[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (656)

Search Parameters:
Keywords = molecular motor

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
12 pages, 2544 KiB  
Article
A Novel MAG Variant Causes Hereditary Spastic Paraplegia in a Consanguineous Pakistani Family
by Rabia Akram, Haseeb Anwar, Humaira Muzaffar, Valentina Turchetti, Tracy Lau, Barbara Vona, Ehtisham Ul Haq Makhdoom, Javed Iqbal, Shahid Mahmood Baig, Ghulam Hussain, Stephanie Efthymiou and Henry Houlden
Genes 2024, 15(9), 1203; https://doi.org/10.3390/genes15091203 - 13 Sep 2024
Viewed by 378
Abstract
Background and objectives: Hereditary spastic paraplegia (HSP) is characterized by unsteady gait, motor incoordination, speech impairment, abnormal eye movement, progressive spasticity and lower limb weakness. Spastic paraplegia 75 (SPG75) results from a mutation in the gene that encodes myelin associated glycoprotein (MAG). Only [...] Read more.
Background and objectives: Hereditary spastic paraplegia (HSP) is characterized by unsteady gait, motor incoordination, speech impairment, abnormal eye movement, progressive spasticity and lower limb weakness. Spastic paraplegia 75 (SPG75) results from a mutation in the gene that encodes myelin associated glycoprotein (MAG). Only a limited number of MAG variants associated with SPG75 in families of European, Middle Eastern, North African, Turkish and Palestinian ancestry have been documented so far. This study aims to provide further insight into the clinical and molecular manifestations of HSP. Methods: Using whole-exome sequencing, we investigated a consanguineous Pakistani family where three individuals presented with clinical signs of HSP. Sanger sequencing was used to carry out segregation analysis on available family members, and a minigene splicing assay was utilized to evaluate the effect of the splicing variant. Results: We identified a novel homozygous pathogenic splice donor variant in MAG (c.46 + 1G > T) associated with SPG75. RNA analysis revealed exon skipping that resulted in the loss of a start codon for ENST00000361922.8 isoform. Affected individuals exhibited variable combinations of nystagmus, developmental delay, cognitive impairments, spasticity, dysarthria, delayed gait and ataxia. The proband displayed a quadrupedal stride, and his siblings experienced frequent falls and ataxic gait as one of the prominent features that have not been previously reported in SPG75. Conclusions: Thus, the present study presents an uncommon manifestation of SPG75, the first from the Pakistani population, and broadens the spectrum of MAG variants. Full article
(This article belongs to the Section Human Genomics and Genetic Diseases)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Five-generation pedigree of the family showing three affected siblings. All affected individuals were homozygous (G &gt; T) while the mother was heterozygous (Proband: V.1; Affected sister: V.2; Affected sister: V.3). The sample for IV.1 was unavailable for further study. (<b>B</b>) Sequence chromatograms of <span class="html-italic">MAG</span> showing a likely pathogenic c.46 + 1G &gt; T variant.</p>
Full article ">Figure 2
<p>RNA functional studies of the <span class="html-italic">MAG</span> c.46 + 1G &gt; T variants. (<b>A</b>) RT-PCR amplicons for the <span class="html-italic">MAG</span> c.46 + 1G wild-type, c.46 + 1T mutant, and empty pSPL3 vector were separated by gel electrophoresis. The PCR and transfection negative controls performed as expected. (<b>B</b>) The in vitro splice assay’s vector construct displays the variant-containing (lower splice profile) and wild-type (upper splice profile) amplicons that are placed between pSPL3 vector’s exons A and B. Each variant’s splicing schematic is displayed below. The c.46 + 1G &gt; A variant causes skipping of exon 3, resulting in a deletion of 46 bp of coding exon 3 (r.1_46del), p.?, including the start codon of the ENST00000361922.8/NM_080600.3 isoforms. (<b>C</b>) Sequencing of the exon–exon junctions for the wild-type (left) and variant, appearing as the empty vector control (right).</p>
Full article ">Figure 3
<p>MAG interacts with RTN4R/NgR and prevents axonal sprouting. MAG contains five Ig domains (1–5) and a intramembrane segment. MAG: Myelin associated glycoprotein; RTN4R: reticulon-4 receptor; NgR: Nogo Receptor; Ig: Immunoglobulin domain; Transmembrane domain: TM. Created with BioRender.com, accessed on 6 September 2024.</p>
Full article ">
17 pages, 3252 KiB  
Article
Polymorphism and Pharmacological Assessment of Carbamazepine
by Alberto Sá Filho, Jose Luis Rodrigues Martins, Rafael Fernandes Costa, Gustavo Rodrigues Pedrino, Vitor Santos Duarte, Osmar Nascimento Silva, Hamilton Barbosa Napolitano and James Oluwagbamigbe Fajemiroye
Int. J. Mol. Sci. 2024, 25(18), 9835; https://doi.org/10.3390/ijms25189835 - 11 Sep 2024
Viewed by 212
Abstract
This work provides insight into carbamazepine polymorphs (Forms I, II, III, IV, and V), with reports on the cytoprotective, exploratory, motor, CNS-depressant, and anticonvulsant properties of carbamazepine (CBZ), carbamazepine formulation (CBZ-F), topiramate (TOP), oxcarbazepine (OXC), and diazepam (DZP) in mice. Structural analysis highlighted [...] Read more.
This work provides insight into carbamazepine polymorphs (Forms I, II, III, IV, and V), with reports on the cytoprotective, exploratory, motor, CNS-depressant, and anticonvulsant properties of carbamazepine (CBZ), carbamazepine formulation (CBZ-F), topiramate (TOP), oxcarbazepine (OXC), and diazepam (DZP) in mice. Structural analysis highlighted the significant difference in molecular conformations, which directly influence the physicochemical properties; and density functional theory description provided indications about CBZ reactivity and stability. In addition to neuron viability assessment in vitro, animals were treated orally with vehicle 10 mL/kg, as well as CBZ, CBZ-F, TOP, OXC, and DZP at the dose of 5 mg/kg and exposed to open-field, rotarod, barbiturate sleep induction and pentylenetetrazol (PTZ 70 mg/kg)-induced seizure. The involvement of GABAergic mechanisms in the activity of these drugs was evaluated with the intraperitoneal pretreatment of flumazenil (2 mg/kg). The CBZ, CBZ-F, and TOP mildly preserved neuronal viability. The CBZ-F and the reference AEDs potentiated barbiturate sleep, altered motor activities, and attenuated PTZ-induced convulsion. However, flumazenil pretreatment blocked these effects. Additional preclinical assessments could further establish the promising utility of CBZ-F in clinical settings while expanding the scope of AED formulations and designs. Full article
(This article belongs to the Section Molecular Pharmacology)
Show Figures

Figure 1

Figure 1
<p>Main interactions of the supramolecular arrangement of the carbamazepine forms. I (<b>a</b>), II (<b>b</b>), III (<b>c</b>), IV (<b>d</b>), and V (<b>e</b>).</p>
Full article ">Figure 2
<p>Molecular packing in the unit cell for carbamazepine forms. I (<b>a</b>), II (<b>b</b>), III (<b>c</b>), IV (<b>d</b>), and V (<b>e</b>).</p>
Full article ">Figure 3
<p>FMO representations for CBZ forms. I (<b>a</b>), II (<b>b</b>), III (<b>c</b>), IV (<b>d</b>), and V (<b>e</b>).</p>
Full article ">Figure 4
<p>MEP map for CBZ forms. I (<b>a</b>), II (<b>b</b>), III (<b>c</b>), IV (<b>d</b>), and V (<b>e</b>).</p>
Full article ">Figure 5
<p>Glutamate- and glycine-induced over-stimulation and excitotoxicity protocol for viability assessment in vitro. Cultured cortical neurons at 37 °C were treated with 10 µM of glutamate/1 µM of glycine for 1 h in the presence of VEH (vehicle), CBZ (carbamazepine), CBZ-F (carbamazepine formulation), TOP (topiramate), OXC (oxcarbazepine), and DZP (diazepam) at 1 mg/mL. The group of untreated cultured neuron without excitotoxic stimulation (sham) was assigned with 100% viability. Data are represented as mean ± SEM, <span class="html-italic">n</span> = 10 different cultures. * <span class="html-italic">p</span> &lt; 0.05 and *** <span class="html-italic">p</span> &lt; 0.001 as compared to the VEH; ### <span class="html-italic">p</span> &lt; 0.001 as compared with the sham group (one-way ANOVA followed by Bonferroni’s post hoc test).</p>
Full article ">Figure 6
<p>Effect of CBZ-F on the exploratory and motor activities in the open field and rotarod tests, respectively. Oral treatments with VEH (vehicle), CBZ (carbamazepine), CBZ-F (carbamazepine formulation), TOP (topiramate), OXC (oxcarbazepine) and DZP (diazepam), all at the dose of 5 mg/kg, were carried out. The parameters evaluated were total number of crossings (<b>A</b>) and number of rearings (<b>B</b>) in the open field as well as the fall latency (<b>C</b>) and the number of falls (<b>D</b>) in the rotarods. Each column represents mean ± S.E.M. (<span class="html-italic">n</span> = 10). One-way ANOVA and Dunnett’s post hoc test for multiple comparisons were performed. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001 for other treatment groups vs. vehicle-treated group.</p>
Full article ">Figure 7
<p>Potentiation and blockade of barbiturate sleep induction. Effect of orally administered vehicle (10 mL/kg), CBZ, CBZ-F, DZP, TOP, and OXC (all 5 mg/kg) on (<b>A</b>) sleep latency and (<b>B</b>) sleep duration of sodium pentobarbital (40 mg/kg)-induced hypno-sedative effect. * Indicates <span class="html-italic">p</span> &lt; 0.05, ** indicates <span class="html-italic">p</span> &lt; 0.01 and *** indicates <span class="html-italic">p</span> &lt; 0.001 as compared with vehicle-treated group (one-way ANOVA followed by Dunnett’s post hoc test). All groups were treated with sodium pentobarbital (40 mg/kg, i.p.). Results are expressed as mean ± SEM; <span class="html-italic">n</span> = 10 in each group. VEH (vehicle), CBZ (carbamazepine), CBZ-F (carbamazepine formulation), TOP (topiramate), OXC (oxcarbazepine), DZP (diazepam), and FLU (flumazenil).</p>
Full article ">Figure 8
<p>Screening for anticonvulsant activity and its pharmacological blockade using pentylenetetrazol-induced seizure test. The effects of the vehicle (10 mL/kg), CBZ, CBZ-F, TOP, OXC, or DZP at 5 mg/kg on the latency (<b>A</b>), duration (<b>B</b>), and severity (<b>C</b>) of seizure were analyzed by either ANOVA and Dunnettʼs post hoc test with results in mean ± S.E.M. (<span class="html-italic">n</span> = 10). Statistical analysis was performed by one-way ANOVA followed by Dunnett’s post hoc test. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001 for other treatment groups vs. VEH group regarding the antiseizure-like effect of CBZ, CBZ-F, TOP, OXC, or DZP. VEH (saline), CBZ (carbamazepine), CBZ-F (carbamazepine formulation), TOP (topiramate), OXC (oxcarbazepine), DZP (diazepam) and FLU (flumazenil).</p>
Full article ">
25 pages, 2418 KiB  
Article
Brain and Serum Membrane Vesicle (Exosome) Profiles in Experimental Alcohol-Related Brain Degeneration: Forging the Path to Non-Invasive Liquid Biopsy Diagnostics
by Suzanne M. De La Monte, Yiwen Yang and Ming Tong
J. Mol. Pathol. 2024, 5(3), 360-384; https://doi.org/10.3390/jmp5030025 - 10 Sep 2024
Viewed by 253
Abstract
Background: Alcohol-related brain degeneration (ARBD) is associated with cognitive–motor impairments that can progress to disability and dementia. White matter (WM) is prominently targeted in ARBD due to chronic neurotoxic and degenerative effects on oligodendrocytes and myelin. Early detection and monitoring of WM pathology [...] Read more.
Background: Alcohol-related brain degeneration (ARBD) is associated with cognitive–motor impairments that can progress to disability and dementia. White matter (WM) is prominently targeted in ARBD due to chronic neurotoxic and degenerative effects on oligodendrocytes and myelin. Early detection and monitoring of WM pathology in ARBD could lead to therapeutic interventions. Objective: This study examines the potential utility of a non-invasive strategy for detecting WM ARBD using exosomes isolated from serum. Comparative analyses were made with paired tissue (Tx) and membrane vesicles (MVs) from the temporal lobe (TL). Methods: Long Evans rats were fed for 8 weeks with isocaloric liquid diets containing 37% or 0% caloric ethanol (n = 8/group). TL-Tx, TL-MVs, and serum exosomes (S-EVs) were used to examine ethanol’s effects on oligodendrocyte glycoprotein, astrocyte, and oxidative stress markers. Results: Ethanol significantly decreased the TL-Tx expression of platelet-derived growth factor receptor alpha (PDGFRA), 2′,3′-cyclic nucleotide 3′ phosphodiesterase (CNPase), proteolipid protein (PLP), myelin oligodendrocyte glycoprotein (MOG), glial fibrillary acidic protein (GFAP), and 8-OHdG, whereas in the TL-MVs, ethanol increased CNPase, PDGFRA, and 8-OHdG, but decreased MOG and GFAP concordantly with TL-Tx. Ethanol modulated the S-EV expression by reducing PLP, nestin, GFAP, and 4-hydroxynonenal (HNE). Conclusion: Chronic ethanol exposures differentially alter the expression of oligodendrocyte/myelin, astrocyte, and oxidative stress markers in the brain, brain MVs, and S-EVs. However, directionally concordant effects across all three compartments were limited. Future studies should advance these efforts by characterizing the relationship between ABRD and molecular pathological changes in brain WM-specific exosomes in serum. Full article
Show Figures

Figure 1

Figure 1
<p>Nanotracker analysis (NTA) size (nm) × concentration (particles/mL) profiles of membrane vesicles/exosomes (EV) isolated from (<b>A</b>,<b>B</b>) temporal lobe (TL) or (<b>C</b>,<b>D</b>) serum of (<b>A</b>,<b>C</b>) control and (<b>B</b>,<b>D</b>) chronic ethanol-fed Long Evans rats. (<b>E</b>–<b>G</b>) NTA summary results (box plots) comparing the MV/EV (<b>E</b>) mean sizes, (<b>F</b>) mode sizes, and (<b>G</b>) nanoparticle concentrations in TL and serum samples from control (Con) and ethanol-fed (EtOH) rats (n = 3/group). Inter-group comparisons were made by two-way ANOVA. The calculated significant (<span class="html-italic">p</span> ≤ 0.05) and statistical trend-wise (0.05 &lt; <span class="html-italic">p</span> &lt; 0.10; italics) differences by post hoc Tukey tests are displayed.</p>
Full article ">Figure 2
<p>Temporal lobe membrane vesicle (TL-MV) and serum exosome (S-EV) tetraspanin immunoreactivity. ELISAs measured CD9, CD63, and CD81 immunoreactivity in (<b>A</b>) TL-MVs and (<b>B</b>) S-EVs with levels normalized to HSP70. Results from 8 control (Con) and 8 ethanol-fed (Et) rats per group are depicted with violin plots. Inter-group comparisons were made with two-way ANOVA tests (see <a href="#jmp-05-00025-t003" class="html-table">Table 3</a>). Software-calculated significant <span class="html-italic">p</span>-values (≤0.05) are displayed.</p>
Full article ">Figure 3
<p>Immature oligodendrocyte–myelin glycoproteins. ELISAs measured immunoreactivity to (<b>A</b>–<b>C</b>) CNPase and (<b>D</b>–<b>F</b>) PLP in TL-Tx homogenates, TL membrane vesicles (TL-MVs), and serum exosomes (S-EVs) from control (Con) and chronic ethanol-fed (EtOH) rats. TL immunoreactivity was normalized to RPLPO. TL-MV and S-EV immunoreactivities were normalized to HSP70. Each group included 8 samples that were analyzed in triplicate. The software-calculated significant (<span class="html-italic">p</span> ≤ 0.05) and statistical trend-wise (0.05 &lt; <span class="html-italic">p</span> &lt; 0.10) <span class="html-italic">p</span>-values from repeated measures <span class="html-italic">t</span>-tests are displayed in the panels or reported with the results.</p>
Full article ">Figure 4
<p>Pre-myelinating oligodendrocyte/myelin glycoproteins. (<b>A</b>–<b>C</b>) PDGFRA and (<b>D</b>–<b>F</b>) GALC immunoreactivities were measured by ELISA in (<b>A</b>,<b>D</b>) TL-Tx, (<b>B</b>,<b>E</b>) TL membrane vesicles (TL-MVs), and (<b>C</b>,<b>F</b>) serum exosomes (S-EVs) from control (Con) and ethanol-fed (EtOH) rats (n = 8/group). The software-calculated significant <span class="html-italic">p</span>-values (<span class="html-italic">p</span> ≤ 0.05) from repeated measures <span class="html-italic">t</span>-tests are displayed in the panels.</p>
Full article ">Figure 5
<p>Mature oligodendrocyte/myelin glycoproteins. ELISAs were used to measure (<b>A</b>–<b>C</b>) MAG, (<b>D</b>–<b>F</b>) MOG, and (<b>G</b>–<b>I</b>) MBP immunoreactivities in (<b>A</b>,<b>D</b>,<b>G</b>) TL-Tx, (<b>B</b>,<b>E</b>,<b>H</b>) TL membrane vesicles (TL-MVs), and (<b>C</b>,<b>F</b>,<b>I</b>) serum exosomes (S-EVs) from control (Con) and ethanol-fed (EtOH) rats (n = 8/group). The TL-Tx results were normalized to RPLPO, and the MV/EV results were normalized to HSP70. Statistically significant (<span class="html-italic">p</span> ≤ 0.05) results of repeated measures <span class="html-italic">t</span>-tests are depicted.</p>
Full article ">Figure 6
<p>Glial/astrocytic markers. (<b>A</b>–<b>C</b>) Nestin, (<b>D</b>–<b>F</b>) vimentin, and (<b>G</b>–<b>I</b>) GFAP immunoreactivities were measured in (<b>A</b>,<b>D</b>,<b>G</b>) TL-Tx homogenates, (<b>B</b>,<b>E</b>,<b>H</b>) TL-MVs, and (<b>C</b>,<b>F</b>,<b>I</b>) S-EV by ELISA. Samples were isolated from control (Con) and Ethanol-fed (EtOH) rats (n = 8/group) and analyzed in triplicate. TL results were normalized to RPLPO, and the MV/EV results were normalized to HSP70. Significant (<span class="html-italic">p</span> ≤ 0.05) and statistical trend-wise (0.05 &lt; <span class="html-italic">p</span> &lt; 0.10;) differences by repeated measures <span class="html-italic">t</span>-test analysis are displayed in the panels and described with the results.</p>
Full article ">Figure 7
<p>Lipid peroxidation and oxidative stress/DNA damage. ELISAs measured (<b>A</b>–<b>C</b>) HNE and (<b>D</b>–<b>F</b>) 8-OHdG immunoreactivities in (<b>A</b>,<b>D</b>) TL-Tx, (<b>B</b>,<b>E</b>) TL-MVs, and (<b>C</b>,<b>F</b>) S-EVs from control (Con) and chronic ethanol-fed (EtOH) rats (n = 8/group). Assays were performed in triplicate with TL-Tx results normalized to RPLPO, and MV/EV results normalized to HSP70. Significant (<span class="html-italic">p</span> ≤ 0.05) and statistical trend-wise (0.05 &lt; <span class="html-italic">p</span> &lt; 0.10) differences by repeated measures <span class="html-italic">t</span>-tests are displayed.</p>
Full article ">Figure 8
<p>Bidirectional heatmaps depicting overall and composite (<b>A</b>) quantitative and (<b>B</b>) qualitative effects of ethanol on white matter biomarker expression in TL-Tx, TL-MV, and S-EV. (<b>A</b>) ELISA results were analyzed by two-way ANOVA with post hoc Tukey tests. The calculated relative levels of immunoreactivity normalized to RPLPO (for TL-Tx) or HSP70 (for TL-MV and S-EV) are displayed in the boxes. Significant inter-group differences are asterisked (* <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001; **** <span class="html-italic">p</span> &lt; 0.0001), with orange font for reduced and blue for increased levels of immunoreactivity in ethanol relative to control samples. (<b>B</b>) Summary of significant ethanol-related qualitative shifts in immunoreactivity relative to control, illustrating concordant, neutral, or discordant responses across the TL-Tx, TL-MV, and S-EV samples. This qualitative heatmap was generated by scoring ethanol’s effects as follows: ±1 for <span class="html-italic">p</span> &lt; 0.0001; ±0.75 for <span class="html-italic">p</span> &lt; 0.001; ±0.05 for <span class="html-italic">p</span> &lt; 0.05. Blue reflects increased expression, orange corresponds to decreased expression, and yellow indicates no significant effect of ethanol. The heatmaps were generated with GraphPad Prism 10.2 software.</p>
Full article ">
17 pages, 2822 KiB  
Article
Preparation of Rutin–Whey Protein Pickering Emulsion and Its Effect on Zebrafish Skeletal Muscle Movement Ability
by Yiting Zhang, Wenyun Xiong, Yijing Ren, Jian Huang, Xiaoying Wang, Ou Wang and Shengbao Cai
Nutrients 2024, 16(18), 3050; https://doi.org/10.3390/nu16183050 - 10 Sep 2024
Viewed by 442
Abstract
Nutritional supplementation enriched with protein and antioxidants has been demonstrated to effectively strengthen skeletal muscle function and mitigate the risk of sarcopenia. Dietary protein has also been a common carrier to establish bioactive delivery system. Therefore, in this study, a Pickering emulsion delivery [...] Read more.
Nutritional supplementation enriched with protein and antioxidants has been demonstrated to effectively strengthen skeletal muscle function and mitigate the risk of sarcopenia. Dietary protein has also been a common carrier to establish bioactive delivery system. Therefore, in this study, a Pickering emulsion delivery system for rutin was constructed with whey protein, and its structural characteristics, bioaccessibility, and molecular interactions were investigated. In the in vivo study, zebrafish (n = 10 in each group), which have a high genetic homology to humans, were treated with dexamethasone to induce sarcopenia symptoms and were administered with rutin, whey protein and the Pickering emulsion, respectively, for muscle movement ability evaluation, and zebrafish treated with or without dexamethasone was used as the model and the control groups, respectively. Results showed that the Pickering emulsion was homogeneous in particle size with a rutin encapsulation rate of 71.16 ± 0.15% and loading efficiency of 44.48 ± 0.11%. Rutin in the Pickering emulsion exhibited a significantly higher bioaccessibility than the free form. The interaction forces between rutin and the two components of whey proteins (α-LA and β-LG) were mainly van der Waals forces and hydrogen bonds. After treatment for 96 h, the zebrafish in Picking emulsion groups showed a significantly increased high-speed movement time and frequency, an increased level of ATP, prolonged peripheral motor nerve length, and normalized muscular histological structure compared with those of the model group (p < 0.05). The results of this study developed a new strategy for rutin utilization and provide scientific evidence for sarcopenia prevention with a food-derived resource. Full article
(This article belongs to the Section Sports Nutrition)
Show Figures

Figure 1

Figure 1
<p>Characterization of rutin–whey protein Pickering emulsion (RWP). (<b>A</b>) Particle size and zeta potential. (<b>B</b>) Particle size distribution. (<b>C</b>) Embedding rate and loading rate. (<b>D</b>) Apparent images of fresh emulsions and emulsions stored for seven days (top) and confocal laser scanning microscopy images (bottom). Data are expressed as mean ± SD (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 2
<p>The first-order kinetics of RWP and rutin solution stored at 4 °C and 25 °C for 35 days. RWP: rutin–whey protein Pickering emulsion. Data are expressed as mean ± SD (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 3
<p>In vitro digestion simulation. (<b>A</b>) Retention rate. (<b>B</b>) Bioaccessibility. RWP: rutin–whey protein Pickering emulsion. Data are expressed as mean ± SD (<span class="html-italic">n</span> = 3). Different letters indicate significant differences (<span class="html-italic">p</span> &lt; 0.05). Uppercase letters in (<b>A</b>) indicate significant differences in gastric digestion and lowercase letters indicate significant differences in enteric digestion.</p>
Full article ">Figure 4
<p>Molecular docking results. (<b>A</b>) α-LA + rutin. (<b>B</b>) β-LG + rutin. α-LA: α-lactalbumin; β-LG: β-lactoglobulin; panel labels 1, 2, and 3 represent the 3D binding site map, 3D hydrogen bonding map, and 2D interaction map, respectively.</p>
Full article ">Figure 5
<p>Movement performance of zebrafish. (<b>A</b>) High-speed movement time. (<b>B</b>) High-speed movement frequency. (<b>C</b>) Total movement distance. (<b>D</b>) Trace map of the movement. A moving speed lower than 4 mm/s is marked in green; 4–20 mm/s is marked in black; and higher than 20 mm/s is marked in red. C, the control group; M, the model group; P, the positive control group; R1 and R2, the rutin treatments at 25.0 ng/fish and 50.0 ng/fish, respectively; W1 and W2, the whey protein treatments at 200.0 ng/fish and 400.0 ng/fish, respectively; RWP1 and RWP2, the rutin–whey protein Pickering emulsion treatments at 25% and 50% of the initial concentration, respectively. Data are expressed as mean ± SE (<span class="html-italic">n</span> = 10). The result of each group was compared with the M group, * indicates <span class="html-italic">p</span> &lt; 0.05 and ** indicates <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 6
<p>ATP levels in different groups. C, the control group; M, the model group; P, the positive control group; R1 and R2, the rutin treatments at 25.0 ng/fish and 50.0 ng/fish, respectively; W1 and W2, the whey protein treatments at 200.0 ng/fish and 400.0 ng/fish, respectively; RWP1 and RWP2, the rutin–whey protein Pickering emulsion treatments at 25% and 50% of the initial concentration, respectively. Data are expressed as mean ± SE (<span class="html-italic">n</span> = 10). The result of each group is compared with the M group, * indicates <span class="html-italic">p</span> &lt; 0.05 and ** indicates <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 7
<p>The peripheral motor nerve length measurement results and images (120×). (<b>A</b>) Peripheral motor nerve length. (<b>B</b>) Schematic representation of the peripheral motor nerve length analysis region. (<b>C</b>) Peripheral motor nerve length measurement areas of each treatment. C, the control group; M, the model group; P, the positive control group; R1 and R2, the rutin treatments at 25.0 ng/fish and 50.0 ng/fish, respectively; W1 and W2, the whey protein treatments at 200.0 ng/fish and 400.0 ng/fish, respectively; RWP1 and RWP2, the rutin–whey protein Pickering emulsion treatments at 25% and 50% of the initial concentration, respectively. Data are expressed as mean ± SE (<span class="html-italic">n</span> = 10). The result of each group was compared with the M group, * indicates <span class="html-italic">p</span> &lt; 0.05 and ** indicates <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 8
<p>The histopathological analysis of zebrafish skeletal muscle (400×). C, the control group; M, the model group; P, the positive control group; R1 and R2, the rutin treatments at 25.0 ng/fish and 50.0 ng/fish, respectively; W1 and W2, the whey protein treatments at 200.0 ng/fish and 400.0 ng/fish, respectively; RWP1 and RWP2, the rutin–whey protein Pickering emulsion treatments at 25% and 50% of the initial concentration, respectively.</p>
Full article ">
11 pages, 4375 KiB  
Case Report
A 13-Year-Old Girl Affected by Melanocytic Tumors of the Central Nervous System—The Case
by Emilia Nowosławska, Magdalena Zakrzewska, Beata Sikorska, Jakub Zakrzewski and Bartosz Polis
Int. J. Mol. Sci. 2024, 25(17), 9628; https://doi.org/10.3390/ijms25179628 - 5 Sep 2024
Viewed by 293
Abstract
Primary intracranial melanoma is a very rare brain tumor, especially when accompanied by benign intramedullary melanocytoma. Distinguishing between a primary central nervous system (CNS) lesion and metastatic melanoma is extremely difficult, especially when the primary cutaneous lesion is not visible. Here we report [...] Read more.
Primary intracranial melanoma is a very rare brain tumor, especially when accompanied by benign intramedullary melanocytoma. Distinguishing between a primary central nervous system (CNS) lesion and metastatic melanoma is extremely difficult, especially when the primary cutaneous lesion is not visible. Here we report a 13-year-old girl admitted to the Neurosurgery Department of the Institute of Polish Mother’s Health Centre in Lodz due to upper limb paresis. An intramedullary tumor of the cervical C3–C4 and an accompanying syringomyelic cavity C1–C7 were revealed. The child underwent partial removal of the tumor due to the risk of damage to spinal cord motor centers. The removed part of the tumor was diagnosed as melanocytoma. Eight months later, a neurological examination revealed paresis of the right sixth cranial nerve, accompanied by bilateral optic disc edema. Diagnostic imaging revealed a brain tumor. The girl underwent resection of both detected the tumors and an additional satellite lesion revealed during the surgery. The removed tumors were diagnosed as malignant melanomas in pathomorphological examination. Molecular analysis revealed NRASQ61K mutation in both the intracranial and the intramedullary tumor. It should be noted that in cases where available evidence is inconclusive, an integrative diagnostic process is essential to reach a definitive diagnosis. Full article
(This article belongs to the Section Molecular Neurobiology)
Show Figures

Figure 1

Figure 1
<p>Contrast-enhanced T1 MRI of the patient’s cervical spine. (<b>A</b>) The axial section at the C2 level; (<b>B</b>) the sagittal section. The photos were taken before the procedure. An intramedullary tumor is located in the spinal canal between the C2–C4 levels (white arrow), accompanied by a syringomyelic cavity extending throughout the spinal cord from the hindbrain to the cervical spine at the C7 level (black arrow). Due to the dynamic neurological symptoms, characteristic of the expansive process in the spinal cord, and the need for urgent surgery, imaging diagnostics were limited to the spinal cord and the cervicocranial border. An MRI of the brain was performed at a later stage.</p>
Full article ">Figure 2
<p>Photos of the patient’s skin pigment spots: (<b>A</b>) a skin mole on the neck removed after dermatoscopy (melanoma was excluded at the time); (<b>B</b>) the remaining part of the skin moles on the neck; (<b>C</b>) birthmarks on the skin of the back; (<b>D</b>) a hairy birthmark on the left thigh; (<b>E</b>) a nodule birthmark on the skin of the right thigh removed after dermatoscopy (melanoma was excluded at the time).</p>
Full article ">Figure 3
<p>Imaging diagnostics of the patient’s cervical spine using T1-weighted MRI with contrast enhancement 48 h after the procedure: (<b>A</b>) an axial section of the cervical spine at the C2 level with clearly visible reduction of the syringomyelic cavity (arrow); (<b>B</b>) a sagittal scan of the cervical spine; the remaining part of the tumor mass in the cervical spine is visible as a hyperintense focus in contrast to the image of the spinal cord (arrow).</p>
Full article ">Figure 4
<p>Microscopic images of the cervical spinal cord tumor: (<b>A</b>) H&amp;E staining shows oval and spindled cells containing pigment (white arrows), magnification 200×; (<b>B</b>) Ki67 immunohistochemical staining showing a low proliferation index of the tumor cells (&lt;5% of positive cells), magnification 200×.</p>
Full article ">Figure 5
<p>Imaging of the brain tumor: (<b>A</b>) T1-weighted MRI of the patient’s head without contrast enhancement—the brain tumor is visible as a hypointense lesion in contrast to the white matter, located in the right frontal lobe; (<b>B</b>) contrast-enhanced T1-weighted MRI—the mass in the right frontal region shows intense contrast enhancement; (<b>C</b>) T2-weighted head MRI—the tumor in the right frontal area is visible as a hyperintense lesion in contrast to the white matter; (<b>D</b>) CT diagnostic imaging after contrast enhancement—the tumor in the right frontal region is visible as a hyperdense lesion in contrast to the white matter of the brain. The arrows indicate the tumor location.</p>
Full article ">Figure 6
<p>Diagnostic T1-weighted MRI images of the patient’s cervical spine: (<b>A</b>) An axial section at the C2 level of the patient’s cervical spine made after administration of a contrast agent. The remainder of the tumor showed no apparent progression compared to the MRI performed 48 h after the first treatment (arrow). (<b>B</b>) A sagittal section of the patient’s cervical spine (without contrast). The photo shows decompression of the syringomyelic cavity (black arrows) and the remains of the partially removed tumor, visible as an isointense lesion in the spine (white arrow). The reserve of cerebrospinal fluid in the spinal canal is clearly increased (red arrow). The photo shows no signs of tumor progression.</p>
Full article ">Figure 7
<p>An intraoperative view of the surgical field. The right frontal lobe of the brain is exposed through a right frontal craniotomy: (<b>A</b>) fresh blood in the tumor mass located above the subarachnoid black satellite lesion; (<b>B</b>) the main focus of the tumor after removal of the blood clot, visible above the dark subarachnoid lesion.</p>
Full article ">Figure 8
<p>Microscopic images of the intracranial tumor: (<b>A</b>) H&amp;E staining shows pleomorphic epithelioid cells, magnification 400×; (<b>B</b>) immunoexpression of MelanA in the tumor cells, magnification 200×; (<b>C</b>) Ki67 immunohistochemical staining showing a high proliferation index of the tumor cells (&gt;30% of positive cells), magnification 200×; (<b>D</b>,<b>E</b>) transmission electron microscopy images of the melanotic tumor cells (arrows), magnification 25,000×.</p>
Full article ">Figure 9
<p>Two-dimensional plots of ddPCR analysis of <span class="html-italic">NRASQ61K</span> mutation: (<b>A</b>) intramedullary tumor sample, ratio = 1.49; (<b>B</b>) brain tumor sample, ratio = 0.688; (<b>C</b>) fascial sample, ratio = 0.687. Purple lines indicate thresholds for distinguishing positive (blue cluster), positive/wild-type (orange cluster), negative (gray cluster), and wild-type (green cluster) signals. The mutation frequency of genes was quantified by the ratio of mutant droplets (only HEX positive) to wild-type droplets (HEX/FAM double positive).</p>
Full article ">Figure 10
<p>Contrast-enhanced T1 MRI images of the head and cervical spine of the 13-year-old girl before the valve implantation: (<b>A</b>) an axial section of the head. The photo does not show progression of the tumor (arrow). (<b>B</b>) MRI of the cervical spine—an axial section. Visible remnants of the tumor diagnosed as melanocytoma (arrow). (<b>C</b>) An MRI—sagittal section of the cervical spine. The residual tumor mass shows no enlargement compared to previous diagnostic images (arrow).</p>
Full article ">
18 pages, 29602 KiB  
Article
Slc4a7 Regulates Retina Development in Zebrafish
by Youyuan Zhuang, Dandan Li, Cheng Tang, Xinyi Zhao, Ruting Wang, Di Tao, Xiufeng Huang and Xinting Liu
Int. J. Mol. Sci. 2024, 25(17), 9613; https://doi.org/10.3390/ijms25179613 - 5 Sep 2024
Viewed by 370
Abstract
Inherited retinal degenerations (IRDs) are a group of genetic disorders characterized by the progressive degeneration of retinal cells, leading to irreversible vision loss. SLC4A7 has emerged as a candidate gene associated with IRDs, yet its mechanisms remain largely unknown. This study aims to [...] Read more.
Inherited retinal degenerations (IRDs) are a group of genetic disorders characterized by the progressive degeneration of retinal cells, leading to irreversible vision loss. SLC4A7 has emerged as a candidate gene associated with IRDs, yet its mechanisms remain largely unknown. This study aims to investigate the role of slc4a7 in retinal development and its associated molecular pathogenesis in zebrafish. Morpholino oligonucleotide knockdown, CRISPR/Cas9 genome editing, quantitative RT-PCR, eye morphometric measurements, immunofluorescent staining, TUNEL assays, visual motor responses, optokinetic responses, rescue experiments, and bulk RNA sequencing were used to assess the impact of slc4a7 deficiency on retinal development. Our results demonstrated that the knockdown of slc4a7 resulted in a dose-dependent reduction in eye axial length, ocular area, and eye-to-body-length ratio. The fluorescence observations showed a significant decrease in immunofluorescence signals from photoreceptors and in mCherry fluorescence from RPE in slc4a7-silenced morphants. TUNEL staining uncovered the extensive apoptosis of retinal cells induced by slc4a7 knockdown. Visual behaviors were significantly impaired in the slc4a7-deficient larvae. GO and KEGG pathway analyses reveal that differentially expressed genes are predominantly linked to aspects of vision, ion channels, and phototransduction. This study demonstrates that the loss of slc4a7 in larvae led to profound visual impairments, providing additional insights into the genetic mechanisms predisposing individuals to IRDs caused by SLC4A7 deficiency. Full article
(This article belongs to the Section Molecular Genetics and Genomics)
Show Figures

Figure 1

Figure 1
<p>The expression profile of <span class="html-italic">SLC4A7 (slc4a7</span>) in both humans and zebrafish. (<b>A</b>) Profiling of human single-cell RNA sequencing unveils <span class="html-italic">SLC4A7</span> expression in various ocular tissues, including the retina, cornea, iris, sclera, and choroid. (<b>B</b>) Single-cell transcriptome profiling and <span class="html-italic">SLC4A7</span> gene signatures of the human retina. The upper tSNE plot showing major cell subsets in human retina. The lower tSNE plot of all cells colored by enrichment of <span class="html-italic">SLC4A7</span> gene signatures. (<b>C</b>) Detailed expression pattern of <span class="html-italic">SLC4A7</span> across various cell types in the human retina. (<b>D</b>) Human subcellular localization of <span class="html-italic">SLC4A7</span> provided by COMPARTMENTS. (<b>E</b>,<b>F</b>) qRT-PCR display the time series (<b>E</b>) and the tissue-specific (<b>F</b>) expression pattern of <span class="html-italic">slc4a7</span> in zebrafish larvae. Bar plots are shown as the mean ± std. Label of human and zebrafish indicating the species source of the data. The normalized expression values refer to the standardized measure of gene expression levels across different cells or cell types in a single-cell RNA sequencing dataset, as calculated and provided by the online analysis tool, Single Cell Portal. The relative expression values were calculated relative to the expression of the reference gene β-actin. The dashed lines indicate detailed information about the corresponding tissues or figures. Human or zebrafish icons indicate the species origin of the data.</p>
Full article ">Figure 2
<p><span class="html-italic">Slc4a7</span>-deficient zebrafish morphants exhibited marked ocular changes. (<b>A</b>) Whole-body view (scale bar = 500 μm), lateral view (scale bar = 100 μm), and vertical view (scale bar = 100 μm) of zebrafish larvae at 3 dpf. The ocular area is indicated by white circles. The ocular axis is indicated by a red dashed line. The equatorial axis is indicated by a blue dashed line. (<b>B</b>–<b>E</b>) The measurement of ocular axis length, equatorial length, ocular area, and the ratio of ocular length to body length. <span class="html-italic">n</span> = 25 in each group. Bar plots are shown as the mean ± s.e.m. Data were analyzed using one-way ANOVA followed by Tukey’s post hoc tests, *** <span class="html-italic">p</span> &lt; 0.001, and **** <span class="html-italic">p</span> &lt; 0.0001 indicate significant differences from the control 1.00 ng group. The scattered circles, squares, and triangles of various colors in the bar chart denote individual values of zebrafish eye-related parameters for different groups.</p>
Full article ">Figure 2 Cont.
<p><span class="html-italic">Slc4a7</span>-deficient zebrafish morphants exhibited marked ocular changes. (<b>A</b>) Whole-body view (scale bar = 500 μm), lateral view (scale bar = 100 μm), and vertical view (scale bar = 100 μm) of zebrafish larvae at 3 dpf. The ocular area is indicated by white circles. The ocular axis is indicated by a red dashed line. The equatorial axis is indicated by a blue dashed line. (<b>B</b>–<b>E</b>) The measurement of ocular axis length, equatorial length, ocular area, and the ratio of ocular length to body length. <span class="html-italic">n</span> = 25 in each group. Bar plots are shown as the mean ± s.e.m. Data were analyzed using one-way ANOVA followed by Tukey’s post hoc tests, *** <span class="html-italic">p</span> &lt; 0.001, and **** <span class="html-italic">p</span> &lt; 0.0001 indicate significant differences from the control 1.00 ng group. The scattered circles, squares, and triangles of various colors in the bar chart denote individual values of zebrafish eye-related parameters for different groups.</p>
Full article ">Figure 3
<p>Loss of <span class="html-italic">slc4a7</span> leads to retinal abnormalities in zebrafish larvae. (<b>A</b>) Immunostaining for RCVRN in <span class="html-italic">slc4a7</span>-deficient and control AB zebrafish strains at 5 dpf. (<b>B</b>) The fluorescence imaging depicts retinal pigment epithelium cells and amacrine cells in <span class="html-italic">slc4a7</span>-deficient and control Tg (gad1b:mCherry) strain at 5 dpf. (<b>C</b>) The fluorescence imaging focusing on blood vessel endothelial cells in <span class="html-italic">slc4a7</span>-deficient and control Tg(kdrl:mCherry) strains at 5 dpf. (<b>D</b>) The fluorescence imaging illustrates Müller cells in <span class="html-italic">slc4a7</span>-deficient and control Tg (gfap:egfp) strains at 5 dpf. (<b>E</b>) Statistical results for normalized fluorescence intensity of RCVRN. (<b>F</b>,<b>G</b>) Statistical results for normalized fluorescence intensity of mCherry signal in amacrine cells (ACs), (<b>F</b>), and the RPE layer (<b>G</b>). (<b>H</b>) Statistical results for vascular network thickness surrounding photoreceptors. (<b>I</b>) Statistical results for the normalized fluorescence intensity of the eGFP signal in Müller cell layer. Scale bar = 50 μm. Bar plots are shown as the mean ± s.e.m. The <span class="html-italic">t</span>-test was performed between the two groups. * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001. GCL: ganglion cell layer; IPL: inner plexiform layer; INL: inner nuclear layer; OPL: outer plexiform layer; ONL: outer nuclear layer.</p>
Full article ">Figure 4
<p>Knockdown of <span class="html-italic">slc4a7</span> resulted in a marked elevation of apoptosis in the zebrafish retina. TUNEL assay was used to detect apoptosis in larval retinas at 5 dpf. Knockdown of <span class="html-italic">slc4a7</span> led to a marked elevation in the number of TUNEL<sup>+</sup> cells in the retina at 5 days dpf. Scale bars = 20 μm. Blue indicates cell nucleus stained by DAPI. Red indicates cells with a positive TUNEL reaction.</p>
Full article ">Figure 5
<p>Zebrafish morphants lacking <span class="html-italic">slc4a7</span> exhibited impaired visual behaviors. (<b>A</b>,<b>B</b>) VMR testing results are depicted in the line charts for <span class="html-italic">slc4a7</span>-deficient zebrafish morphants. (<b>C</b>,<b>D</b>) Scatter plots with bar show the quantification of ON responses and OFF responses. Each point represents the average activity of 12 larvae within the group at the moment of each light environment change. Error bars represent standard deviation (STD). (<b>E</b>,<b>F</b>) Frequency distributions of eye movements (times/minute) in zebrafish larvae are illustrated in the stacked boxes. Data were analyzed using one-way ANOVA with Tukey’s post hoc tests, * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001. The scattered circles of various colors in the bar chart denote individual values of zebrafish eye-related parameters for different groups.</p>
Full article ">Figure 6
<p>Transcriptional profiling of bulbus oculi from <span class="html-italic">slc4a7</span>-deficient zebrafish morphants. (<b>A</b>) Principal component analysis (PCA) of the expressed genes showing <span class="html-italic">slc4a7</span> and control sample separation. (<b>B</b>) Heatmap of sample-to-sample distances. (<b>C</b>) Volcano plot showing highlights of DEGs from <span class="html-italic">slc4a7</span> KD eyes compared with control eyes. (<b>D</b>) GO analysis identified the top 30 most significant GO terms in the <span class="html-italic">slc4a7</span> 1.00 ng group compared to the control 1.00 ng group. (<b>E</b>) Significantly enriched KEGG pathways (top 20) in the <span class="html-italic">slc4a7</span> 1.00 ng group compared to the control 1.00 ng group.</p>
Full article ">Figure 7
<p><span class="html-italic">Slc4a7</span> mRNA compensation rescues phenotypes in slc4a7-deficient morphants. (<b>A</b>) Enlarged vertical and lateral views of larval eyeballs. Larvae injected with slc4a7 MO and full-length slc4a7 mRNA exhibit normal-sized eyeballs at 3 dpf. Scale bar = 100 mm. (<b>B</b>–<b>D</b>) Statistical analysis of axial length and ocular area. (<b>E</b>,<b>F</b>) VMR testing demonstrates significant recovery of both ON and OFF responses in slc4a7-deficient zebrafish compensated with mRNA compared to those without compensation. <span class="html-italic">n</span> = 20 in each group. Rescue experiments were repeated three times. Bar plots represent the mean ± s.e.m. Data were analyzed using one-way ANOVA followed by Tukey’s post hoc test, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001, indicate significant differences from the <span class="html-italic">slc4a7</span> 1.00 ng group. ns means no significance. The scattered circles of various colors in the bar chart denote individual values of zebrafish eye-related parameters for different groups.</p>
Full article ">
28 pages, 8501 KiB  
Article
The Bifunctional Dimer Caffeine-Indan Attenuates α-Synuclein Misfolding, Neurodegeneration and Behavioral Deficits after Chronic Stimulation of Adenosine A1 Receptors
by Elisabet Jakova, Omozojie P. Aigbogun, Mohamed Taha Moutaoufik, Kevin J. H. Allen, Omer Munir, Devin Brown, Changiz Taghibiglou, Mohan Babu, Chris P. Phenix, Ed S. Krol and Francisco S. Cayabyab
Int. J. Mol. Sci. 2024, 25(17), 9386; https://doi.org/10.3390/ijms25179386 - 29 Aug 2024
Viewed by 490
Abstract
We previously found that chronic adenosine A1 receptor stimulation with N6-Cyclopentyladenosine increased α-synuclein misfolding and neurodegeneration in a novel α-synucleinopathy model, a hallmark of Parkinson’s disease. Here, we aimed to synthesize a dimer caffeine-indan linked by a 6-carbon chain to cross [...] Read more.
We previously found that chronic adenosine A1 receptor stimulation with N6-Cyclopentyladenosine increased α-synuclein misfolding and neurodegeneration in a novel α-synucleinopathy model, a hallmark of Parkinson’s disease. Here, we aimed to synthesize a dimer caffeine-indan linked by a 6-carbon chain to cross the blood–brain barrier and tested its ability to bind α-synuclein, reducing misfolding, behavioral abnormalities, and neurodegeneration in our rodent model. Behavioral tests and histological stains assessed neuroprotective effects of the dimer compound. A rapid synthesis of the 18F-labeled analogue enabled Positron Emission Tomography and Computed Tomography imaging for biodistribution measurement. Molecular docking analysis showed that the dimer binds to α-synuclein N- and C-termini and the non-amyloid-β-component (NAC) domain, similar to 1-aminoindan, and this binding promotes a neuroprotective α-synuclein “loop” conformation. The dimer also binds to the orthosteric binding site for adenosine within the adenosine A1 receptor. Immunohistochemistry and confocal imaging showed the dimer abolished α-synuclein upregulation and aggregation in the substantia nigra and hippocampus, and the dimer mitigated cognitive deficits, anxiety, despair, and motor abnormalities. The 18F-labeled dimer remained stable post-injection and distributed in various organs, notably in the brain, suggesting its potential as a Positron Emission Tomography tracer for α-synuclein and adenosine A1 receptor in Parkinson’s disease therapy. Full article
Show Figures

Figure 1

Figure 1
<p>Radiosynthesis of the <sup>18</sup>F-C<sub>8</sub>–6–I dimer and in vivo studies with the CD-1 mice. (<b>a</b>) Formation reaction of <sup>18</sup>F-C<sub>8</sub>–6–I from C<sub>8</sub>–6–I–OMs in 23 ± 5% rcy (decay corrected). (<b>b</b>) All steps involved in the (<b>b1</b>) radiosynthesis of <sup>18</sup>F-C<sub>8</sub>–6–I from [<sup>18</sup>F]fluorine purification, (<b>b2</b>) nucleophilic reaction with Kryptofix/K<sup>18</sup>F and (<b>b3</b>) semi-preparative HPLC purification, to (<b>b4</b>) PET-imaging and (<b>b5</b>) biodistribution of the <sup>18</sup>F- C<sub>8</sub>–6–I in major organs. (<b>b6</b>) The data were then analyzed and graphed using GraphPad Prism 8 (San Diego, CA, USA). Created using <a href="http://BioRender.com" target="_blank">BioRender.com</a> (URL accessed on 8 July 2024).</p>
Full article ">Figure 2
<p>Behavioral tests conducted for male Sprague-Dawley rats after the 7-day chronic injection of C<sub>8</sub>–6–I at 3 and 5 mg/kg. (<b>a</b>). Y-maze test values of the 7-day chronic C<sub>8</sub>–6–I dimer (3 mg/kg and 5 mg/kg) as percentage of time spent in each of the arms: S-arm (“start” arm), O-arm (“old” arm), and N-arm (“new’ arm). The percentages of the time spent in each arm were calculated from the 5 min trial. Open field test values of the 7-day chronic C<sub>8</sub>–6–I dimer (3 mg/kg and 5 mg/kg) as percentage of time spent in the center square. The animal is placed in the center square of the grid and left free to explore the field for 10 min. (<b>b</b>). The percentage of the time spent in the red center square. (<b>c</b>). The total fecal boli count. Forced swim test results of the 7-day chronic C<sub>8</sub>–6–I dimer (3 mg/kg and 5 mg/kg) treatment groups as measurements of swimming vigorously and successfully. The animals were placed in the forced swim tank and let free to swim for 10 min. Once the test was conducted the animals were scored for (<b>d</b>) vigor, the ability to purposely swim and use all limbs and (<b>e</b>) success, the ability to keep their head above water. (<b>f</b>). The total time spent immobile was also measured to assess learned helplessness and despair. Each dimer treatment was repeated at least 10 times per treatment (<span class="html-italic">n</span> = 10) and the average of each treatment is presented in bar graphs as means ± SEM. Significances were determined using One-way ANOVA, followed by Student–Newman–Keuls multiple comparison tests with * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; and *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 3
<p>Molecular docking simulation of α-Syn structures ((<b>a</b>). C1, (<b>b</b>). C2, (<b>c</b>). C3, and (<b>d</b>). C8) bound to C<sub>8</sub>–6–I. Bold black dashed lines and amino acid residues indicate hydrogen bonding, while the grey dashed lines and amino acid residues indicate hydrophobic interactions. (<b>a</b>) The bifunctional dimer compound forms a hydrogen bond with T81 located in the NAC region of α-Syn, and additional hydrophobic interactions are found with the α-Syn N-terminus and NAC region. (<b>b</b>) The dimer compound interacts via hydrogen bonding to the α-Syn N-terminus, and additional hydrophobic interactions occur with amino acid residues in the N-terminus and NAC region. (<b>c</b>) The dimer compound binds via hydrogen bonds to amino acid residues in the α-Syn N-terminus and NAC region, and additional hydrophobic binding occurs with amino acids located in the distal N-terminus and NAC region. (<b>d</b>) The dimer compound binds via hydrogen bonding with NAC amino acid residues (V66, T72) and through hydrophobic interactions with amino acid residues in the N-terminus and NAC domain. C1, C2, C3 and C8 α-Syn structures bind to the dimer compound, which is predicted to form a “loop” conformation of α-Syn.</p>
Full article ">Figure 4
<p>Molecular docking of C<sub>8</sub>–6–I with A1R and A2AR. (<b>a</b>) Amino acid sequence alignment of human A1R and A2AR with distinct binding of C<sub>8</sub>–6–I to A1R and A2AR indicated (red asterisks, A1R binding; blue asterisks, A2AR binding). Amino acid residues shaded in red are conserved or identical amino acid sequences, while amino acids in red font are mostly classified under non-polar aliphatic residues (AVLIM). Other amino acids highlighted in red font are classified as follows: HKR are polar positive; DE are polar negative; STNQ are polar neutral; FYW are nonpolar aromatic. (<b>b</b>) Molecular docking of C<sub>8</sub>–6–I with A1R showing binding to amino acid residues that are similarly found within the A1R orthosteric binding site for adenosine. (<b>c</b>) Molecular docking showing C<sub>8</sub>–6–I binding to amino acid residues that do not resemble those associated with A2AR orthosteric binding site.</p>
Full article ">Figure 5
<p>Summary of the surface area analysis of the dimer study’s pars compacta region of the substantia nigra of DAPI, TH, and α-Syn. (<b>a</b>) Representative images of 40 μm pars compacta region of substantia nigra taken with 63X oil immersion objective of a confocal microscope (126 times magnification). Separate channels of 7-day chronic intraperitoneal injections with 3 mg/kg of the following treatments: Control (DMSO/Saline), CPA, C<sub>8</sub>–6–I (3 mg/kg) + CPA, and C<sub>8</sub>–6–I (5 mg/kg) + CPA. Slices were probed for DAPI (Blue), Thioflavin S (Thio-S, green), and α-Syn (Red, Alexa Fluor 647). Arrows indicate neuronal somas and processes with high localization of aggregated α-Syn. Scale 50 μm. (<b>b</b>) Bar charts showing the mean area intensities of α-Syn and Thioflavin S in the pars compacta region of the substantia nigra. Similar areas of 100-by-100 μm ROI coordinates for lateral pars compacta of SN were quantified, respectively, for each slice and normalized by subtracting F0 (50 by 50 μm ROI coordinates) values of the background (non-cell body bottom area). The average intensity values in bars represent the average mean ± SEM from <span class="html-italic">n</span> = 4 independent experiments. Significances were determined using One-way ANOVA, followed by Student–Newman–Keuls multiple comparison tests with ** <span class="html-italic">p</span> &lt; 0.01; and *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 6
<p>Summary of the surface area analysis of the CA1 region of the hippocampus of DAPI, α-Syn and Thio-S. (<b>a</b>) Representative images from 40 μm hippocampal rat brain slices after probing for DAPI, anti-α-Synuclein and Thioflavin S (Thio-S) taken at 63-times magnification with a confocal microscope for the following treatments: Control (DMSO/Saline), CPA, C<sub>8</sub>–6–I (3 mg/kg) + CPA, and C<sub>8</sub>–6–I (5 mg/kg) + CPA. Arrows indicate high colocalization of Thioflavin S and α-Syn in CA1 hippocampal neuronal somas and dendritic processes. Scale 50 μm. (<b>b</b>) Bar charts showing the mean area intensities of α-Syn and Thioflavin S in the CA1 region of the hippocampus. Fluorescence intensities from a 100-by-100 μm ROI from the CA1 pyramidal cell layer of the hippocampus were quantified using a similar method to that employed for the pars compacta region. The average intensity values in bars represent the average mean ± SEM from <span class="html-italic">n</span> = 4 independent experiments. Significances were determined using One-way ANOVA, followed by Student–Newman–Keuls multiple comparison test with * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; and *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 7
<p>Fluoro-Jade C (FJC) staining in the SN pars compacta and in the hippocampus CA1 region of rats with 7-day chronic intraperitoneal injection of Control (DMSO/saline), CPA, C<sub>8</sub>–6–I (3 mg/kg) + CPA, and C<sub>8</sub>–6–I (5 mg/kg) + CPA. Representative images with 50 μm scale bar for the (<b>a</b>) SN pars compacta and (<b>b</b>) the CA1 region of the hippocampus. FJC fluorescence intensity in a 100 × 100 μm<sup>2</sup> region was normalized to the control group (100%). Values are shown as mean ± SEM. The average FJC fluorescence values were obtained from <span class="html-italic">n</span> = 4 independent experiments. * <span class="html-italic">p</span> &lt; 0.05; and ** <span class="html-italic">p</span> &lt; 0.01 (one-way ANOVA followed by Student–Newman–Keuls post-hoc multiple comparison test).</p>
Full article ">Figure 8
<p>Ex vivo stability and biodistribution in CD-1 mice at five different time points (5, 10, 20, 40, and 60 min) and major organs. (<b>a</b>) HPLC co-registration profiles of <sup>18</sup>F–C<sub>8</sub>–6–I and <sup>19</sup>F–C<sub>8</sub>–6–I based on a radio detector and ultraviolet detector, respectively. Analytical radio-HPLC chromatograms in mouse (<b>b</b>) liver and (<b>c</b>) pancreas extracts at 40 min after injection of <sup>18</sup>F–C<sub>8</sub>–6–I. (<b>d</b>) The distribution of <sup>18</sup>F–C<sub>8</sub>–6–I was calculated as a percentage of the injected dose per gram of tissue (% ID/g) and the results are displayed into two groups: lower distribution—blood, heart, bone, and brain (left panel), and higher distribution—liver, duodenum, kidneys, spleen, lungs, and large intestine (right panel). The data were obtained from <span class="html-italic">n</span> = 3 independent animals. Values are presented as mean ± SEM. Significances are indicated as follows: * <span class="html-italic">p</span> &lt; 0.05; and *** <span class="html-italic">p</span> &lt; 0.001 (one-way ANOVA followed by Student–Newman–Keuls post-hoc multiple comparison test).</p>
Full article ">Figure 9
<p>Representative PET/CT images in CD-1 mice at different time points throughout an hour, as well as time–activity curve in the brain. (<b>a</b>) PET images at different time points (1 min, 5 min, 10 min, 20 min, 30 min, and 50 min). (<b>b</b>) PET summation images for 60 min and (<b>c</b>) 120 min dynamic imaging. (<b>d</b>) Time–activity curve (TAC) of <sup>18</sup>F–C<sub>8</sub>–6–I for whole brain and three regions consisting of the cortex, midbrain, and hippocampus from PET/CT imaging. Values are presented as the standardized uptake value (SUV). The data were obtained from <span class="html-italic">n</span> = 3 independent animals.</p>
Full article ">
21 pages, 6758 KiB  
Article
NeuroAiDTM-II (MLC901) Promoted Neurogenesis by Activating the PI3K/AKT/GSK-3β Signaling Pathway in Rat Spinal Cord Injury Models
by Anam Anjum, Muhammad Dain Yazid, Muhammad Fauzi Daud, Jalilah Idris, Angela Min Hwei Ng, Amaramalar Selvi Naicker, Ohnmar Htwe Rashidah Ismail, Ramesh Kumar Athi Kumar and Yogeswaran Lokanathan
Biomedicines 2024, 12(8), 1920; https://doi.org/10.3390/biomedicines12081920 - 21 Aug 2024
Viewed by 470
Abstract
Traumatic damage to the spinal cord (SCI) frequently leads to irreversible neurological deficits, which may be related to apoptotic neurodegeneration in nerve tissue. The MLC901 treatment possesses neuroprotective and neuroregenerative activity. This study aimed to explore the regenerative potential of MLC901 and the [...] Read more.
Traumatic damage to the spinal cord (SCI) frequently leads to irreversible neurological deficits, which may be related to apoptotic neurodegeneration in nerve tissue. The MLC901 treatment possesses neuroprotective and neuroregenerative activity. This study aimed to explore the regenerative potential of MLC901 and the molecular mechanisms promoting neurogenesis and functional recovery after SCI in rats. A calibrated forceps compression injury for 15 s was used to induce SCI in rats, followed by an examination of the impacts of MLC901 on functional recovery. The Basso, Beattie, and Bresnahan (BBB) scores were utilized to assess neuronal functional recovery; H&E and immunohistochemistry (IHC) staining were also used to observe pathological changes in the lesion area. Somatosensory Evoked Potentials (SEPs) were measured using the Nicolet® Viking Quest™ apparatus. Additionally, we employed the Western blot assay to identify PI3K/AKT/GSK-3β pathway-related proteins and to assess the levels of GAP-43 and GFAP through immunohistochemistry staining. The study findings revealed that MLC901 improved hind-limb motor function recovery, alleviating the pathological damage induced by SCI. Moreover, MLC901 significantly enhanced locomotor activity, SEPs waveform, latency, amplitude, and nerve conduction velocity. The treatment also promoted GAP-43 expression and reduced reactive astrocytes (GFAP). MLC901 treatment activated p-AKT reduced p-GSK-3β expression levels and showed a normalized ratio (fold changes) relative to β-tubulin. Specifically, p-AKT exhibited a 4-fold increase, while p-GSK-3β showed a 2-fold decrease in T rats compared to UT rats. In conclusion, these results suggest that the treatment mitigates pathological tissue damage and effectively improves neural functional recovery following SCI, primarily by alleviating apoptosis and promoting neurogenesis. The underlying molecular mechanism of this treatment mainly involves the activation of the PI3K/AKT/GSK-3β pathway. Full article
(This article belongs to the Special Issue Spinal Cord Compression: Molecular, Cellular and Therapeutic Aspects)
Show Figures

Figure 1

Figure 1
<p>Diagrammatic representation of the step-by-step procedure for creating a mechanical spinal cord injury (SCI) model using the calibrated forceps compression method. (<b>a</b>) Sublime Animal Position: The animal is positioned in a prone orientation on the surgical table, ensuring stability and access to the spinal region. (<b>b</b>) Marking T10, T12, and T13 vertebra: Identification of the vertebrae to be targeted and marked for precise surgical intervention. (<b>c</b>) subcutaneous cut: An incision is made through the skin to gain access to the underlying tissues. (<b>d</b>) Removing Muscles: The overlying muscles are carefully removed to expose the spinal column while minimizing damage to surrounding tissues. (<b>e</b>) Exposing Spinal Cord: The spinal column is accessed, providing visibility to the spinal cord. (<b>f</b>) Removing T12 Vertebrae: The T12 vertebra is removed to allow direct compression of the spinal cord. (<b>g</b>) Compression of the Spinal Cord for 15 Sec: The spinal cord is compressed using calibrated forceps for a precise duration to induce injury, the blue box indicates spinal cord exposure and the compression site. (<b>h</b>) Wound Closing (Suture of Tissue and Skin): The surgical wound is closed with sutures, including tissue layers and skin, to complete the procedure, <span class="html-italic">n</span> = 6: Indicates the number of animals used for this procedure.</p>
Full article ">Figure 2
<p>Health evaluation after compression injury. (<b>a</b>) Changes in body weight following spinal cord injury: Injured rats initially lost more body weight on Days 3 and 7 than sham (H) rats, but later began gaining weight. Treated (T) rats lost less weight compared to untreated (UT) animals on Days 3 and 7 (mean ± standard deviation [SD]; one-way analysis of variance [ANOVA] with post hoc test, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01) (<span class="html-italic">n</span> = 6). (<b>b</b>) Urine volume per void: Manual bladder voiding by the experimenter was required for 1 week (Day 7) post-SCI, after which spontaneous recovery began. No significant difference was observed between UT and T rats. Data were analyzed using Student’s <span class="html-italic">t</span>-test for non-parametric and unpaired comparisons (<span class="html-italic">n</span> = 6). (<b>c</b>) Length of the incision area: The length of the incision area was measured up to Day 28 post-injury. Wound recovery continued until Day 28, with no significant difference observed between UT and T rats. Data were analyzed using Student’s <span class="html-italic">t</span>-test for unpaired comparisons (<span class="html-italic">n</span> = 6).</p>
Full article ">Figure 3
<p>Assessment of locomotor function following MLC901 treatment (<b>a</b>) Basso, Beattie, and Bresnahan (BBB) scores, (<b>b</b>) distance traveled in the open field test (OFT), (<b>c</b>) performance in the running wheel test, (<b>d</b>) grid walk assessment, (<b>e</b>) grid distance traveled, (<b>f</b>) inverted grid (grip strength test), (<b>g</b>) total number of footsteps taken, and (<b>h</b>) fore- and hind-limb faults. Assessments were conducted on Day 0 (pre-injury) and Days 3, 7, 14, 21, and 28 post-injuries. Data are presented as means with error bars indicating the standard error of the mean (<span class="html-italic">n</span> = 6). Statistical analysis was performed using one-way ANOVA followed by post hoc tests. Treated (T) rats demonstrated significant recovery and improvement compared to controls. Statistical significance is indicated as * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001 for comparisons among the three groups (sham [H], treated [T], and untreated [UT]).</p>
Full article ">Figure 4
<p>Somatosensory Evoked Potential (SEP) Analysis. (<b>a</b>) Waveform: SEPs were obtained by stimulating the left hindlimb in sham (H), untreated (UT), and treated (T) rats with moderate compression injury on Day 0 (pre-injury), Day 14, and Day 28 post-injury. The waveform data show the characteristic peaks and changes over time. (<b>b</b>) Amplitude: Amplitude of SEPs was significantly reduced in UT rats compared to sham controls, with treated (T) rats showing a significant improvement. Statistical significance (<span class="html-italic">p</span> &lt; 0.05) was observed between UT and T rats on Day 14 and Day 28. (<b>c</b>) Latency: The latency period of SEPs was significantly increased in injured rats compared to sham controls, with no significant differences between UT and T rats at Days 14 and 28. (<b>d</b>) Duration: The duration of SEPs, inversely related to nerve conduction velocity, was longer in UT rats compared to T rats. No significant differences were found between groups at Days 14 and 28. Data are expressed as mean ± standard deviation (SD) and were analyzed using one-way analysis of variance (ANOVA) followed by post hoc tests (<span class="html-italic">n</span> = 6). Statistical significance is indicated by * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 5
<p>Histological analysis of spinal cord: (<b>a</b>) Transverse Section: Transverse spinal cord sections stained with hematoxylin and eosin (H&amp;E) at 4 weeks post-injury. This view highlights the extent of tissue damage and demyelination in untreated (UT), treated (T), and sham (H) rats. Images are captured at 4× magnification with a scale bar of 500 µm. Significant differences in tissue integrity and lesion characteristics are evident between groups. Sham (H) rats show normal spinal cord morphology, while UT rats display extensive tissue damage and large cystic cavities. Treated (T) rats show reduced lesion size and less severe tissue degeneration compared to UT rats. (<b>b</b>) Longitudinal Section: Longitudinal spinal cord sections stained with H&amp;E, provide a detailed view of demyelination and histopathological features across the length of the injury. Images are taken at 10× magnification with a scale bar of 500 µm. Similar to the transverse sections, UT rats exhibit pronounced tissue loss and hemorrhagic foci, while T rats demonstrate reduced tissue damage and smaller lesions. Sham (H) rats present with normal spinal cord structure. (<b>c</b>) Relative Tissue Loss: Quantitative analysis of tissue loss in the center of the lesion, normalized to spinal cord sections from sham (H) rats without lesions. Bars represent the means and standard deviation (SD) of tissue loss measurements. Significant differences are observed between treatment groups, with T rats showing less tissue loss compared to UT rats (<span class="html-italic">p</span> &lt; 0.05). Data were analyzed by one-way analysis of variance (ANOVA) with the Dunnett post hoc test. (<b>d</b>) Lesion Size: Measurement of lesion size in the center of the injury site, comparing untreated (UT), treated (T), and sham (H) rats. Bars represent the means and standard deviation (SD) of lesion size measurements. Statistical Significance: Significant reduction in lesion size in T rats compared to UT rats (<span class="html-italic">p</span> &lt; 0.05 *). Data were analyzed using one-way ANOVA with the Dunnett post hoc test. These findings confirm that MLC901 treatment effectively mitigates spinal cord damage and supports tissue repair following mechanical compression injury.</p>
Full article ">Figure 6
<p>Comparison of GAP-43 and GFAP expression 28 days post-compression spinal cord injury with MLC901 treatment. (<b>a</b>,<b>b</b>) GAP-43 Expression: Immunofluorescence staining revealed that the intensity of Growth-associated protein 43 (GAP-43, green) was significantly higher in MLC901-treated (T) rats compared to untreated (UT) rats 28 days post-injury ((<b>a</b>,<b>b</b>); <span class="html-italic">p</span> &lt; 0.05 *). This increased GAP-43 immunoreactivity indicates enhanced neurogenesis and axonal growth in the T group. (<b>c</b>,<b>d</b>) GFAP Expression: Conversely, the intensity of Glial fibrillary acidic protein (GFAP, red) immunoreactivity was notably lower in the T rats compared to the UT rats ((<b>c</b>,<b>d</b>); <span class="html-italic">p</span> &lt; 0.05 *). Reduced GFAP staining suggests that MLC901 treatment effectively mitigates astrocytic scar formation and inflammation, contributing to a more favorable environment for neuroprotection and recovery. Nuclei Staining: Nuclei are stained with DAPI (blue), which helps to visualize cell bodies in the spinal cord sections. The higher magnification images (20×) on the right side of the figures showed morphological changes and detailed regions of demyelination with double staining for GAP-43-DAPI and GFAP-DAPI, highlighting the distinct areas of neurogenesis and astrocytic response. Scale bars are 1000 µm and 100 µm, providing context for the images’ magnification. Data are expressed as mean ± standard deviation (SD). Statistical significance was determined using Student’s <span class="html-italic">t</span>-test (<span class="html-italic">n</span> = 6). Significance is indicated as * <span class="html-italic">p</span> &lt; 0.05. These findings support that MLC901 treatment enhances neurogenesis while reducing astrocytic scar formation, contributing to improved functional recovery following spinal cord injury.</p>
Full article ">Figure 7
<p>MLC901 activated the PI3K/AKT/GSK-3β signaling pathway. (<b>a</b>) Western Blot Analysis: To assess the relative expression levels of p-AKT, AKT, p-GSK-3β, and GAP-43 proteins in spinal cord tissues from sham (H), treated (T), and untreated (UT) rats with β-Tubulin as a loading control. (<b>b</b>) Mean Fluorescence Intensity (MFI): To measure the expression levels of p-AKT/AKT, p-GSK-3β, and GAP-43 (neurogenesis marker) in spinal cord tissues of sham (H), treated (T), and untreated (UT) rats. (i) p-AKT/AKT Ratio: T rats showed higher expression compared to UT rats, indicating enhanced activation of the PI3K/AKT pathway. (ii) p-GSK-3β expression: UT rats showed higher expression compared to T rats, suggesting that MLC901 treatment inhibits GSK-3β activity and reduces apoptosis. (iii) GAP-43 Expression: Significantly higher expression in T rats compared to UT rats (<span class="html-italic">p</span> &lt; 0.05), indicating that MLC901 promotes neurogenesis. Statistical Analysis: Data are presented as mean ± standard deviation (SD) (<span class="html-italic">n</span> = 6/group). Statistical significance was determined using a one-way analysis of variance (ANOVA), followed by a post hoc test. * <span class="html-italic">p</span> &lt; 0.05 indicates significant differences between T rats and the UT SCI group.</p>
Full article ">Figure 8
<p>Histopathological Evaluation of Liver and Kidney Tissues. (<b>a</b>) Hepatic Tissue (Liver): Hematoxylin and eosin (H&amp;E) staining of liver tissue sections to assess potential hepatotoxicity associated with MLC901 treatment, observed under different magnifications (10×, 20×, and 40×), using scale bars 50, 100, and 500 µm. The images showed no sign of sinusoidal dilatation (enlargement of the hepatic capillaries), necrosis, hemorrhage, or congestion observed in treated (T) rats associated with MLC901 treatment, in liver tissue. MLC901 treatment did not induce hepatotoxicity after 28 days, indicating the liver remained healthy under the treatment conditions, <span class="html-italic">n</span> = 6. (<b>b</b>) Renal Tissue (Kidney): Hematoxylin and eosin (H&amp;E) staining of kidney tissue sections to evaluate potential nephrotoxicity associated with MLC901 treatment, observed under different magnifications (10×, 20×, and 40×), using scale bars 50, 100, and 500 µm. The images showed no vacuolization in tubular cells, focal necrosis, or hemorrhage observed in the kidney tissue of T rats. MLC901 treatment did not induce nephrotoxicity after 28 days, confirming that the kidney function remained intact, <span class="html-italic">n</span> = 6. Both liver and kidney tissues in T rats showed no adverse effects such as hepatotoxicity or nephrotoxicity, suggesting that MLC901 is safe for these organs after 28 days of treatment, UT represents untreated rats without MLC901 treatment and receiving normal saline.</p>
Full article ">
21 pages, 3875 KiB  
Review
Ubiquitination Insight from Spinal Muscular Atrophy—From Pathogenesis to Therapy: A Muscle Perspective
by Alfonso Bolado-Carrancio, Olga Tapia and José C. Rodríguez-Rey
Int. J. Mol. Sci. 2024, 25(16), 8800; https://doi.org/10.3390/ijms25168800 - 13 Aug 2024
Viewed by 784
Abstract
Spinal muscular atrophy (SMA) is one of the most frequent causes of death in childhood. The disease’s molecular basis is deletion or mutations in the SMN1 gene, which produces reduced survival motor neuron protein (SMN) levels. As a result, there is spinal motor [...] Read more.
Spinal muscular atrophy (SMA) is one of the most frequent causes of death in childhood. The disease’s molecular basis is deletion or mutations in the SMN1 gene, which produces reduced survival motor neuron protein (SMN) levels. As a result, there is spinal motor neuron degeneration and a large increase in muscle atrophy, in which the ubiquitin–proteasome system (UPS) plays a significant role. In humans, a paralogue of SMN1, SMN2 encodes the truncated protein SMNΔ7. Structural differences between SMN and SMNΔ7 affect the interaction of the proteins with UPS and decrease the stability of the truncated protein. SMN loss affects the general ubiquitination process by lowering the levels of UBA1, one of the main enzymes in the ubiquitination process. We discuss how SMN loss affects both SMN stability and the general ubiquitination process, and how the proteins involved in ubiquitination could be used as future targets for SMA treatment. Full article
(This article belongs to the Special Issue Advances in Neurodevelopmental-Related Disorders)
Show Figures

Figure 1

Figure 1
<p>Generation of SMN transcripts from <span class="html-italic">SMN1</span> and <span class="html-italic">SMN2</span> genes. (<b>A</b>) In healthy individuals, <span class="html-italic">SMN1</span> and <span class="html-italic">SMN2</span> gene transcripts are translated into the full-length SMN (SMN-FL) and SMNΔ7 proteins, respectively. A small percentage of <span class="html-italic">SMN2</span> is also translated into SMN-FL. (<b>B</b>) In SMA patients, transcripts from <span class="html-italic">SMN1</span> are absent. Most <span class="html-italic">SMN2</span> transcripts are translated into SMNΔ7 and mostly degraded. See the text for a more detailed explanation. Figure created with <a href="http://BioRender.com" target="_blank">BioRender.com</a>.</p>
Full article ">Figure 2
<p>Diagrammatic representation of the domain structure of SMN-FL. The <b>SMN protein</b> comprises several highly conserved motifs: a basic lysine-rich domain (K-rich), a Tudor domain, a poly-L-proline-rich domain (P-stretch), and a Y/G box in close proximity to the C-terminus, that itself mediates self-oligomerization and stability. The SMN protein is highly modified through phosphorylation (P), methylation (Me), acetylation (Ac), SUMOylation (S), and ubiquitination (Ub). Depicted are a summary of sites modified by the indicated PTMs identified through MS/proteomics and other methods (see [<a href="#B19-ijms-25-08800" class="html-bibr">19</a>] for a complete list of SMN PTMs sites). Some of the well-known proteins that interact with SMN-FL and the corresponding function are depicted below the corresponding interacting domain (see [<a href="#B20-ijms-25-08800" class="html-bibr">20</a>] for a review on SMN interactors and functional implications). Also indicated are the known interaction sites of SMN with the E3 UBLs Mib1, Itch and SCF<sup>Slmb</sup> and the DUB Bap1 (see <a href="#sec7-ijms-25-08800" class="html-sec">Section 7</a> for a detailed discussion). Figure created with <a href="http://BioRender.com" target="_blank">BioRender.com</a>.</p>
Full article ">Figure 3
<p>Outline of the ubiquitination process. An E1 ubiquitin-activating adenylates ubiquitin and forms an E1-Ub intermediate. Then, ubiquitin is transferred to the E2 ubiquitin-conjugating enzyme through a transthiolation reaction performed by E1. The E3 ligase forms an isopeptide bond between the substrate’s lysine side chain and the ubiquitin molecule’s C-terminal glycine. Monoubiquitinated molecules can either lose their ubiquitin moiety by the action of a deubiquitinase or become polyubiquitinated and further degraded in the proteasome. Figure created with <a href="http://BioRender.com" target="_blank">BioRender.com</a>.</p>
Full article ">Figure 4
<p>(<b>A</b>) Mono-ubiquitination of SMN would not preclude its incorporation into the SMN complex. The incorporation of SMN into the complex, in turn, would prevent SMN from polyubiquitination and proteasomal degradation. Mono-ubiquitination can be reversed by Usp9x and Bap1. (<b>B</b>) The inability of SMNΔ7 to form stable complexes, its decreased affinity for Usp9x and its different compartment distribution would make SMNΔ7 more prone to polyubiquitination and degradation by the UPS. Figure created with <a href="http://Biorender.com" target="_blank">Biorender.com</a>.</p>
Full article ">Figure 5
<p>A model to explain the functional link between SMN levels and ubiquitination and the changes in SMA. (<b>A</b>) Functional SMN complexes and correct UBE1 splicing render normal levels of UBA1 protein, which drive the ubiquitination and degradation of β-catenin and block the transcription of atrogenes. AKT activation reinforces the effect by phosphorylating FOXO and making it more sensitive to ubiquitination by MDM2. (<b>B</b>) In SMA patients, reduced levels of SMN result in low levels of UBA1, the accumulation of undegraded β-catenin, and increased atrogene expression. Figure created with <a href="http://Biorender.com" target="_blank">Biorender.com</a>.</p>
Full article ">
11 pages, 4821 KiB  
Article
Formation and Long-Term Culture of hiPSC-Derived Sensory Nerve Organoids Using Microfluidic Devices
by Takuma Ogawa, Souichi Yamada, Shuetsu Fukushi, Yuya Imai, Jiro Kawada, Kazutaka Ikeda, Seii Ohka and Shohei Kaneda
Bioengineering 2024, 11(8), 794; https://doi.org/10.3390/bioengineering11080794 - 5 Aug 2024
Viewed by 967
Abstract
Although methods for generating human induced pluripotent stem cell (hiPSC)-derived motor nerve organoids are well established, those for sensory nerve organoids are not. Therefore, this study investigated the feasibility of generating sensory nerve organoids composed of hiPSC-derived sensory neurons using a microfluidic approach. [...] Read more.
Although methods for generating human induced pluripotent stem cell (hiPSC)-derived motor nerve organoids are well established, those for sensory nerve organoids are not. Therefore, this study investigated the feasibility of generating sensory nerve organoids composed of hiPSC-derived sensory neurons using a microfluidic approach. Notably, sensory neuronal axons from neurospheres containing 100,000 cells were unidirectionally elongated to form sensory nerve organoids over 6 mm long axon bundles within 14 days using I-shaped microchannels in microfluidic devices composed of polydimethylsiloxane (PDMS) chips and glass substrates. Additionally, the organoids were successfully cultured for more than 60 days by exchanging the culture medium. The percentage of nuclei located in the distal part of the axon bundles (the region 3−6 mm from the entrance of the microchannel) compared to the total number of cells in the neurosphere was 0.005% for live cells and 0.008% for dead cells. Molecular characterization confirmed the presence of the sensory neuron marker ISL LIM homeobox 1 (ISL1) and the capsaicin receptor transient receptor potential vanilloid 1 (TRPV1). Moreover, capsaicin stimulation activated TRPV1 in organoids, as evidenced by significant calcium ion influx. Conclusively, this study demonstrated the feasibility of long-term organoid culture and the potential applications of sensory nerve organoids in bioengineered nociceptive sensors. Full article
Show Figures

Figure 1

Figure 1
<p>Microfluidic device for generating sensory nerve organoids. (<b>a</b>) Device design. (<b>b</b>) Photograph of the fabricated device. Scale bar: 5 mm.</p>
Full article ">Figure 2
<p>The formation of sensory nerve organoids. (<b>a</b>) Axon elongation in the microchannel. The 1–4 mm regions from the entrance of the microchannel are magnified. Scale bar: 1 mm. (<b>b</b>) The length of the leading axons in the microchannels. The vertical axis value of 6 mm indicates an axon length of 6 mm or more. The error bars indicate ± the standard error of the mean (SEM), <span class="html-italic">n</span> = 12.</p>
Full article ">Figure 3
<p>Representative images of a series of long-term cultured sensory nerve organoids. Axon bundle diameters at 3 mm from the entrance of the microchannel are indicated by two arrows. The 1–4 mm regions from the entrance of the microchannel are magnified. Scale bar: 1 mm.</p>
Full article ">Figure 4
<p>Analysis of live cell contamination in the axon bundles. (<b>a</b>) Visualization of the nuclei using fluorescence probes. The 1–6 mm regions from the entrance of the microchannel are magnified. Scale bar: 1 mm. (<b>b</b>) The number of nuclei derived from live and dead cells in each region of the axon bundles. The position indicates the distance from the entrance of the microchannels. Organoids cultured for 19 days were used. The error bars indicate ± SEM, <span class="html-italic">n</span> = 6.</p>
Full article ">Figure 5
<p>Characterization of neurospheres and axon bundles using fluorescent immunostaining. (<b>a</b>,<b>b</b>) TUJ1 as a neuron marker and (<b>a</b>) ISL1 as a sensory neuron marker. An organoid cultured for 39 days was used. (<b>b</b>) TRPV1 as a nociceptor marker. An organoid cultured for 28 days was used. White-colored scale bars: 500 μm. Gray-colored scale bars in close-up panels for axon bundles: 100 μm.</p>
Full article ">Figure 6
<p>Response of sensory nerve organoids to capsaicin stimulation. (<b>a</b>) Fluo 4-AM intensity indicating intracellular calcium ion influx following capsaicin (upper panel) and DMSO (control; lower panel) treatments. Capsaicin solution (100 μM) or DMSO was poured into reservoir of device at 0 s. Scale bars: 1 mm. (<b>b</b>) Time course of Fluo 4-AM fluorescent intensity in neurospheres and axon bundles. Timing for pouring 100 μM capsaicin solution or DMSO at 0 s is indicated by an arrow. ROIs for intensity measurements are dotted ellipses for neurospheres and dotted rectangles for axon bundles in (<b>a</b>). Representative data of organoids cultured for 32 days were used.</p>
Full article ">
13 pages, 533 KiB  
Review
Therapeutic Application of Modulators of Endogenous Cannabinoid System in Parkinson’s Disease
by Leonid G. Khaspekov and Sergey N. Illarioshkin
Int. J. Mol. Sci. 2024, 25(15), 8520; https://doi.org/10.3390/ijms25158520 - 5 Aug 2024
Viewed by 1027
Abstract
The endogenous cannabinoid system (ECS) of the brain plays an important role in the molecular pathogenesis of Parkinson’s disease (PD). It is involved in the formation of numerous clinical manifestations of the disease by regulating the level of endogenous cannabinoids and changing the [...] Read more.
The endogenous cannabinoid system (ECS) of the brain plays an important role in the molecular pathogenesis of Parkinson’s disease (PD). It is involved in the formation of numerous clinical manifestations of the disease by regulating the level of endogenous cannabinoids and changing the activation of cannabinoid receptors (CBRs). Therefore, ECS modulation with new drugs specifically designed for this purpose may be a promising strategy in the treatment of PD. However, fine regulation of the ECS is quite a complex task due to the functional diversity of CBRs in the basal ganglia and other parts of the central nervous system. In this review, the effects of ECS modulators in various experimental models of PD in vivo and in vitro, as well as in patients with PD, are analyzed. Prospects for the development of new cannabinoid drugs for the treatment of motor and non-motor symptoms in PD are presented. Full article
Show Figures

Figure 1

Figure 1
<p>ECS and clinical manifestations of PD. The predominant localization of CBR1s (red) and CBR2s (blue) are shown on the left. In the center and on the right, the relationships between the biological effects of ECS and the clinical symptoms of PD are shown. Possible “clinical targets” for ECS modulators are indicated by arrows.</p>
Full article ">
8 pages, 7630 KiB  
Case Report
Case Report: Molecular Analyses of Cell-Cycle-Related Genes in Cortical Brain Tissue of a Patient with Rasmussen Encephalitis
by João Ismael Budelon Gonçalves, Vinicius Rosa de Castro, William Alves Martins, Fernando Antonio Costa Xavier, Jaderson Costa Da Costa, Eliseu Paglioli Neto, André Palmini and Daniel Rodrigo Marinowic
Int. J. Mol. Sci. 2024, 25(15), 8487; https://doi.org/10.3390/ijms25158487 - 3 Aug 2024
Viewed by 700
Abstract
Rasmussen’s encephalitis (RE) stands as a rare neurological disorder marked by progressive cerebral hemiatrophy and epilepsy resistant to medical treatment. Despite extensive study, the primary cause of RE remains elusive, while its histopathological features encompass cortical inflammation, neuronal degeneration, and gliosis. The underlying [...] Read more.
Rasmussen’s encephalitis (RE) stands as a rare neurological disorder marked by progressive cerebral hemiatrophy and epilepsy resistant to medical treatment. Despite extensive study, the primary cause of RE remains elusive, while its histopathological features encompass cortical inflammation, neuronal degeneration, and gliosis. The underlying molecular mechanisms driving disease progression remain largely unexplored. In this case study, we present a patient with RE who underwent hemispherotomy and has remained seizure-free for over six months, experiencing gradual motor improvement. Furthermore, we conducted molecular analysis on the excised brain tissue, unveiling a decrease in the expression of cell-cycle-associated genes coupled with elevated levels of BDNF and TNF-α proteins. These findings suggest the potential involvement of cell cycle regulators in the progression of RE. Full article
(This article belongs to the Special Issue Molecular Advances in Epilepsy and Seizures)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>–<b>G</b>) MRI and EEG from patient with RE: (<b>A</b>–<b>C</b>) brain MR images highlighting T2 and T2 FLAIR. Hyperintensity in the right frontotemporal-insular region suggests subacute encephalitis with some degree of cortical atrophy; (<b>D</b>–<b>F</b>) brain MRI three months later shows extensive right hemisphere atrophy, predominating in the perisylvian region and affecting the caudate; (<b>G</b>) representative EEG with continuous rhythmic activity detected in the Fp2-F8 region.</p>
Full article ">Figure 2
<p>Molecular analysis of brain tissue of patient with RE. (<b>A</b>) Heatmap from RT 2 Profiler™ PCR Array human cell cycle pathway genes. Data are representative of log2 fold change difference from a control group (<span class="html-italic">n</span> = 5) and a Rasmussen patient. (<b>B</b>) Volcano plot for control group versus Rasmussen patient showing significantly down-regulated genes which passed the −1.3 (<span class="html-italic">p</span> &lt; 0.05) threshold for log2 fold change difference. <span class="html-italic">p</span>-values are calculated based on a Student’s <span class="html-italic">t</span>-test of the mean 2 −ΔCT values for each gene in the control group versus the Rasmussen patient. (<b>C</b>) Log2 fold change from top five down-regulated genes. (<b>D</b>) Protein levels from Luminex immunoassay.</p>
Full article ">
19 pages, 6204 KiB  
Article
Age-Related Effects of Inhalational Anesthetics in B4galnt1-Null and Cuprizone-Treated Mice: Clinically Relevant Insights into Demyelinating Diseases
by Ozana Katarina Tot, Stefan Mrđenović, Vedrana Ivić, Robert Rončević, Jakov Milić, Barbara Viljetić and Marija Heffer
Curr. Issues Mol. Biol. 2024, 46(8), 8376-8394; https://doi.org/10.3390/cimb46080494 - 1 Aug 2024
Viewed by 425
Abstract
Anesthetics are essential agents that are frequently used in clinical practice to induce a reversible loss of consciousness and sensation by depressing the central nervous system. The inhalational anesthetics isoflurane and sevoflurane are preferred due to their rapid induction and recovery times and [...] Read more.
Anesthetics are essential agents that are frequently used in clinical practice to induce a reversible loss of consciousness and sensation by depressing the central nervous system. The inhalational anesthetics isoflurane and sevoflurane are preferred due to their rapid induction and recovery times and ease of administration. Despite their widespread use, the exact molecular mechanisms by which these anesthetics induce anesthesia are not yet fully understood. In this study, the age-dependent effects of inhalational anesthetics on two demyelination models were investigated: congenital (B4galnt1-null) and chemically induced (cuprizone). Various motor and cognitive tests were used to determine sensitivity to isoflurane and sevoflurane anesthesia. B4galnt1-null mice, which exhibit severe motor deficits due to defects in ganglioside synthesis, showed significant impairments in motor coordination and balance in all motor tests, which were exacerbated by both anesthetics. Cuprizone-treated mice, which mimic the demyelination in B4galnt1-null mice, also showed altered, age-dependent sensitivity to anesthesia. The study showed that older mice exhibited more pronounced deficits, with B4galnt1-null mice showing the greatest susceptibility to sevoflurane. These differential responses to anesthetics suggest that age and underlying myelin pathology significantly influence anesthetic effects. Full article
(This article belongs to the Special Issue Membrane Transporters and Channels in Health and Diseases)
Show Figures

Figure 1

Figure 1
<p>Assessment of the locomotor abilities of wild-type (WT), cuprizone-treated wild-type (WT + CPZ) and <span class="html-italic">B4galnt1</span>-null (KO) mice aged 6 and 12 months (6 m and 12 m). (<b>A</b>) Basso–Beattie–Bresnahan (BBB) assessment of the locomotor performance of the mice before motor and cognitive tests (median with interquartile range; Kruskal–Wallis test followed by Dunn’s multiple comparison test; ** <span class="html-italic">p</span> &lt; 0.01, **** <span class="html-italic">p</span> &lt; 0.0001); (<b>B</b>) Times necessary for the induction of and awakening from anesthesia with isoflurane or sevoflurane (minimum to maximum values; the line in the box represents the median; Mann–Whitney U-test; * <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>Analyses of behavioral tests of locomotor assessment in wild-type (WT), cuprizone-treated wild-type (WT + CPZ) and <span class="html-italic">B4galnt1</span>-null (KO) mice aged 6 and 12 months (6 m and 12 m). (<b>A</b>) Hindlimb extension reflex test in non-anesthetized mice (NA) and mice after anesthesia with isoflurane or sevoflurane (Kruskal–Wallis test followed by Dunn’s multiple comparison test; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, **** <span class="html-italic">p</span> &lt; 0.0001). Kaplan–Meier survival analysis using the log–rank test (Mantel–Cox test) to compare forelimb grip strength (<b>B</b>) and rotarod endurance (<b>C</b>) of non-anesthetized mice (NA) and mice after anesthesia with isoflurane or sevoflurane.</p>
Full article ">Figure 3
<p>Analyses of cognitive behavioral tests in wild-type (WT), cuprizone-treated wild-type (WT + CPZ) and <span class="html-italic">B4galnt1</span>-null (KO) mice aged 6 and 12 months (6 m and 12 m). (<b>A</b>) Modified Lashley III maze test from non-anesthetized mice (NA) and mice after anesthesia with isoflurane or sevoflurane (Kruskal–Wallis test, followed by Dunn’s multiple comparison test for comparisons between groups with respect to the anesthetic used and the Mann–Whitney U test for pairwise comparisons between groups treated with different anesthetics; * <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). (<b>B</b>) The Kaplan–Meier survival analysis using the log–rank test (Mantel–Cox test) to compare the passive avoidance of non-anesthetized mice (NA) and mice after anesthesia with isoflurane or sevoflurane.</p>
Full article ">Figure 4
<p>Immunohistochemical staining of myelination markers (MAG, MBP and CNPase) in the brains of wild-type (WT), cuprizone-treated wild-type (WT + CPZ) and <span class="html-italic">B4galnt1</span>-null (KO) mice on coronal sections of the forebrain. The arrows indicate the sites with the greatest differences in the distribution of the individual markers.</p>
Full article ">Figure 5
<p>Expression level of myelination markers (MAG, MBP and CNPase) in the <span class="html-italic">corpus callosum</span> of the brains of wild-type (WT), cuprizone-treated wild-type (WT + CPZ) and <span class="html-italic">B4galnt1</span>-null (KO) mice. (<b>A</b>) Immunohistochemical staining of myelination markers (total magnification is 100×) and (<b>B</b>) their evaluation (median and interquartile range of integrated optical density). Mann–Whitney U test; * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 6
<p>Immunohistochemical staining of the four major brain gangliosides (GM1, GD1a, GD1b and GT1b) in the brains of wild-type (WT) and cuprizone-treated wild-type (WT + CPZ) mice on coronal sections of the forebrain.</p>
Full article ">Figure 7
<p>Expression level of the four major brain gangliosides (GM1, GD1a, GD1b and GT1b) in the <span class="html-italic">corpus callosum</span> of the brains of wild-type (WT) and cuprizone-treated wild-type (WT + CPZ) mice. (<b>A</b>) Immunohistochemical staining of gangliosides (total magnification is 100×) and (<b>B</b>) their evaluation (median and interquartile range of integrated optical density). Kruskal–Wallis test followed by Dunn’s multiple comparison test; * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">
20 pages, 5147 KiB  
Article
Plastic Fly: What Drosophila melanogaster Can Tell Us about the Biological Effects and the Carcinogenic Potential of Nanopolystyrene
by Massimo Aloisi, Daniela Grifoni, Osvaldo Zarivi, Sabrina Colafarina, Patrizia Morciano and Anna Maria Giuseppina Poma
Int. J. Mol. Sci. 2024, 25(14), 7965; https://doi.org/10.3390/ijms25147965 - 21 Jul 2024
Viewed by 1053
Abstract
Today, plastic pollution is one of the biggest threats to the environment and public health. In the tissues of exposed species, micro- and nano-fragments accumulate, leading to genotoxicity, altered metabolism, and decreased lifespan. A model to investigate the genotoxic and tumor-promoting potential of [...] Read more.
Today, plastic pollution is one of the biggest threats to the environment and public health. In the tissues of exposed species, micro- and nano-fragments accumulate, leading to genotoxicity, altered metabolism, and decreased lifespan. A model to investigate the genotoxic and tumor-promoting potential of nanoplastics (NPs) is Drosophila melanogaster. Here we tested polystyrene, which is commonly used in food packaging, is not well recycled, and makes up at least 30% of landfills. In order to investigate the biological effects and carcinogenic potential of 100 µm polystyrene nanoparticles (PSNPs), we raised Oregon [R] wild-type flies on contaminated food. After prolonged exposure, fluorescent PSNPs accumulated in the gut and fat bodies. Furthermore, PSNP-fed flies showed considerable alterations in weight, developmental time, and lifespan, as well as a compromised ability to recover from starvation. Additionally, we noticed a decrease in motor activity in DNAlig4 mutants fed with PSNPs, which are known to be susceptible to dietary stressors. A qPCR molecular investigation of the larval intestines revealed a markedly elevated expression of the genes drice and p53, suggesting a response to cell damage. Lastly, we used warts-defective mutants to assess the carcinogenic potential of PSNPs and discovered that exposed flies had more aberrant masses than untreated ones. In summary, our findings support the notion that ingested nanopolystyrene triggers metabolic and genetic modifications in the exposed organisms, eventually delaying development and accelerating death and disease. Full article
(This article belongs to the Section Molecular Biology)
Show Figures

Figure 1

Figure 1
<p>Morphological analysis (<b>A</b>,<b>B</b>) and energy dispersive X-ray analysis (<b>C</b>,<b>D</b>) of polystyrene nanoparticles, non-fluorescent (<b>A</b>,<b>C</b>) and fluorescent (<b>C</b>,<b>D</b>).</p>
Full article ">Figure 2
<p>SEM analysis of PSNPs in feces from untreated flies (CTRL) and three different concentrations (50 μg/mL, 100 μg/mL, and 750 μg/mL). Left column has 172 X magnification for CTRL and 750 μg/mL, 20.00 K X for 50 μg/mL, and 1.00 K X for 100 μg/mL; the right column magnification is 50.00 K X for every condition. White arrows point to NPs.</p>
Full article ">Figure 3
<p>Fluorescence microscopy of fPSNPs in OR-R third instar larvae guts and fat bodies in two conditions: negative control (no fPSNPs, 0 μg/mL) and 750 μg/mL (fPSNPs). (<b>A</b>,<b>C</b>) dissected guts and fat bodies, respectively (400X); (<b>B</b>,<b>D</b>) whole living larvae (100X).</p>
Full article ">Figure 4
<p>Analysis of development traits of OR-R flies: pupae formation (<b>A</b>) and adult timing (<b>B</b>). Treatments: Control (0 µg/mL); 50 µg/mL, 100 µg/mL, 750 µg/mL. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.005; *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 5
<p>Weight distribution of 3 days old OR-R flies. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.005.</p>
Full article ">Figure 6
<p>Starvation assay with OR-R flies. Three days-old adults chronically exposed were starved in empty vials and checked hourly for the first 12 h (<a href="#app1-ijms-25-07965" class="html-app">Figure S2A,B Supplementary Materials</a>) and after 24 h (<b>A</b>,<b>B</b>). Treatments: CTRL (0 µg/mL); 50 µg/mL; 100 µg/mL; 750 µg/mL. * <span class="html-italic">p</span> &lt; 0.05; *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 7
<p>Climbing assay of OR-R (<b>A</b>,<b>B</b>,<b>E</b>,<b>F</b>) and DNA<span class="html-italic">lig4</span> (<b>C</b>,<b>D</b>,<b>G</b>,<b>H</b>) flies after 3 (<b>A</b>–<b>D</b>) and 18 (<b>E</b>–<b>H</b>) days from eclosion, divided into males (<b>A</b>,<b>C</b>,<b>E</b>,<b>G</b>) and females (<b>B</b>,<b>D</b>,<b>F</b>,<b>H</b>). * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.005; *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 8
<p>Trypan blue assay in wild-type third instar larvae gut fed with PSNPs. (<b>A</b>) dissected guts were incubated with Trypan blue and then scored under a stereoscope based on blue intensity and spreading; (<b>B</b>) violin plot and box plot showing the final scoring of the four conditions used (0, 50, 100, and 750 μg/mL). * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 9
<p>A schematic drawing of the apoptotic process. Image realized with Biorender.com. Gene expression of apoptosis (<b>A</b>) and endocytosis (<b>B</b>) biomarkers in third instar larvae guts. Student’s <span class="html-italic">t</span> test was employed to compare the values of untreated (ctrl) and treated. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.005.</p>
Full article ">Figure 10
<p>Dissected <span class="html-italic">OR-R</span> third instar larvae guts were homogenized to obtain a cell suspension that was subjected to an electrophoretic run and subsequent nuclear stain with ethidium bromide to observe DNA breaks under a fluorescence microscope (<b>A</b>). The percentage of DNA detected in comet tails, the tail moment and the olive tail moment are reported (<b>B</b>–<b>D</b>). Non-parametric one-way ANOVA was used for significance analysis. Non-significant conditions were grouped with the same letter; different letters represent significance between conditions. Treatment: Control (0 µg/mL); 50 µg/mL; 100 µg/mL.</p>
Full article ">Figure 11
<p><span class="html-italic">warts</span> assay showing carcinogenic potential of PSNPS. (<b>A</b>) representative images of tumors in 3 days old <span class="html-italic">wts</span> defective flies (white arrows); (<b>B</b>) table resuming the percentage of masses in the <span class="html-italic">wts</span> fly population; (<b>C</b>), bar plot showing the percentage of total alterations. Treatment: CTRL (0 µg/mL); 50 µg/mL, 100 µg/mL, 750 µg/mL. * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">
10 pages, 2508 KiB  
Article
CDC’s Laboratory Activities to Support Newborn Screening for Spinal Muscular Atrophy
by Francis K. Lee, Christopher Greene, Kristina Mercer, Jennifer Taylor, Golriz Yazdanpanah, Robert Vogt, Rachel Lee, Carla Cuthbert and Suzanne Cordovado
Int. J. Neonatal Screen. 2024, 10(3), 51; https://doi.org/10.3390/ijns10030051 - 17 Jul 2024
Viewed by 815
Abstract
Spinal muscular atrophy (SMA) was added to the HHS Secretary’s Recommended Uniform Screening Panel for newborn screening (NBS) in 2018, enabling early diagnosis and treatment of impacted infants to prevent irreversible motor neuron damage. In anticipation of supporting SMA newborn screening, scientists at [...] Read more.
Spinal muscular atrophy (SMA) was added to the HHS Secretary’s Recommended Uniform Screening Panel for newborn screening (NBS) in 2018, enabling early diagnosis and treatment of impacted infants to prevent irreversible motor neuron damage. In anticipation of supporting SMA newborn screening, scientists at the U.S. Centers for Disease Control and Prevention (CDC) have worked towards building resources for public health laboratories in four phases since 2013. In Phase 1, CDC established a real-time PCR assay, which uses a locked nucleic acid probe to attain the needed specificity, to detect SMN1 exon 7. In Phase 2, we developed quality assurance dried blood spot materials made with transduced lymphoblast cell lines established from de-identified SMA patients, carriers, and unaffected donors. In 2021, CDC implemented Phase 3, a proficiency testing program, that now supports 115 NBS labs around the world. We are currently completing Phase 4, which includes the implementation of an external SMA quality control material program. Also, during this time, CDC has provided individual technical assistance to NBS programs and bench training to NBS scientists during our annual molecular workshop. These CDC-led activities have contributed to the rapid and full implementation of SMA screening in all 50 U.S. states as of February 2024. Full article
(This article belongs to the Special Issue Newborn Screening for SMA—State of the Art)
Show Figures

Figure 1

Figure 1
<p>Real-time PCR amplification plots using the CDC assay from SMA quality assurance materials from cell lines established from de-identified donors with (<b>A</b>) normal <span class="html-italic">SMN1</span> exon 7 sequence, (<b>B</b>) heterozygous deletion of <span class="html-italic">SMN1</span> exon 7 sequence, and (<b>C</b>) homozygous deletion of <span class="html-italic">SMN1</span> exon 7 sequences.</p>
Full article ">Figure 1 Cont.
<p>Real-time PCR amplification plots using the CDC assay from SMA quality assurance materials from cell lines established from de-identified donors with (<b>A</b>) normal <span class="html-italic">SMN1</span> exon 7 sequence, (<b>B</b>) heterozygous deletion of <span class="html-italic">SMN1</span> exon 7 sequence, and (<b>C</b>) homozygous deletion of <span class="html-italic">SMN1</span> exon 7 sequences.</p>
Full article ">Figure 2
<p>Number of participants in CDC’s Newborn Screening Quality Assurance Program, NSQAP, for SMA. Black bars represent U.S. participants, and grey represents the number of international participants.</p>
Full article ">Figure 3
<p>SMA PT misclassifications by year.</p>
Full article ">
Back to TopTop