar = 100 µm) using TUNEL staining. (<b>E</b>) Cells treated with (<b>B</b>) were analyzed via flow cytometry to determine the number of apoptotic cell populations presented via Annexin V-FITC binding and propidium iodide (PI) staining. (<b>F</b>) Changes in levels of apoptosis-related proteins in cells treated with (<b>B</b>) were analyzed via Western blotting. β-tubulin was used to normalize the level of total protein in each group.</p>
Full article ">Figure 3
<p>SRG pretreatment decreased the production of cellular ROS and increased the autophagy in the 6-OHDA-exposed SH-SY5Y cell model. After 24 h of SRG pretreatment, cells were exposed to 100 μM 6-OHDA for an additional 18 h. (<b>A</b>) The ROS production in cells was analyzed via H2DCFDA probe and spectrophotometer. (<b>B</b>) The acidic vacuoles level in the cells was stained with acridine orange and then observed using a fluorescence microscope (scale bar = 50 µm). Fluorescence intensity was quantified via ImageJ software. (<b>C</b>) Autophagy activity in the cells was analyzed via LC3-II staining and spectrophotometer. Signal intensity was quantified via ImageJ. (<b>D)</b> After SRG pretreatment for 24 h, SH-SY5Y cells were exposed to 100 μM 6-OHDA for 12 h, and finally, the expression of autophagy-related genes was analyzed via Western blotting. The histogram is the result of quantitative analysis of the measured signal using ImageJ. β-tubulin was used to normalize the level of total protein in each group.</p>
Full article ">Figure 3 Cont.
<p>SRG pretreatment decreased the production of cellular ROS and increased the autophagy in the 6-OHDA-exposed SH-SY5Y cell model. After 24 h of SRG pretreatment, cells were exposed to 100 μM 6-OHDA for an additional 18 h. (<b>A</b>) The ROS production in cells was analyzed via H2DCFDA probe and spectrophotometer. (<b>B</b>) The acidic vacuoles level in the cells was stained with acridine orange and then observed using a fluorescence microscope (scale bar = 50 µm). Fluorescence intensity was quantified via ImageJ software. (<b>C</b>) Autophagy activity in the cells was analyzed via LC3-II staining and spectrophotometer. Signal intensity was quantified via ImageJ. (<b>D)</b> After SRG pretreatment for 24 h, SH-SY5Y cells were exposed to 100 μM 6-OHDA for 12 h, and finally, the expression of autophagy-related genes was analyzed via Western blotting. The histogram is the result of quantitative analysis of the measured signal using ImageJ. β-tubulin was used to normalize the level of total protein in each group.</p>
Full article ">Figure 4
<p>Treatment with wortmannin and bafilomycin A1 abolished the character of SRG to reverse the protein expression associated with the down-regulation of autophagy and enhancement of apoptosis caused by 6-OHDA exposure. SH-SY5Y cells were pretreated with SRG for 24 h after adding 0.5 μM wortmannin (wort) or 0.5 nM bafilomycin A1 (baf-1) for 1 h and then incubated with 6-OHDA for an additional 12 h. (<b>A</b>) The protein level of PI3 kinase p100, Beclin-1, Atg7, and the ratio of LC3-II/LC3-I were measured via Western blotting. (<b>B</b>) The level of apoptosis-related proteins was detected via Western blotting. β-tubulin was used as a loading control. Alterations in protein levels were determined via ImageJ software.</p>
Full article ">Figure 5
<p>The 6-OHDA caused the downregulation of the SIRT1 protein expression, and the acetylation of Beclin-1 was reversed via SRG pretreatment in SH-SY5Y cells. (<b>A</b>) The expression of the SIRT1 protein in SH-SY5Y cells was analyzed via Western blotting. β-tubulin was used as a loading control. The protein expression was estimated via ImageJ software. (<b>B</b>) The lysate of SH-SY5Y cells was immunoprecipitated with a Beclin-1 antibody and then analyzed via Western blotting using an anti-acetylated lysine residue antibody. In addition, β-tubulin was used as the lysate input control. Consistent Beclin-1 expression in the lysate of each sample was used as the benchmark for determining the degree of acetylation.</p>
Full article ">Figure 6
<p>SRG reduces the upregulation of miR-34a expression induced via 6-OHDA exposure to restore SIRT1 expression in SH-SY5Y cells. (<b>A</b>) RT-qPCR analysis of <span class="html-italic">SIRT1</span> mRNA expression in SH-SY5Y cells of each group. (<b>B</b>) RT-qPCR analysis of miR-34a levels in SH-SY5Y cells of each group. (<b>C</b>,<b>D</b>) The role of miR-34a in the SRG-enhanced SIRT1 pathway was confirmed using (<b>C</b>) anti-miR-34a (miR-34a inhibitor) and (<b>D</b>) miR-34a mimics. The level of the SIRT1 protein in SH-SY5Y cells of each group was analyzed via Western blotting. β-tubulin was used as an internal loading control, and the signal was quantified via ImageJ software.</p>
Full article ">Figure 7
<p>The 6-OHDA neurotoxicity-induced deficits in DA neuron integrity, food-sensitive behavior, and lifespan in <span class="html-italic">C. elegans</span> were ameliorated via SRG pretreatment. (<b>A</b>) A food clearance test was used to obtain an appropriate dose of SRG for treating worms. The L1-stage worms (approximately 20) of the used strains were placed in 96-well plates containing <span class="html-italic">Escherichia coli</span> OP50 (OD A<sub>595</sub> = 0.6) and different concentrations of SRG in an S medium for 6 days (20 °C). The OD value of each well was recorded every day. (<b>B</b>) Fluorescent images of GFP-labeled DA neurons in the head of BZ555 worms from each group. Fluorescence signals were calculated using ImageJ software. Scale bar = 50 µm. (<b>C</b>) The phenotype deficit of DA neurons in (<b>B</b>) was visually assessed and expressed as a percentage. If the axonal fluorescence signal was punctated or disappeared, it was considered degeneration. (<b>D</b>) The effect of SRG on the restoration of the DA neuronal function in 6-OHDA-exposed N2 worms was evaluated in a food-sensitivity behavioral assay. The slowing rate is expressed as the percentage reduction in the frequency of the S-shaped movement (20 s) of the worms moving from the empty lawn to the bacterial lawn. Fifty worms per group were counted. (<b>E</b>) The lifespan of N2 worms was assessed via cumulative survival curves. The worms in each group were replaced with fresh plates every three days, and the number of surviving worms was calculated every day. Fifty worms per group were counted.</p>
Full article ">Figure 7 Cont.
<p>The 6-OHDA neurotoxicity-induced deficits in DA neuron integrity, food-sensitive behavior, and lifespan in <span class="html-italic">C. elegans</span> were ameliorated via SRG pretreatment. (<b>A</b>) A food clearance test was used to obtain an appropriate dose of SRG for treating worms. The L1-stage worms (approximately 20) of the used strains were placed in 96-well plates containing <span class="html-italic">Escherichia coli</span> OP50 (OD A<sub>595</sub> = 0.6) and different concentrations of SRG in an S medium for 6 days (20 °C). The OD value of each well was recorded every day. (<b>B</b>) Fluorescent images of GFP-labeled DA neurons in the head of BZ555 worms from each group. Fluorescence signals were calculated using ImageJ software. Scale bar = 50 µm. (<b>C</b>) The phenotype deficit of DA neurons in (<b>B</b>) was visually assessed and expressed as a percentage. If the axonal fluorescence signal was punctated or disappeared, it was considered degeneration. (<b>D</b>) The effect of SRG on the restoration of the DA neuronal function in 6-OHDA-exposed N2 worms was evaluated in a food-sensitivity behavioral assay. The slowing rate is expressed as the percentage reduction in the frequency of the S-shaped movement (20 s) of the worms moving from the empty lawn to the bacterial lawn. Fifty worms per group were counted. (<b>E</b>) The lifespan of N2 worms was assessed via cumulative survival curves. The worms in each group were replaced with fresh plates every three days, and the number of surviving worms was calculated every day. Fifty worms per group were counted.</p>
Full article ">Figure 8
<p>SRG-induced autophagic activity relieves α-synuclein accumulation in muscle cells of worms. (<b>A</b>) Muscle cells of NL5901 worms expressing YFP-fused human α-synuclein were used to assess α-synuclein accumulation. Fifty worms were tested in each group. The fluorescence signals of YFP were calculated via ImageJ software. (scale bar = 100 μm) (<b>B</b>) The expression of YFP-fused human α-synuclein from (<b>A</b>) was analyzed via Western blotting. The signal intensity was quantified via ImageJ software. (<b>C</b>) DA2123 worms expressing the LGG-1 promoter-inducible GFP-fused LGG-1 protein in hypodermal seam cells were used to assess autophagic activity in worms. Fluorescent punctate signals represent the formation of autophagosomes (autophagy activation) and are counted from at least 10 seam cells per worm. Fifty worms were tested in each group. (scale bar = 200 μm).</p>
Full article ">Figure 9
<p>SRG pretreatment inhibited 6-OHDA-caused ROS upregulation, and promoted the <span class="html-italic">sir-2.1</span> pathway in N2 worms. (<b>A</b>) Thirty worms from each group were placed in the wells of a 96-well plate to detect the total ROS level using an H2DCFDA probe and a spectrophotometer. (<b>B</b>) Quantitative detection of sir-2.1 expression in worms treated with SRG via RT-qPCR. (<b>C</b>) The effect of SRG pretreatment on sir-2.1 expression in N2 worms exposed to 6-OHDA was quantified via RT-qPCR. (<b>D</b>) The VC199 strain (the mutant of sir2.1) was used to assess the effect of sir-2.1 on SRG by improving food sensitivity behaviors impaired via 6-OHDA in worms.</p>
Full article ">
15 pages, 3303 KiB  
Article
Lymphatic Defects in Zebrafish sox18 Mutants Are Exacerbated by Perturbed VEGFC Signaling, While Masked by Elevated sox7 Expression
by Silvia Moleri, Sara Mercurio, Alex Pezzotta, Donatella D’Angelo, Alessia Brix, Alice Plebani, Giulia Lini, Marialaura Di Fuorti and Monica Beltrame
Cells 2023, 12(18), 2309; https://doi.org/10.3390/cells12182309 - 19 Sep 2023
Cited by 2 | Viewed by 1453
Abstract
Mutations in the transcription factor-coding gene SOX18, the growth factor-coding gene VEGFC and its receptor-coding gene VEGFR3/FLT4 cause primary lymphedema in humans. In mammals, SOX18, together with COUP-TFII/NR2F2, activates the expression of Prox1, a master regulator in lymphatic identity and development. [...] Read more.
Mutations in the transcription factor-coding gene SOX18, the growth factor-coding gene VEGFC and its receptor-coding gene VEGFR3/FLT4 cause primary lymphedema in humans. In mammals, SOX18, together with COUP-TFII/NR2F2, activates the expression of Prox1, a master regulator in lymphatic identity and development. Knockdown studies have also suggested an involvement of Sox18, Coup-tfII/Nr2f2, and Prox1 in zebrafish lymphatic development. Mutants in the corresponding genes initially failed to recapitulate the lymphatic defects observed in morphants. In this paper, we describe a novel zebrafish sox18 mutant allele, sa12315, which behaves as a null. The formation of the lymphatic thoracic duct is affected in sox18 homozygous mutants, but defects are milder in both zygotic and maternal-zygotic sox18 mutants than in sox18 morphants. Remarkably, in sox18 mutants, the expression of the closely related sox7 gene is elevated where lymphatic precursors arise. Sox7 could thus mask the absence of a functional Sox18 protein and account for the mild lymphatic phenotype in sox18 mutants, as shown in mice. Partial knockdown of vegfc exacerbates lymphatic defects in sox18 mutants, making them visible in heterozygotes. Our data thus reinforce the genetic interaction between Sox18 and Vegfc in lymphatic development, previously suggested by knockdown studies, and highlight the ability of Sox7 to compensate for Sox18 lymphatic dysfunction. Full article
(This article belongs to the Special Issue Modeling Developmental Processes and Disorders in Zebrafish)
Show Figures

Figure 1

Figure 1
<p>The <span class="html-italic">sox18<sup>sa12315</sup></span> mutant behaves as expected for a null allele. (<b>A</b>) On the left is a schematic representation of the Sox18 protein, with the HMG-box domain in blue. The G &gt; A transition and the premature stop codon introduced in the <span class="html-italic">sa12315</span> mutant are indicated. Fragments of the electropherograms derived using Sanger sequencing of the region surrounding the mutation in wt (<span class="html-italic">sox18<sup>+/+</sup></span>), heterozygous (<span class="html-italic">sox18<sup>+/−</sup></span>) or homozygous mutants (<span class="html-italic">sox18<sup>−/−</sup></span>) are reported on the right. The restriction site for the BstNI/MvaI enzymes (boxed sequence) is disrupted by the mutation. (<b>B</b>) Embryos derived from <span class="html-italic">sa12315</span> heterozygote matings and injected with subcritical doses of <span class="html-italic">sox7</span>-MO were collected in several independent experiments and analyzed in vivo at 2 dpf and 3 dpf or fixed at around 30 hpf for ISH, as shown in (<b>C</b>). The histogram on the right shows the trunk–tail circulatory phenotypes observed at 3 dpf. In control embryos, i.e., uninjected or injected with a standard control MO (first and second bars, respectively), trunk–tail circulatory defects are present in a small percentage of embryos. On the contrary, the partial knockdown of <span class="html-italic">sox7</span> causes a blockage in trunk–tail circulation in a dose-dependent manner (third and fourth bars). Circulatory defects are genotype-dependent (see <a href="#app1-cells-12-02309" class="html-app">Table S1</a>). (<b>C</b>) ISHs were performed on embryos derived from <span class="html-italic">sa12315</span> heterozygote matings and injected with subcritical doses of <span class="html-italic">sox7</span>-MO, as shown in (<b>B</b>), fixed at around 30 hpf. Upper panels show control ISH performed with the endothelial marker <span class="html-italic">cdh5,</span> showing no gross alteration in embryos of the three different genotypes. Lower panels show ISHs performed with a probe for <span class="html-italic">vsg1/plvapb</span>, whose expression was particularly downregulated in double partial <span class="html-italic">sox7/sox18</span> morphants [<a href="#B29-cells-12-02309" class="html-bibr">29</a>]. Higher magnification images of the trunk–tail regions of the embryos are also shown. Experiments were repeated twice; all <span class="html-italic">plavpb</span> stained embryos and a subset of <span class="html-italic">cdh5</span> stained embryos were genotyped; numbers in each image refer to a single experiment. Lateral views, anterior to the left. Pictures were taken at 40× and 63× magnification, for lower and higher magnification images respectively.</p>
Full article ">Figure 2
<p>Homozygous <span class="html-italic">sox18</span> mutants show subtle but statistically significant defects in thoracic duct (TD) formation. (<b>A</b>) Confocal trunk images representing wt (+/+) and <span class="html-italic">sox18<sup>sa12315</sup></span> homozygous mutant larvae (−/−) in the <span class="html-italic">Tg(lyve1b:DsRed)</span> line at 5dpf. Large and small white arrowheads point to TD+ segments (of typical or thinner aspect, respectively) while asterisks indicate the absence of TD. (<b>B</b>) The graph reports the mean number of TD+ segments, counted along 10 consecutive trunk segments, together with the Standard Error of the Mean (SEM), in all analyzed embryos of the three genotypes (wt: <span class="html-italic">sox18</span><sup>+/+</sup>, het: <span class="html-italic">sox18</span><sup>+/−</sup>, hom: <span class="html-italic">sox18</span><sup>−/−</sup>). Data were gathered in three independent experiments, and each symbol represents the number of TD+ segments of a single analyzed larva. n = number of larvae, ** = <span class="html-italic">p</span> &lt; 0.01. TD = thoracic duct; DA = dorsal aorta; PCV = posterior cardinal vein. Lateral view, anterior to the left.</p>
Full article ">Figure 3
<p>TD formation defects are exacerbated upon slight perturbation of Vegfc signaling. The progeny of <span class="html-italic">sox18<sup>sa12315</sup></span> heterozygote matings in the <span class="html-italic">Tg(fli1a:EGFP)<sup>y1</sup></span> line were injected with a subcritical dose of <span class="html-italic">vegfc</span>-MO or left uninjected; TD formation was analyzed at 5dpf. (<b>A</b>) Confocal trunk images of uninjected wt (+/+) and <span class="html-italic">sox18</span> homozygous mutant (−/−) larvae. Arrowheads point to TD+ segments, while asterisks indicate the absence of TD; a smaller arrowhead marks a thinner TD+ segment. (<b>B</b>) The graph reports the mean number of TD+ segments, counted along 10 consecutive trunk segments, together with the SEM, for all analyzed larvae of each genotype (wt: sox18<sup>+/+</sup>, het: sox18<sup>+/−</sup>, hom: sox18<sup>−/−</sup>). Each symbol represents the number of TD+ segments of a single larva; data were gathered in several independent experiments. Uninjected larvae on the left are compared to larvae with partially reduced Vegfc on the right. n = number of larvae, * = <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. TD = thoracic duct; DA = dorsal aorta; PCV = posterior cardinal vein. Lateral view, anterior to the left.</p>
Full article ">Figure 4
<p>The expression of <span class="html-italic">sox7</span> in the PCV is upregulated in <span class="html-italic">sox18</span> mutants, but not in <span class="html-italic">sox18</span> morphants. (<b>A</b>) Representative images of <span class="html-italic">sox7</span> ISH on embryos at around 26 hpf derived from matings of <span class="html-italic">sox18<sup>sa12315</sup></span> heterozygotes in the <span class="html-italic">Tg(lyve1b:DsRed)</span> line. Higher magnifications of the trunk region are shown below the full size images. Compared to wt embryos, the <span class="html-italic">sox7</span> ISH signal in the PCV is elevated in the great majority of <span class="html-italic">sox18</span><sup>−/−</sup> homozygotes and, to a lesser extent, in <span class="html-italic">sox18</span><sup>+/−</sup> heterozygotes. Numbers in each image state the number of embryos with the reported phenotype over the total analyzed embryos in one representative experiment. Lateral views, anterior to the left. Pictures were taken at 40× and 63× magnification, for lower and higher magnification images respectively. (<b>B</b>) Left, representative ImageJ-modified images, used to perform the quantification of the <span class="html-italic">sox7</span> ISH signal in the PCV and the DA (as described in Materials and Methods) on 26 hpf wt (+/+), <span class="html-italic">sox18<sup>sa12315</sup></span> heterozygotes (+/−) or homozygous mutants (−/−). The graph on the right shows the calculated PCV/DA ratio in each embryo; embryos are grouped based on their genotypes: mean values and SEM are indicated. (<b>C</b>) The same analysis was performed on <span class="html-italic">sox7</span> ISH of <span class="html-italic">sox18</span> morphants and control embryos, as shown in <a href="#app1-cells-12-02309" class="html-app">Figure S6</a>. The calculated PCV/DA ratio of each embryo is shown in the graph; mean values and SEM for std-MO injected embryos and <span class="html-italic">sox18</span> morphants are indicated. n = number of embryos, ** = <span class="html-italic">p</span> &lt; 0.01; DA = dorsal aorta; PCV = posterior cardinal vein. The analysis was also repeated on ISHs of <span class="html-italic">sox18<sup>sa12315</sup></span> mutants in the <span class="html-italic">Tg(fli1a:EGFP)<sup>y1</sup></span> reporter line with similar results. ISH experiments on <span class="html-italic">sox18<sup>sa12315</sup></span> mutants were repeated at least three times. Data shown in A and B were generated on different clutches of embryos.</p>
Full article ">
24 pages, 2343 KiB  
Article
Altered Epigenetic Marks and Gene Expression in Fetal Brain, and Postnatal Behavioural Disorders, Following Prenatal Exposure of Ogg1 Knockout Mice to Saline or Ethanol
by Shama Bhatia, David Bodenstein, Ashley P. Cheng and Peter G. Wells
Cells 2023, 12(18), 2308; https://doi.org/10.3390/cells12182308 - 19 Sep 2023
Viewed by 1373
Abstract
Oxoguanine glycosylase 1 (OGG1) is widely known to repair the reactive oxygen species (ROS)-initiated DNA lesion 8-oxoguanine (8-oxoG), and more recently was shown to act as an epigenetic modifier. We have previously shown that saline-exposed Ogg1 −/− knockout progeny exhibited learning and memory [...] Read more.
Oxoguanine glycosylase 1 (OGG1) is widely known to repair the reactive oxygen species (ROS)-initiated DNA lesion 8-oxoguanine (8-oxoG), and more recently was shown to act as an epigenetic modifier. We have previously shown that saline-exposed Ogg1 −/− knockout progeny exhibited learning and memory deficits, which were enhanced by in utero exposure to a single low dose of ethanol (EtOH) in both Ogg1 +/+ and −/− progeny, but more so in Ogg1 −/− progeny. Herein, OGG1-deficient progeny exposed in utero to a single low dose of EtOH or its saline vehicle exhibited OGG1- and/or EtOH-dependent alterations in global histone methylation and acetylation, DNA methylation and gene expression (Tet1 (Tet Methylcytosine Dioxygenase 1), Nlgn3 (Neuroligin 3), Hdac2 (Histone Deacetylase 2), Reln (Reelin) and Esr1 (Estrogen Receptor 1)) in fetal brains, and behavioural changes in open field activity, social interaction and ultrasonic vocalization, but not prepulse inhibition. OGG1- and EtOH-dependent changes in Esr1 and Esr2 mRNA and protein levels were sex-dependent, as was the association of Esr1 gene expression with gene activation mark histone H3 lysine 4 trimethylation (H3K4me3) and gene repression mark histone H3 lysine 27 trimethylation (H3K27me3) measured via ChIP-qPCR. The OGG1-dependent changes in global epigenetic marks and gene/protein expression in fetal brains, and postnatal behavioural changes, observed in both saline- and EtOH-exposed progeny, suggest the involvement of epigenetic mechanisms in developmental disorders mediated by 8-oxoG and/or OGG1. Epigenetic effects of OGG1 may be involved in ESR1-mediated gene regulation, which may be altered by physiological and EtOH-enhanced levels of ROS formation, possibly contributing to sex-dependent developmental disorders observed in Ogg1 knockout mice. The OGG1- and EtOH-dependent associations provide a basis for more comprehensive mechanistic studies to determine the causal involvement of oxidative DNA damage and epigenetic changes in ROS-mediated neurodevelopmental disorders. Full article
Show Figures

Figure 1

Figure 1
<p><b>Dosing and behavioural testing timeline</b>. <span class="html-italic">Ogg1</span> heterozygous mice were mated, and on gestational day (<b>GD</b>) 17, the dams received a single dose of EtOH (2 g/kg i.p.) or its saline vehicle. The pups were delivered spontaneously, weaned after 3 weeks and subjected to a series of behavioural tests including interaction-induced ultrasonic vocalization (USV) at 3–4 weeks of age, open field activity at 6 weeks of age, social interaction with a novel mouse at 7–8 weeks of age, female-induced USV at 4–5 months of age and prepulse inhibition at approximately 5 months of age.</p>
Full article ">Figure 2
<p><b>Altered histone and DNA modifications in fetal brains of <span class="html-italic">Ogg1</span> −/− mice exposed in utero to EtOH.</b> GD 17 fetal brains exposed in utero to a single dose of EtOH (2 g/kg i.p.) or its saline vehicle were extracted 1, 6 and 24 h later from <span class="html-italic">Ogg1</span> +/+ and −/− littermates and were assessed for histone and DNA modifications. Fetal brains from at least three litters were used to minimize potential litter effects, and the number of fetal brains for each group is shown in parentheses. (<b>A</b>). The following histone modifications were analyzed: H3K9ac (activation mark), H3K9me3 and H3K27me3 (repressive marks). See <a href="#app1-cells-12-02308" class="html-app">Supplementary Figure S1</a> for 1 h results for H3K9ac and H3K9me3 (no differences observed) and for 6 and 24 h results for H3K4me3 (activation mark, no differences observed). (<b>B</b>). Increase in 5-methylcytosine (5-mC) levels in saline-exposed <span class="html-italic">Ogg1</span> −/− vs. +/+ fetal brains at 24 h, with a similar trend in EtOH-exposed fetal brains. The same DNA sample was used for both 5-mC and 5-hmC measurements. The significance of differences was determined by two-way ANOVA and a post hoc Tukey’s test. Abbreviations: 5-mC: 5-methylcytosine; 5-hmC: 5-hydroxymethylcytosine; H3K4me3: histone 3 lysine 9 trimethylation; H3K9ac: histone 3 lysine 9 acetylation; H3K9me3: histone 3 lysine 9 trimethylation; H3K27me3: histone 3 lysine 27 trimethylation.</p>
Full article ">Figure 3
<p><b>Altered gene expression in <span class="html-italic">Ogg1</span> −/− fetal brains exposed in utero to EtOH.</b> GD 17 fetal brains exposed in utero to a single dose of EtOH (2 g/kg i.p.) or its saline vehicle were extracted 6 and 24 h later from <span class="html-italic">Ogg1</span> +/+ and −/− littermates. Fetal brains from at least three litters were used to minimize potential litter effects, and the number of fetal brains for each group is shown in parentheses. Fetal brains were homogenized, RNA was extracted using the TRIzol method and mRNA expression levels were measured via RT-qPCR. <span class="html-italic">Gapdh</span> was used as a control. Expression levels of various learning and memory candidate genes were measured (see <a href="#app1-cells-12-02308" class="html-app">Supplementary Figure S2</a>). Above are the results for OGG1- and EtOH-dependent differences in mRNA levels in fetal brains. The significance of differences was determined by two-way ANOVA and a post hoc Tukey’s test. Abbreviations: <span class="html-italic">Hdac2:</span> histone deacetylase 2; <span class="html-italic">Nlgn3</span>: neuroligin 3; <span class="html-italic">Reln:</span> Reelin; <span class="html-italic">Tet1:</span> Ten-eleven translocation methylcytosine dioxygenase 1.</p>
Full article ">Figure 4
<p><b>OGG1- and sex-dependent differences in <span class="html-italic">Esr1</span> and <span class="html-italic">Esr2</span> mRNA levels and their ratios in fetal brains exposed in utero to EtOH</b>. GD 17 fetal brains exposed in utero to a single dose of EtOH (2 g/kg i.p.) or its saline vehicle were extracted 6 and 24 h later from <span class="html-italic">Ogg1</span> +/+ and −/− littermates. Fetal brains from at least three litters were used to minimize potential litter effects, and the number of fetal brains for each group is shown in parentheses. Fetal brains were homogenized, RNA was extracted using the TRIzol method and mRNA expression levels were measured via RT-qPCR. Gapdh was used as a control. Above are the results for <span class="html-italic">Esr1</span> and <span class="html-italic">Esr2</span> mRNA levels as well as their ratios in fetal brains. The significance of differences was determined by two-way ANOVA and a post hoc Tukey’s test. Abbreviations: <span class="html-italic">Esr1</span>: estrogen receptor 1; <span class="html-italic">Esr2</span>: estrogen receptor 2.</p>
Full article ">Figure 5
<p><b>OGG1- and sex-dependent differences in ESR1 and ESR2 protein levels and their ratios in fetal brains exposed in utero to EtOH</b>. GD 17 fetal brains exposed in utero to a single dose of EtOH (2 g/kg i.p.) or its saline vehicle were extracted 6 and 24 h later from <span class="html-italic">Ogg1</span> +/+ and −/− littermates. Fetal brains from at least three litters were used to minimize potential litter effects, and the number of fetal brains for each group is shown in parentheses. Fetal brains were homogenized, and protein levels were quantified via western blot. GAPDH was used as a loading control. Above are the results for ESR1 and ESR2 proteins levels as well as their ratios in fetal brains. The significance of differences for each sex was determined by two-way ANOVA and a post hoc Tukey’s test. Abbreviations: ESR1: estrogen receptor 1; ESR2: estrogen receptor 2.</p>
Full article ">Figure 6
<p><b>EtOH-mediated increased association of H3K27me3 with <span class="html-italic">Esr1</span> gene expression in <span class="html-italic">Ogg1</span> +/+ but not −/− fetal brains.</b> GD 17 fetal brains exposed in utero to a single dose of EtOH (2 g/kg i.p.) or its saline vehicle were extracted 6 h later from <span class="html-italic">Ogg1</span> +/+ and −/− littermates. Fetal brains from at least three litters were used to minimize potential litter effects, and the number of fetal brains for each group is shown in parentheses. (<b>A</b>). ChIP was performed using fetal brains, and extracted DNA was used to perform quantitative PCR using five different sets of primers directed against various regions of the <span class="html-italic">Esr1</span> gene. The <span class="html-italic">Esr1</span> gene has two promoter regions, marked with arrows, which can generate four transcript variants via gene splicing. The locations of the exons and introns on chromosome 10 are marked. The primer locations are relative to the first transcription start site (<b>TSS</b>). Regions 1–5 were chosen based on the Ensembl database reporting their association with activation or repressive epigenetic marks. Each of the regions was amplified after chromatin was immunoprecipitated using antibodies against H3K4me3 (active promoter) and H3K27me3 (inactive promoter) and histone H3 (control). This figure is not drawn to scale. (<b>B</b>). The association of the H3K27me3:H3 ratio was normalized to 1% input in various regions of <span class="html-italic">Esr1</span> of fetal brains exposed in utero to saline or EtOH. (<b>C</b>). The association of the H3K4me3:H3 ratio was normalized to 1% input in various regions of <span class="html-italic">Esr1</span> of fetal brains exposed in utero to saline or EtOH. The significance of differences was determined by two-way ANOVA and a post hoc Tukey’s test. See <a href="#app1-cells-12-02308" class="html-app">Supplementary Figure S3</a> for controls and <a href="#app1-cells-12-02308" class="html-app">Figures S4 and S5</a> for sex-separated data.</p>
Full article ">Figure 7
<p>Increased hyperactivity for saline- but not EtOH-exposed <span class="html-italic">Ogg1</span> −/− vs. +/+ females, and decreased centre time for EtOH-exposed <span class="html-italic">Ogg1</span> +/+ but not −/− female mice. For all behavioural studies, pregnant dams were treated with single dose of EtOH (2 g/kg i.p.) or its saline vehicle on GD 17, as shown in <a href="#cells-12-02308-f001" class="html-fig">Figure 1</a>, and progeny were delivered spontaneously. Fetal brains from at least three litters were used to minimize potential litter effects, and the number of mice tested for each group is shown in parentheses. Results show total distance travelled during the entire test (1 h), total distance travelled during the last 30 min of the test and total time spent in the centre zone (10 × 10 inches). The significance of differences was determined by two-way ANOVA and a post hoc Tukey’s test. See <a href="#app1-cells-12-02308" class="html-app">Supplementary Figure S6a</a> for data for the nonsocial zone.</p>
Full article ">Figure 8
<p><b>OGG1- and sex-dependent effect on social interaction and interaction-induced ultrasonic vocalizations.</b> For all behavioural studies, pregnant dams were treated with a single dose of EtOH (2 g/kg i.p.) or its saline vehicle on GD 17, as described in <a href="#cells-12-02308-f001" class="html-fig">Figure 1</a>, and progeny were delivered spontaneously. Fetal brains from at least three litters were used to minimize potential litter effects, and the number of mice tested for each group is shown in parentheses. (Panel <b>A</b>). For social interaction, EtOH vs. saline decreased velocity and track length in <span class="html-italic">Ogg1</span> +/− progeny, but not in +/+ or −/− littermates, although the latter exhibited a similar non-significant trend. (Panel <b>B</b>). As with social interaction, EtOH vs. saline increased interaction-induced ultrasonic vocalizations in <span class="html-italic">Ogg1</span> +/− progeny, but not in +/+ or −/− littermates. The significance of differences was determined by two-way ANOVA and a post hoc Tukey’s test.</p>
Full article ">Figure 9
<p><b>Postulated epigenetic role of oxidative DNA damage initiated by reactive oxygen species (ROS) in neurodevelopmental disorders caused by physiological or ethanol-enhanced levels of ROS formation</b>. ROS, which are naturally produced in the body and essential for normal development and life, can oxidatively damage DNA, resulting in multiple types of lesions. The most prevalent DNA lesion, 8-oxoguanine (8-oxoG), is developmentally pathogenic, and is repaired by oxoguanine glycosylase 1 (OGG1). Heterozygous (+/−) and particularly homozygous (−/−) <span class="html-italic">Ogg1</span> knockout progeny have decreased DNA repair activity compared to their wild-type (+/+) littermates, and can accumulate oxidative DNA damage due to even physiological levels of ROS formation, leading to epigenetic changes in DNA methylation and modifications to histone proteins, exemplified by histone methylation and acetylation. These epigenetic changes can alter the expression of developmentally important genes, leading to neurodevelopmental disorders. Prenatal exposure to ROS-enhancing drugs like alcohol (ethanol, EtOH) can further enhance both the complexity and magnitude of epigenetic changes initiated by oxidative DNA damage, increasing the spectrum and severity of neurodevelopmental disorders.</p>
Full article ">
19 pages, 1623 KiB  
Review
Astrocytes Are a Key Target for Neurotropic Viral Infection
by Maja Potokar, Robert Zorec and Jernej Jorgačevski
Cells 2023, 12(18), 2307; https://doi.org/10.3390/cells12182307 - 19 Sep 2023
Cited by 2 | Viewed by 1560
Abstract
Astrocytes are increasingly recognized as important viral host cells in the central nervous system. These cells can produce relatively high quantities of new virions. In part, this can be attributed to the characteristics of astrocyte metabolism and its abundant and dynamic cytoskeleton network. [...] Read more.
Astrocytes are increasingly recognized as important viral host cells in the central nervous system. These cells can produce relatively high quantities of new virions. In part, this can be attributed to the characteristics of astrocyte metabolism and its abundant and dynamic cytoskeleton network. Astrocytes are anatomically localized adjacent to interfaces between blood capillaries and brain parenchyma and between blood capillaries and brain ventricles. Moreover, astrocytes exhibit a larger membrane interface with the extracellular space than neurons. These properties, together with the expression of various and numerous viral entry receptors, a relatively high rate of endocytosis, and morphological plasticity of intracellular organelles, render astrocytes important target cells in neurotropic infections. In this review, we describe factors that mediate the high susceptibility of astrocytes to viral infection and replication, including the anatomic localization of astrocytes, morphology, expression of viral entry receptors, and various forms of autophagy. Full article
(This article belongs to the Special Issue Astrocytes in CNS Disorders)
Show Figures

Figure 1

Figure 1
<p>Entry of viruses through the epithelial barrier, the lymphatic spread of viruses and neuroinvasion through the blood–brain barrier. (<b>A</b>) After entry through the epithelium, viruses may continue with subepithelial invasion by exploiting the network of lymphatics to enter lymph nodes. Viruses enter porous lymphatic capillaries and/or efferent lymphatic vessels to access the venous system. Circulating blood is the fastest and most effective way of viral dissemination to different tissues, including the central nervous system (CNS). (<b>B</b>) Entry into the delicate CNS parenchyma is restricted by an elaborate barrier network, most notably by the blood–brain barrier (BBB). Certain viruses can pass the BBB by the transcellular route by transferring through the lining of microvascular endothelial cells, where they may or may not replicate. In addition, viruses can pass the BBB by paracellular traversal between adjacent endothelial cells, with or without disruption of tight junctions. Viruses can also penetrate the BBB by transmigration within infected haematopoietic cells, known as the “Trojan horse” mechanism.</p>
Full article ">Figure 2
<p>Viral entry receptors in astrocytes. Diverse plasma membrane receptors can mediate virus attachment to and entry into astrocytes. These include neuropilin-1 (NRP1), angiotensin-converting enzyme 2 (ACE2), type II transmembrane serine protease (TMPRSS2), TAM family receptors (TYRO-3 and AXL), dendritic cell (DC)-specific intercellular adhesion molecule-3-grabbing nonintegrin (DC-SIGN), T-cell Ig and mucin domain 1 (TIM-1), heparan sulphate proteoglycans (HSPGs) comprising syndecans and glypicans, αvβ3 integrin, dipeptidylpeptidase 4 (DPP4), and cluster of differentiation 147 (CD147).</p>
Full article ">Figure 3
<p>Tick-borne encephalitis virus (TBEV) and West Nile virus (WNV), but not mosquito-only flavivirus (MOF), increase autophagy in human astrocytes. (<b>a</b>) Representative fluorescence micrographs documenting mock-infected human astrocytes (Control) and astrocytes after exposure to selected flaviviruses (TBEV, WNV, and MOF) for 48 h. Cells were expressing mRFP-EGFP-LC3, where LC3, a marker of autophagic compartments, is tandem fluorescent-tagged with mRFP (red fluorescence) and EGFP (green fluorescence). The pH of autophagosomes (AP) is close to neutral, which facilitates the fluorescence of both fluorophores, resulting in yellow objects. Fusion of AP with lysosomes yields autolysosomes (AL), i.e., organelles with acidic pH, where the EGFP fluorescence is quenched, resulting in red-only objects. Selected rectangular areas within the cells, enlarged at the corners (bottom, right), show superimposed images of mRFP and EGFP fluorescence. Arrows indicate AL. Adjacent, smaller panels display mRFP and EGFP fluorescence and co-localization masks (Col.mask) (co-localized objects correspond to AP) of the enlarged images. The white outlines in the large panels show the cell shape. (<b>b</b>) The total number of autophagic compartments (<b>i</b>) and the ratio of AL to AP (which is a measure of autophagic degradation activity) (<b>ii</b>) in mock-infected cells and after infection with TBEV, WNV, and MOF at 12, 24, and 48 h post-infection (hpi). Infection with TBEV increases the total number of autophagic compartments at all three time points tested, compared with mock-infected cells (i.e., controls at 48 hpi, which were confirmed to be comparable to controls at 12 and 24 hpi) (* <span class="html-italic">p</span> &lt; 0.05, one-way ANOVA followed by Dunn’s test), but does not affect the AL:AP ratio (<span class="html-italic">p</span>  &gt;  0.05, one-way ANOVA). WNV infection induces an increase in the total number of autophagic compartments at 48 hpi (* <span class="html-italic">p</span> &lt;  0.05, one-way ANOVA followed by Dunn’s test) and does not affect the ratio AL:AP (<span class="html-italic">p</span> &gt; 0.05, one-way ANOVA). MOF infection does not affect the number of autophagic compartments or the AL: AP ratio compared with mock-infected cells at any time point tested (<span class="html-italic">p</span> &gt;  0.05, one-way ANOVA). (<b>c</b>) Diameter of AP and AL in mock-infected cells and 48 hpi with selected flaviviruses. Infection with TBEV, WNV and MOF does not affect the size of the autophagic compartment compared with mock-infected cells (<span class="html-italic">p</span>  &gt;  0.05, one-way ANOVA). ALs are larger than APs in all experimental conditions (*** <span class="html-italic">p</span> &lt; 0.001, Mann-Whitney U test). Full lines in the boxplots represent median values, and the dotted lines correspond to average values. The numbers below the boxplots are the number of cells (<b>b</b>) or compartments (<b>c</b>) analysed for each condition. Cells were infected with TBEV and MOF at an MOI of 0.1 and with WNV at an MOI of 1. Figure and figure legend reproduced from Tavcar Verdev et al. [<a href="#B46-cells-12-02307" class="html-bibr">46</a>] with permission, licensed under a Creative Commons Attribution 4.0 International License (<a href="http://creativecommons.org/licenses/by/4.0" target="_blank">http://creativecommons.org/licenses/by/4.0</a> (accessed on 1 August 2023)).</p>
Full article ">
23 pages, 2332 KiB  
Article
Impact of Elevated Brain IL-6 in Transgenic Mice on the Behavioral and Neurochemical Consequences of Chronic Alcohol Exposure
by Donna L. Gruol, Delilah Calderon, Salvador Huitron-Resendiz, Chelsea Cates-Gatto and Amanda J. Roberts
Cells 2023, 12(18), 2306; https://doi.org/10.3390/cells12182306 - 19 Sep 2023
Cited by 1 | Viewed by 1117
Abstract
Alcohol consumption activates the neuroimmune system of the brain, a system in which brain astrocytes and microglia play dominant roles. These glial cells normally produce low levels of neuroimmune factors, which are important signaling factors and regulators of brain function. Alcohol activation of [...] Read more.
Alcohol consumption activates the neuroimmune system of the brain, a system in which brain astrocytes and microglia play dominant roles. These glial cells normally produce low levels of neuroimmune factors, which are important signaling factors and regulators of brain function. Alcohol activation of the neuroimmune system is known to dysregulate the production of neuroimmune factors, such as the cytokine IL-6, thereby changing the neuroimmune status of the brain, which could impact the actions of alcohol. The consequences of neuroimmune–alcohol interactions are not fully known. In the current studies we investigated this issue in transgenic (TG) mice with altered neuroimmune status relative to IL-6. The TG mice express elevated levels of astrocyte-produced IL-6, a condition known to occur with alcohol exposure. Standard behavioral tests of alcohol drinking and negative affect/emotionality were carried out in homozygous and heterozygous TG mice and control mice to assess the impact of neuroimmune status on the actions of chronic intermittent alcohol (ethanol) (CIE) exposure on these behaviors. The expressions of signal transduction and synaptic proteins were also assessed by Western blot to identify the impact of alcohol–neuroimmune interactions on brain neurochemistry. The results from these studies show that neuroimmune status with respect to IL-6 significantly impacts the effects of alcohol on multiple levels. Full article
(This article belongs to the Special Issue Alcohol and Neuroimmunology)
Show Figures

Figure 1

Figure 1
<p>Diagram of alcohol exposure protocol.</p>
Full article ">Figure 2
<p>Effects of CIE on 2BC alcohol drinking. (<b>A</b>,<b>B</b>). Graphs showing effect of cycle on average weekly alcohol intake (mean ± SEM) in male (<b>A</b>) and female (<b>B</b>) WT, +/−TG, and +/+TG mice in 2BC and 2BC-CIE groups during baseline and four cycles of exposure to air (air 1–4) or CIE (CIE 1–4). * = significant increase in drinking relative to baseline drinking (blue, WT; green, +/−TG; red, +/+TG; data from +/+TG mice were not included in the statistical analyses but are shown for reference). Numbers in parentheses are number of animals studied. In this and all other figures, a statistically significant difference is defined as <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 3
<p>Graphs showing effects of 2BC and 2BC-CIE on behavioral tests for negative affect/emotionality during the abstinence period. CIE actions are reflected in differences between effects of 2BC and 2BC-CIE. Behavioral tests include light/dark transfer (measures of time in light chamber) (<b>A1</b>), and total number of transitions between dark and light chamber (<b>A2</b>), open field test (distance traveled) (<b>B</b>) and immobility time (forced swim test) (<b>C</b>). In this and other figures, bar graphs show mean ± SEM values and individual data points. Numbers within bars or under bars are number of animals tested. Bars underneath symbols for significance indicate 2BC plus 2BC-CIE data are combined (i.e., no treatment effect). # = significantly different from WT, &amp; = significantly different from +/−TG, @ = overall 2BC-CIE significantly lower than 2BC.</p>
Full article ">Figure 4
<p>Effects of 2BC and 2BC-CIE on neuroimmune-related proteins. (<b>A</b>,<b>B</b>). Mean (± SEM) values for levels of IL-6 (<b>A</b>) and TNF-alpha (<b>B</b>) measured by ELISA in hippocampi from 2BC- and 2BC-CIE-treated WT, +/−TG, and +/+TG mice. (<b>C</b>,<b>D</b>). Mean (± SEM) levels of IL-6 signal transduction partners STAT3 (<b>C1</b>), pSTAT3 (<b>C2</b>), p42MAPK (<b>D1</b>), and pp42MAPK (<b>D2</b>), measured by Western blot analysis in hippocampi from 2BC- and 2BC-CIE-treated WT, +/−TG, and +/+TG mice. Representative Western blots are inserted above the corresponding data bars. Β = beta-actin (42 kD). The source Western blots are included in the <a href="#app1-cells-12-02306" class="html-app">Supplemental section</a> (<a href="#app1-cells-12-02306" class="html-app">Figure S1</a>). # = significantly different from WT. * = significant difference between 2BC and 2BC-CIE of the same genotype. Note, on these and other graphs showing data points, some data points may be masked by overlapping values. Numbers within or above the bars show the number of animals in the sample.</p>
Full article ">Figure 5
<p>Effects of 2BC and 2BC-CIE on synaptic proteins. (<b>A</b>–<b>G</b>). Mean (± SEM) values for levels of GABA<sub>A</sub>R alpha-1 (<b>A1</b>) and alpha-5 (<b>A2</b>), VGAT (<b>B</b>), GAD65 (<b>C1</b>) and GAD 67 (<b>C2</b>), gephyrin (<b>D</b>), GluR1 naïve, NR1 (<b>F</b>), and PSD-95 (<b>G</b>) determined by Western blot in WT, +/−TG, and +/+TG mice exposed to 2BC/abstinence or 2BC-CIE-abstinence protocols. The source Western blots are included in the <a href="#app1-cells-12-02306" class="html-app">Supplemental section</a> (<a href="#app1-cells-12-02306" class="html-app">Figure S2</a>). # = significantly different from WT, &amp; = significantly different from +/−TG, @ = overall 2BC-CIE significantly lower than 2BC.</p>
Full article ">
21 pages, 21543 KiB  
Article
scRNA-seq Reveals the Mechanism of Fatty Acid Desaturase 2 Mutation to Repress Leaf Growth in Peanut (Arachis hypogaea L.)
by Puxuan Du, Quanqing Deng, Wenyi Wang, Vanika Garg, Qing Lu, Lu Huang, Runfeng Wang, Haifen Li, Dongxin Huai, Xiaoping Chen, Rajeev K. Varshney, Yanbin Hong and Hao Liu
Cells 2023, 12(18), 2305; https://doi.org/10.3390/cells12182305 - 19 Sep 2023
Cited by 1 | Viewed by 1713
Abstract
Fatty Acid Desaturase 2 (FAD2) controls the conversion of oleic acids into linoleic acids. Mutations in FAD2 not only increase the high-oleic content, but also repress the leaf growth. However, the mechanism by which FAD2 regulates the growth pathway has not [...] Read more.
Fatty Acid Desaturase 2 (FAD2) controls the conversion of oleic acids into linoleic acids. Mutations in FAD2 not only increase the high-oleic content, but also repress the leaf growth. However, the mechanism by which FAD2 regulates the growth pathway has not been elucidated in peanut leaves with single-cell resolution. In this study, we isolated fad2 mutant leaf protoplast cells to perform single-cell RNA sequencing. Approximately 24,988 individual cells with 10,249 expressed genes were classified into five major cell types. A comparative analysis of 3495 differentially expressed genes (DEGs) in distinct cell types demonstrated that fad2 inhibited the expression of the cytokinin synthesis gene LOG in vascular cells, thereby repressing leaf growth. Further, pseudo-time trajectory analysis indicated that fad2 repressed leaf cell differentiation, and cell-cycle evidence displayed that fad2 perturbed the normal cell cycle to induce the majority of cells to drop into the S phase. Additionally, important transcription factors were filtered from the DEG profiles that connected the network involved in high-oleic acid accumulation (WRKY6), activated the hormone pathway (WRKY23, ERF109), and potentially regulated leaf growth (ERF6, MYB102, WRKY30). Collectively, our study describes different gene atlases in high-oleic and normal peanut seedling leaves, providing novel biological insights to elucidate the molecular mechanism of the high-oleic peanut-associated agronomic trait at the single-cell level. Full article
(This article belongs to the Collection Feature Papers in Plant, Algae and Fungi Cell Biology)
Show Figures

Figure 1

Figure 1
<p>Phenotypic variations of peanut seedlings between high-OA cultivar Yueyou271 and normal-OA cultivar Yueyou43. (<b>A</b>) The growth phenotype of seedlings of the two varieties on day 3, 5, and 7 after sowing. (<b>B</b>–<b>D</b>) Leaf area, stem length, leaf length, and width of seedlings of the two varieties on day 3, day 5, and day 7 (<span class="html-italic">n</span> = 5). (<b>E</b>) Comparison of the cytokinin derivatives contents in seedling leaves between Yueyou43 and Yueyou271. Histograms depict the mean ± SD of three biological replicates. cZ, <span class="html-italic">cis</span>-Zeatin; tZ, <span class="html-italic">trans</span>-Zeatin; tZR, <span class="html-italic">trans</span>-Zeatin riboside; IPR, N6-isopentenyladenosine; DHZR, Dihydrozeatin ribonucleoside. The asterisks indicate significant differences between the two varieties (T-test, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01). (<b>F</b>) A model representing the high-OA accumulation mediated by <span class="html-italic">fad2</span> mutation to repress the cytokinin pathway.</p>
Full article ">Figure 2
<p>Construction of single-cell transcriptome atlas and annotation of peanut leaf clusters. (<b>A</b>) An overview of the scRNA-seq workflow used. (<b>B</b>,<b>C</b>) Visualization of 12 cell clusters using t-SNE plot; each dot indicates individual cells colored based on variety and cell clusters. (<b>D</b>) Bar plot depicting distribution of cells from high-OA (Yueyou271) and normal peanut (Yueyou43) in the 12 clusters. (<b>E</b>) Violin plots showing the expression pattern of known marker genes across clusters. (<b>F</b>–<b>H</b>) Circos plots consisted of 20 peanut chromosomes, representing the single-cell gene expression pattern in the leaf cells of total high-OA and normal peanuts, respectively. The outer circle to inner circle represents the cell clusters 0 to 11.</p>
Full article ">Figure 3
<p>scRNA-seq identifies differentially expressed genes (DEGs) in distinct leaf cell types. (<b>A</b>) The number of DEGs identified in each cluster. (<b>B</b>) KEGG pathway enrichment analysis of all DEGs from all clusters. (<b>C</b>) Venn diagram showing the 804 core DEGs across all clusters. (<b>D</b>,<b>E</b>) Expression matrix of 35 hormone signaling DEGs and 32 differentially expressed TFs in each cell cluster. (<b>F</b>) The interaction network of 32 differentially expressed TFs. (<b>G</b>) Bar plots illustrating the up- and down-regulated DEGs in each cell type. (<b>H</b>) Venn diagram showing the 1649 core-DEGs across all cell types. (<b>I</b>) Heatmap depicting the expression level of 65 hub-TFs in each cell type of the two varieties. (<b>J</b>,<b>K</b>) The expression distribution of four <span class="html-italic">LOG</span> genes in all cell clusters, with the gray dots as background representing the cells with no expression of the given transcript. (<b>L</b>) Dot plots show the expression pattern and distribution of four <span class="html-italic">LOG</span> genes in distinct cell types. (<b>M</b>) A putative model illustrating that the down-regulation of <span class="html-italic">LOG</span> in Yueyou271 leads to a decrease in cytokinin content.</p>
Full article ">Figure 4
<p>Pseudo-time trajectory analysis of cell types in one-week-old peanut seedling leaves. (<b>A</b>) The cell ordering along the differentiation trajectory presented by cell differentiation states, samples, and cell clusters. (<b>B</b>) Venn diagram showing the core DEGs from DEGs of cell differentiation trajectory, cell differentiation states, and cell fate. (<b>C</b>) KEGG pathway enrichment analysis of 11,914 core DEGs in leaf development trajectory. (<b>D</b>) Clustering and expression kinetics of 520 TFs in 11,914 core DEGs along cell differentiation states of total leaf cell ontology. (<b>E</b>) KEGG pathway enrichment analysis of 520 TFs in 11,914 core DEGs. (<b>F</b>,<b>G</b>) The cell ordering along the PAGA trajectory is presented by samples and cell clusters. (<b>H</b>) Venn diagram showing the 1251 core DEGs across both cell trajectories result. (<b>I</b>) Dot plots showing the expression pattern of 48 critical TFs in each cell cluster.</p>
Full article ">Figure 5
<p>Separate development trajectory of five cell types. (<b>A</b>) Cell differentiation state, sample, and cell cluster distributions followed the pseudo-time trajectory of mesophyll development. (<b>B</b>,<b>C</b>) Clustering and expression kinetics of 66 TFs in 1773 DEGs along cell differentiation states of the mesophyll cell group. (<b>D</b>) Cell differentiation state, sample, and cell cluster distributions followed the pseudo-time trajectory of primordium development. (<b>E</b>,<b>F</b>) Clustering and expression kinetics of 17 TFs in 386 DEGs along cell differentiation states of the primordium cell group. (<b>G</b>) Cell differentiation state, sample, and cell cluster distributions followed the pseudo-time trajectory of vascular development. (<b>H</b>,<b>I</b>) Clustering and expression kinetics of 361 TFs in 5512 DEGs along cell differentiation states of the vascular cells. (<b>J</b>) PAGA trajectory of differentiation from the epidermal to the guard cells. (<b>K</b>) Expression tendency of 380 DEGs in PAGA trajectory. (<b>L</b>) Expression tendency of 6 TFs in 380 DEGs in the process of epidermis transforming into guard cells.</p>
Full article ">Figure 6
<p>Cell cycle analysis provides important DEGs regulating cell cycle differences between the high-OA and normal seedling leaves. (<b>A</b>) RNA velocity analysis of all cells. The number 0–11 indicates the cell cluster 0–11 in the <span class="html-italic">fad2</span> and normal peanut seedling leaf t-SNE map. (<b>B</b>) Cell cycle phase distribution; NC indicates the non-cycling cell population. The scaleplate of out circular represents the total cell number in each cell cycle phase. (<b>C</b>) The histogram plot shows the distribution of cells from respective cell cycle phases. (<b>D</b>) The histogram plot shows the DEGs in different cell cycle phases. (<b>E</b>) GO enrichment analysis of DEGs in different cell cycle phases. (<b>F</b>) Newly identified genes for distinguishing leaf cell genome replication states. (<b>G</b>) Venn diagram showing 1113 core DEGs between cell-cycle-related DEGs, cell cluster DEGs, and cell development trajectory DEGs. (<b>H</b>) KEGG pathway enrichment analysis of 1113 core DEGs. (<b>I</b>) Cell expression distribution of eight core TFs identified from 1113 core DEGs.</p>
Full article ">Figure 7
<p>Putative model of <span class="html-italic">FAD2</span> mutant regulated the cytokinin pathway to repress the leaf cell development in peanut leaf vascular tissue. DMAPP, dimethylallyl diphosphate; iPRMP, N(6)-(Delta(2)-isopentenyl)adenosine-5′-monophosphate; cZRMP, <span class="html-italic">cis</span>-Zeatin riboside monophosphate; tZRMP, <span class="html-italic">trans</span>-Zeatin riboside monophosphate; cZR, <span class="html-italic">cis</span>-Zeatin riboside; tZR, <span class="html-italic">trans</span>-Zeatin riboside; cZ, <span class="html-italic">cis</span>-Zeatin; tZ, <span class="html-italic">trans</span>-Zeatin; 13-HPOT, 13-hydroperoxide of alpha-linolenic acid; 12,13-EOT, 12,13(S)-epoxy-9(Z),11,15(Z)-octadecatrienoic acid; 12-OPDA, 12-oxophytodienoic acid; OPC8, 3-oxo-2-(2-(Z)-pentenyl) cyclopentane-1-octanoic acid; JA, jasmonic acid; KAS1, 3-ketoacyl-acyl carrier protein synthase 1; KAS2, 3-ketoacyl-acyl carrier protein synthase 2; FAB2, stearoyl-ACP desaturase 2; AOS, allene oxide synthase; AOC, alleneoxide cyclase; ACX1, acyl-CoA oxidase1.</p>
Full article ">
14 pages, 1015 KiB  
Review
Riboflavin and Its Derivates as Potential Photosensitizers in the Photodynamic Treatment of Skin Cancers
by Małgorzata Insińska-Rak, Marek Sikorski and Agnieszka Wolnicka-Glubisz
Cells 2023, 12(18), 2304; https://doi.org/10.3390/cells12182304 - 19 Sep 2023
Cited by 4 | Viewed by 2340
Abstract
Riboflavin, a water-soluble vitamin B2, possesses unique biological and physicochemical properties. Its photosensitizing properties make it suitable for various biological applications, such as pathogen inactivation and photodynamic therapy. However, the effectiveness of riboflavin as a photosensitizer is hindered by its degradation upon exposure [...] Read more.
Riboflavin, a water-soluble vitamin B2, possesses unique biological and physicochemical properties. Its photosensitizing properties make it suitable for various biological applications, such as pathogen inactivation and photodynamic therapy. However, the effectiveness of riboflavin as a photosensitizer is hindered by its degradation upon exposure to light. The review aims to highlight the significance of riboflavin and its derivatives as potential photosensitizers for use in photodynamic therapy. Additionally, a concise overview of photodynamic therapy and utilization of blue light in dermatology is provided, as well as the photochemistry and photobiophysics of riboflavin and its derivatives. Particular emphasis is given to the latest findings on the use of acetylated 3-methyltetraacetyl-riboflavin derivative (3MeTARF) in photodynamic therapy. Full article
Show Figures

Figure 1

Figure 1
<p>Structures of riboflavin and some of its derivatives.</p>
Full article ">Figure 2
<p>Diagram of the biological effects of riboflavin or its derivatives on pathogenic microorganisms or cancer cells after UVA or blue light activation [based on [<a href="#B34-cells-12-02304" class="html-bibr">34</a>,<a href="#B43-cells-12-02304" class="html-bibr">43</a>,<a href="#B84-cells-12-02304" class="html-bibr">84</a>,<a href="#B85-cells-12-02304" class="html-bibr">85</a>,<a href="#B86-cells-12-02304" class="html-bibr">86</a>,<a href="#B87-cells-12-02304" class="html-bibr">87</a>]].</p>
Full article ">Figure 3
<p>Scheme of riboflavin and 3MeTARF action in tumor cells upon blue light activation. RCP: riboflavin-binding protein; p: phosphorylation [based on [<a href="#B34-cells-12-02304" class="html-bibr">34</a>,<a href="#B85-cells-12-02304" class="html-bibr">85</a>]].</p>
Full article ">
19 pages, 1217 KiB  
Review
Multiple Roles of the RUNX Gene Family in Hepatocellular Carcinoma and Their Potential Clinical Implications
by Milena Krajnović, Bojana Kožik, Ana Božović and Snežana Jovanović-Ćupić
Cells 2023, 12(18), 2303; https://doi.org/10.3390/cells12182303 - 19 Sep 2023
Cited by 1 | Viewed by 1291
Abstract
Hepatocellular carcinoma (HCC) is one of the most frequent cancers in humans, characterised by a high resistance to conventional chemotherapy, late diagnosis, and a high mortality rate. It is necessary to elucidate the molecular mechanisms involved in hepatocarcinogenesis to improve diagnosis and treatment [...] Read more.
Hepatocellular carcinoma (HCC) is one of the most frequent cancers in humans, characterised by a high resistance to conventional chemotherapy, late diagnosis, and a high mortality rate. It is necessary to elucidate the molecular mechanisms involved in hepatocarcinogenesis to improve diagnosis and treatment outcomes. The Runt-related (RUNX) family of transcription factors (RUNX1, RUNX2, and RUNX3) participates in cardinal biological processes and plays paramount roles in the pathogenesis of numerous human malignancies. Their role is often controversial as they can act as oncogenes or tumour suppressors and depends on cellular context. Evidence shows that deregulated RUNX genes may be involved in hepatocarcinogenesis from the earliest to the latest stages. In this review, we summarise the topical evidence on the roles of RUNX gene family members in HCC. We discuss their possible application as non-invasive molecular markers for early diagnosis, prognosis, and development of novel treatment strategies in HCC patients. Full article
(This article belongs to the Special Issue Emerging Therapeutic Approaches for Chronic Liver Diseases)
Show Figures

Figure 1

Figure 1
<p>The structures of RUNX1, RUNX2, and RUNX3 genes and proteins. (<b>A</b>) <span class="html-italic">RUNX1</span>, <span class="html-italic">RUNX2</span>, and <span class="html-italic">RUNX3</span> genes’ structure. Rectangles—exons; lines—introns; ATG—start codon; P1 and P2—promoters, RHD—Runt homology domain; and TAD—transactivation domain. Grey rectangles—untranslated regions. (<b>B</b>) RUNX1, RUNX2, and RUNX3 proteins’ structure. Rectangles—protein domains. Numbers—amino acids’ numbers. NLS—nuclear localisation signal; QA—the glutamine/alanine-rich signal, RUNX2 specific; ID—inhibitory domain; and VWRPY—Groucho/TLE binding site. The figure is a not-to-scale drawing. We created the figure under CC BY NC, based on Yi et al., 2022 [<a href="#B10-cells-12-02303" class="html-bibr">10</a>].</p>
Full article ">Figure 2
<p>Roles of <span class="html-italic">RUNX1</span> in hepatocellular carcinoma. COL4A1↑-Collagen type IV alpha 1 chain increases; VEGF↓-Vascular endothelial growth factor decreases.</p>
Full article ">Figure 3
<p><span class="html-italic">RUNX2</span> oncogenic mechanisms in hepatocellular carcinoma.</p>
Full article ">Figure 4
<p>The downregulation of <span class="html-italic">RUNX3</span> in HCC.</p>
Full article ">
19 pages, 13972 KiB  
Article
Pancreatic Islet Viability Assessment Using Hyperspectral Imaging of Autofluorescence
by Jared M. Campbell, Stacey N. Walters, Abbas Habibalahi, Saabah B. Mahbub, Ayad G. Anwer, Shannon Handley, Shane T. Grey and Ewa M. Goldys
Cells 2023, 12(18), 2302; https://doi.org/10.3390/cells12182302 - 19 Sep 2023
Cited by 2 | Viewed by 1492
Abstract
Islets prepared for transplantation into type 1 diabetes patients are exposed to compromising intrinsic and extrinsic factors that contribute to early graft failure, necessitating repeated islet infusions for clinical insulin independence. A lack of reliable pre-transplant measures to determine islet viability severely limits [...] Read more.
Islets prepared for transplantation into type 1 diabetes patients are exposed to compromising intrinsic and extrinsic factors that contribute to early graft failure, necessitating repeated islet infusions for clinical insulin independence. A lack of reliable pre-transplant measures to determine islet viability severely limits the success of islet transplantation and will limit future beta cell replacement strategies. We applied hyperspectral fluorescent microscopy to determine whether we could non-invasively detect islet damage induced by oxidative stress, hypoxia, cytokine injury, and warm ischaemia, and so predict transplant outcomes in a mouse model. In assessing islet spectral signals for NAD(P)H, flavins, collagen-I, and cytochrome-C in intact islets, we distinguished islets compromised by oxidative stress (ROS) (AUC = 1.00), hypoxia (AUC = 0.69), cytokine exposure (AUC = 0.94), and warm ischaemia (AUC = 0.94) compared to islets harvested from pristine anaesthetised heart-beating mouse donors. Significantly, with unsupervised assessment we defined an autofluorescent score for ischaemic islets that accurately predicted the restoration of glucose control in diabetic recipients following transplantation. Similar results were obtained for islet single cell suspensions, suggesting translational utility in the context of emerging beta cell replacement strategies. These data show that the pre-transplant hyperspectral imaging of islet autofluorescence has promise for predicting islet viability and transplant success. Full article
(This article belongs to the Special Issue Islet Transplantation)
Show Figures

Figure 1

Figure 1
<p>Representative spectral images of islets subjected to ROS. (<b>A</b>) Control islet autofluorescence; (<b>B</b>) ROS damage islet autofluorescence. (<b>C</b>) False colour principal component analysis (PCA) image superimposed with brightfield image of control islets. (<b>D</b>) False colour PCA image superimposed with brightfield image of ROS damaged islets. The false colours in the PCA image visualise compressed pixels using different channels resulting from the PCA analysis, to enhance visual understanding of data clusters obtained from multivariate analysis.</p>
Full article ">Figure 2
<p>ROS damage induced by exposing islets to menadione. (<b>A</b>) Canonical discriminant analysis for the discrimination of control (blue), and exposed (red). (<b>B</b>) ROC curve showing complete discrimination (AUC = 1.0). Corresponding 95% confidence intervals are drawn as ellipses. Hypoxia damage induced by exposing islets to DMOG. (<b>C</b>) Canonical discriminant analysis for control (blue) islets and hypoxic (red) islets. Corresponding 95% confidence intervals are drawn as ellipses. (<b>D</b>) ROC curve showing partial discrimination (AUC = 0.69).</p>
Full article ">Figure 3
<p>Impact of ROS exposure (induced by menadione) on relative fluorophore levels. (<b>A</b>) NAD(P)H (indicating the combined signals of NADH and NADPH, of which NADH is most plentiful), (<b>B</b>) flavins (indicating the combined signals of flavin family members, of which the metabolic coenzyme to NADH, FAD, is the most plentiful), (<b>C</b>) Col-I (collagen-I), (<b>D</b>) redox ratio (RR = NAD(P)H/flavins), and (<b>E</b>) cytochrome-C. Data are means with 95% CI, * shows <span class="html-italic">p</span> = 0.0444. In all cases, n = 24 for control and 25 for ROS.</p>
Full article ">Figure 4
<p>Impact of hypoxia-relative fluorophore levels. (<b>A</b>) NAD(P)H, (<b>B</b>) flavins, (<b>C</b>) Col-I (collagen-I, (<b>D</b>) redox ratio (NAD(P)H/flavins), and (<b>E</b>) cytochrome-C. Data are means with 95% CI, * shows <span class="html-italic">p</span> &lt; 0.05, ** shows <span class="html-italic">p</span> &lt; 0.01, *** shows <span class="html-italic">p</span> &lt; 0.001. In all cases, n = 36 for control and 32 for hypoxia.</p>
Full article ">Figure 5
<p>Hyperspectral signature of moderate (4 h) and major (24 h) pro-inflammatory signalling. Canonical discriminant analysis for differentiating islets exposed moderate pro-inflammatory signalling from pristine islets with the corresponding ROC curve is shown in (<b>A</b>,<b>B</b>), the same for differentiating major pro-inflammatory signalling from pristine is shown in (<b>C</b>,<b>D</b>), and the differentiation of moderate from major is shown in (<b>E</b>,<b>F</b>). Corresponding 95% confidence intervals are drawn as ellipses.</p>
Full article ">Figure 6
<p>Impact of pro-inflammatory cytokine exposure on relative fluorophore levels. (<b>A</b>) NAD(P)H, (<b>B</b>) flavins, (<b>C</b>) Col-I (collagen-I), (<b>D</b>) redox ratio (NAD(P)H/flavins), and (<b>E</b>) cytochrome-C. Data are means with 95% confidence interval, in all cases n = 34 for control 4 h (Cont-4 h), 48 for cytokine exposure 4 h (Cyto-4 h), 36 for control 24 h (Cont-24 h), and 50 for cytokine 24 h (Cyto 24 h). *, **, and *** indicate significant differences at <span class="html-italic">p</span> &lt; 0.05, 0.005, and 0.0001.</p>
Full article ">Figure 7
<p>Hyperspectral signature of moderate (30 min) and major (60 min) exposure to warm ischaemia. Canonical discriminant analysis for differentiating islets exposed to moderate warm ischaemia from pristine islets with the corresponding ROC curve is shown in (<b>A</b>,<b>B</b>), the same for differentiating major warm ischaemia from pristine islets is shown in (<b>C</b>,<b>D</b>), and the differentiation of moderate from major is shown in (<b>E</b>,<b>F</b>). Corresponding 95% confidence intervals are drawn as ellipses.</p>
Full article ">Figure 8
<p>Impact of warm ischaemia on relative fluorophore levels. (<b>A</b>) NAD(P)H, (<b>B</b>) flavins, (<b>C</b>) Col-I (collagen-I), (<b>D</b>) redox ratio (RR), and (<b>E</b>) cytochrome-C. Data are means with 95% CI. In all cases n = 79 for control islets, 72 for islets from pancreata left in mice for 30 min post mortem, and 62 for islets left in mice for 60 min post mortem. *, **, and *** indicate significant differences at <span class="html-italic">p</span> &lt; 0.05, 0.005, and 0.0001.</p>
Full article ">Figure 9
<p>Unsupervised assessment of hyperspectral data for the transplanted islets. Principal component analysis (PCA) was applied for multidimensional feature vectors for all islets. (<b>A</b>) Blood glucose control after islet transplantation into mice. Islet sets labelled as treatment group (control or Ischaemic), replicate number: number within replicate. (<b>B</b>) Islets which restored glucose control (functional, blue dots) plotted against islets which did not restore glucose control (red dots) in the space spanned by the first and second PCA component. (<b>C</b>) Box plots of the values of the second PCA component for the different islet sets where C = control and I = ischaemic treated group. A threshold line was able to be drawn, which lay above the median value for all functional islet preparations and below the median value of all non-functional islet preparations, giving 100% accuracy for discrimination at the group level. (<b>D</b>) ROC curve for individual islets, AUC = 0.81.</p>
Full article ">
Previous Issue
Back to TopTop