[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
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

Journals

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

Search Results (379)

Search Parameters:
Keywords = event-related brain potentials

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
14 pages, 3388 KiB  
Article
Impact of Larval Sertraline Exposure on Alternative Splicing in Neural Tissue of Adult Drosophila melanogaster
by Luis Felipe Santos-Cruz, Myriam Campos-Aguilar, Laura Castañeda-Partida, Santiago Cristobal Sigrist-Flores, María Eugenia Heres-Pulido, Irma Elena Dueñas-García, Elías Piedra-Ibarra, Rafael Jiménez-Flores and Alberto Ponciano-Gómez
Int. J. Mol. Sci. 2025, 26(2), 563; https://doi.org/10.3390/ijms26020563 - 10 Jan 2025
Viewed by 433
Abstract
Sertraline, a selective serotonin reuptake inhibitor (SSRI), is commonly used to treat various psychiatric disorders such as depression and anxiety due to its ability to increase serotonin availability in the brain. Recent findings suggest that sertraline may also influence the expression of genes [...] Read more.
Sertraline, a selective serotonin reuptake inhibitor (SSRI), is commonly used to treat various psychiatric disorders such as depression and anxiety due to its ability to increase serotonin availability in the brain. Recent findings suggest that sertraline may also influence the expression of genes related to synaptic plasticity and neuronal signaling pathways. Alternative splicing, a process that allows a single gene to produce multiple protein isoforms, plays a crucial role in the regulation of neuronal functions and plasticity. Dysregulation of alternative splicing events has been linked to various neurodevelopmental and neurodegenerative diseases. This study aims to explore the effects of sertraline on alternative splicing events, including exon inclusion, exon exclusion, and mutually exclusive splicing events, in genes associated with neuronal function in Drosophila melanogaster and to use this model to investigate the molecular impacts of SSRIs on gene regulation in the nervous system. RNA sequencing (RNA-seq) was performed on central nervous system samples from Drosophila melanogaster adults exposed to sertraline for 24 h when they were third instar larvae. Alternative splicing events were analyzed to identify changes in exon inclusion and exclusion, as well as intron retention. Sertraline treatment significantly altered alternative splicing patterns in key genes related to neuronal stability and function. Specifically, sertraline promoted the inclusion of long Ank2 isoforms, suggesting enhanced axonal stability, and favored long ATPalpha isoforms, which support Na+/K+ ATPase activity essential for ionic balance and neuronal excitability. Intron retention in the yuri gene suggests that cytoskeletal reorganization could impact neuronal morphology. Additionally, splicing alterations in sxc and Atg18a indicate a potential influence of sertraline on epigenetic regulation and autophagy processes, fundamental aspects for neuronal plasticity and cellular homeostasis. These findings suggest that sertraline influences alternative splicing in the central nervous system of Drosophila melanogaster, potentially contributing to its therapeutic effects by modulating neuronal stability and adaptability. Full article
(This article belongs to the Special Issue Cell Pathways Underlying Neuronal Differentiation)
Show Figures

Figure 1

Figure 1
<p>Sashimi plots illustrating statistically significant alternative splicing events by exon inclusion in the (<b>A</b>) <span class="html-italic">Ank2</span> and (<b>B</b>) <span class="html-italic">ATPalpha</span> genes. The red histograms represent RNA-seq read coverage for sertraline-treated samples, while the orange histograms correspond to control samples. The black blocks indicate annotated genomic regions, and the connecting lines represent spliced regions. For <span class="html-italic">Ank2</span>, sertraline-treated samples exhibited higher exon inclusion levels compared to controls, suggesting a shift towards the long <span class="html-italic">Ank2</span> isoform. Similarly, for <span class="html-italic">ATPalpha</span>, sertraline treatment promoted the inclusion of the long isoform, which is associated with enhanced neuronal functionality. Abbreviations: IncLevel—inclusion level, RPKM—reads per kilobase million.</p>
Full article ">Figure 2
<p>Sashimi plots displaying significant intron retention events in the genes (<b>A</b>) <span class="html-italic">yuri</span>, (<b>B</b>) <span class="html-italic">sxc</span>, (<b>C</b>) <span class="html-italic">Atg18a</span>, and (<b>D</b>) <span class="html-italic">stmA</span>. The red histograms represent RNA-seq read coverage for sertraline-treated samples, while the orange histograms correspond to control samples. The connecting lines represent spliced regions. For the <span class="html-italic">yuri</span> gene, intron retention was significantly higher in sertraline-treated samples compared to controls, suggesting a splicing alteration that could affect the expression of the functional isoform. For <span class="html-italic">sxc</span>, <span class="html-italic">Atg18a</span>, and <span class="html-italic">stmA</span>, sertraline treatment reduced intron retention, which may favor the expression of functional isoforms. Abbreviations: IncLevel—inclusion level, RPKM—reads per kilobase million.</p>
Full article ">Figure 3
<p>Sashimi plots displaying significant mutually exclusive exon splicing events in the genes (<b>A</b>) Sam-S, (<b>B</b>) Tm1, (<b>C</b>) gish, and (<b>D</b>) Tep2. The red histograms represent RNA-seq read coverage for sertraline-treated samples, while the orange histograms correspond to control samples. The connecting lines represent spliced regions. For Sam-S and Tm1, sertraline treatment increased the inclusion of alternative exons, while in gish, exon exclusion was observed. In Tep2, there was a higher inclusion of alternative exons following treatment. Abbreviations: IncLevel—inclusion level, RPKM—reads per kilobase million.</p>
Full article ">
19 pages, 15090 KiB  
Article
Time Course of Brain Activity Changes Related to Number (Quantity) Processing Triggered by Digits Versus Number Words: An Event-Related Potential (ERP) Study
by Peter Walla and Philipp Klimovic
Appl. Sci. 2025, 15(2), 530; https://doi.org/10.3390/app15020530 - 8 Jan 2025
Viewed by 387
Abstract
The neuroscience of language processing in the human brain has a long history. Strings of letters that form meaningful words trigger lexical and semantic processing, which in turn lead to conscious awareness of what the words mean. However, it is still unclear how [...] Read more.
The neuroscience of language processing in the human brain has a long history. Strings of letters that form meaningful words trigger lexical and semantic processing, which in turn lead to conscious awareness of what the words mean. However, it is still unclear how the brain processes normal words differently from number words and, more interestingly, how the brain processes number words differently from digits, both of which are meant to trigger quantity processing. While much of the literature deals with this topic, the time course of the respective differences in brain activity has been largely ignored. This may be because most studies have used functional magnetic resonance imaging (fMRI), which is known to have limited temporal resolution. This study used electroencephalography (EEG), more specifically event-related potentials (ERPs), to investigate brain potential differences between visual presentations of words, non-words, number words and digits. This approach made it possible to describe the time course of brain activity evoked by these four stimulus categories. Starting at about 200 ms post-stimulus, digits elicited the strongest negative ERP in the right occipito-parietal cortical region. Peaking at around 300 ms after stimulus onset, number words elicited the most negative going ERP in the left occipito-parietal area. Finally, starting at about 400 ms after stimulus onset, digits elicited by far the most negative ERP in the left inferior fronto-temporal area. All of these findings are supported by analytical statistics across all study participants. It is noteworthy that the last effect in the left inferior fronto-temporal area can also be seen for number words, but it is much smaller and not statistically significant. In summary, we found clear differences between brain activity related to the processing of words, non-words, number words, and digits, providing evidence that the left inferior fronto-temporal cortical area is specialised for the processing of quantities. Furthermore, it can be concluded that digits are better symbols for mediating quantity processing in the human brain than number words. Full article
(This article belongs to the Section Applied Neuroscience and Neural Engineering)
Show Figures

Figure 1

Figure 1
<p>Sequence of screen presentations. On top is a stimulus example (‘siebzehn’; German for ‘seventeen’), and the screen in the middle was meant to prompt a key press response (‘Was haben Sie gesehen?’; Germen for ‘What did you see?’).</p>
Full article ">Figure 2
<p>(<b>Top</b>): Selected electrode locations. (<b>Bottom</b>): Event-related potentials (ERPs) generated for all four stimulus categories for the six selected electrode positions. As can be seen below in more detail, analysis of variance (ANOVA) revealed three time-windows showing significant condition effects (see Table 2). First, from 219 ms to 251 ms, second, from 299 ms to 363 ms, and finally, from 427 ms to 587 ms (see time-windows marked in light blue colour. At the earliest time-window, digits elicited the most negative brain potentials with a maximum effect at electrode location P8 (right inferior parietal; see Table 3 showing respective <span class="html-italic">t</span>-test results). During the time-window from 299 ms to 363 ms, number words and non-words elicited the most negative brain potentials (maximum effect at electrode location P7; see Table 4). Finally, during the time-window from 427 ms to 587 ms, actual numbers elicited by far the most negative brain potentials with a maximum effect at electrode location F7. See Table 5 for respective <span class="html-italic">t</span>-test results that confirm and underline these findings.</p>
Full article ">Figure 3
<p>Topographical maps including data from all electrode locations created for all 4 stimulus conditions for the time point 227 ms post-stimulus. Note that the largest negative-going ERPs were elicited by digits in the right parietal region (see location marked with a red circle).</p>
Full article ">Figure 4
<p>Topographical maps including data from all electrode locations created for all 4 stimulus conditions for the time point 307 ms post-stimulus. Note that the largest negative-going ERPs were elicited by words in the left temporo-parietal region (see location marked with a red circle).</p>
Full article ">Figure 5
<p>Topographical maps including data from all electrode locations created for all 4 stimulus conditions for the time point 467 ms post-stimulus. Note that the largest negative-going ERPs were elicited by digits in the left frontal region (see location marked with a red circle). This is followed by number words. Both areas are marked with red circles.</p>
Full article ">
32 pages, 2920 KiB  
Review
EEG in Education: A Scoping Review of Hardware, Software, and Methodological Aspects
by Christos Orovas, Theodosios Sapounidis, Christina Volioti and Euclid Keramopoulos
Sensors 2025, 25(1), 182; https://doi.org/10.3390/s25010182 - 31 Dec 2024
Viewed by 760
Abstract
Education is an activity that involves great cognitive load for learning, understanding, concentrating, and other high-level cognitive tasks. The use of the electroencephalogram (EEG) and other brain imaging techniques in education has opened the scientific field of neuroeducation. Insights about the brain mechanisms [...] Read more.
Education is an activity that involves great cognitive load for learning, understanding, concentrating, and other high-level cognitive tasks. The use of the electroencephalogram (EEG) and other brain imaging techniques in education has opened the scientific field of neuroeducation. Insights about the brain mechanisms involved in learning and assistance in the evaluation and optimization of education methodologies according to student brain responses is the main target of this field. Being a multidisciplinary field, neuroeducation requires expertise in various fields such as education, neuroinformatics, psychology, cognitive science, and neuroscience. The need for a comprehensive guide where various important issues are presented and examples of their application in neuroeducation research projects are given is apparent. This paper presents an overview of the current hardware and software options, discusses methodological issues, and gives examples of best practices as found in the recent literature. These were selected by applying the PRISMA statement to results returned by searching PubMed, Scopus, and Google Scholar with the keywords “EEG and neuroeducation” for projects published in the last six years (2018–2024). Apart from the basic background knowledge, two research questions regarding methodological aspects (experimental settings and hardware and software used) and the subject of the research and type of information used from the EEG signals are addressed and discussed. Full article
(This article belongs to the Special Issue Smart Educational Systems: Hardware and Software Aspects)
Show Figures

Figure 1

Figure 1
<p>Example of 5 s EEG recording (source: sample data plot from EEGLAB [<a href="#B14-sensors-25-00182" class="html-bibr">14</a>]). There are 32 channels, with their naming derived from the 10–20 system, and the scale refers to microvolts (μV).</p>
Full article ">Figure 2
<p>The conceptual framework for the usage of software in EEG applications.</p>
Full article ">Figure 3
<p>The PRISMA flow diagram.</p>
Full article ">Figure 4
<p>Counts of the sample sizes in groups of ten.</p>
Full article ">Figure 5
<p>Counts of wired and wireless EEG recordings in each group of sample sizes.</p>
Full article ">Figure 6
<p>Counts of use of EEG devices and configurations (amplifier and caps).</p>
Full article ">Figure 7
<p>The way in which EEG signals were used in the presented projects.</p>
Full article ">
10 pages, 497 KiB  
Article
Olfactory Evoked Potentials and Brain MRI Outcomes in Multiple Sclerosis Patients: A Cross-Sectional Study
by Rosella Ciurleo, Simona De Salvo, Fabrizia Caminiti, Annalisa Militi and Lilla Bonanno
J. Clin. Med. 2025, 14(1), 141; https://doi.org/10.3390/jcm14010141 - 29 Dec 2024
Viewed by 565
Abstract
Background: Olfactory dysfunction (OD) is an underestimated symptom in multiple sclerosis (MS). Multiple factors may play a role in the OD reported by MS patients, such as ongoing inflammation in the central nervous system (CNS), damage to the olfactory bulbs due to demyelination, [...] Read more.
Background: Olfactory dysfunction (OD) is an underestimated symptom in multiple sclerosis (MS). Multiple factors may play a role in the OD reported by MS patients, such as ongoing inflammation in the central nervous system (CNS), damage to the olfactory bulbs due to demyelination, and the presence of plaques in brain areas associated with the olfactory system. Indeed, neuroimaging studies in MS have shown a clear association of the OD with the number and activity of MS-related plaques in frontal and temporal brain regions. However, these studies have used only psychophysical tests to evaluate the OD in MS patients. Olfactory Event-Related Potentials (OERPs) are a method to assess olfaction with the clear advantage of its objectivity in comparison with psychophysical tests. The aim of this study was to investigate the association between the parameters of OERP components (latency and amplitude) and the lesion load of the brain regions which are involved in olfaction in a cohort of relapsing-remitting (RR) MS patients. Methods: In this cross-sectional study, we enrolled 30 RRMS patients and 30 healthy controls. The parameters of OERP components and magnetic resonance imaging data (lesions in the CNS) were analyzed in RRMS patients. Results: The association found between the RRMS patient groups with and without OERPs and the number of lesions in the frontal area as well as the correlation between the lesion load in the temporal area and OERP parameters suggest how brain alterations may impact on olfactory performance in MS. In addition, the predictive value of the number of lesions in the frontal and parietal areas for P2 amplitude also highlights the potential for OERP measures to serve as markers for disease progression in MS. Conclusions: This approach to assess the olfaction in MS could improve our understanding of the disease’s neurological impact and contribute to the development of new targeted interventions to mitigate olfactory sensory deficits. Full article
Show Figures

Figure 1

Figure 1
<p>Study protocol and RRMS patients’ selection. <b>Legend:</b> OERPs = Olfactory Event Related Potentials; RRMS = relapsing-remitting multiple sclerosis; MRI = magnetic resonance imaging.</p>
Full article ">
17 pages, 1828 KiB  
Article
Dynamic Neuro-Glial-Vascular Responses in a Mouse Model of Vascular Cognitive Impairment
by Ki Jung Kim, Rachel E. Patterson, Juan Ramiro Diaz, Philip O’Herron, Weston Bush, Ferdinand Althammer, Javier E. Stern, Michael W. Brands, Zsolt Bagi and Jessica A. Filosa
Neuroglia 2024, 5(4), 505-521; https://doi.org/10.3390/neuroglia5040032 - 19 Dec 2024
Viewed by 917
Abstract
Background: Chronic hypoperfusion is a risk factor for neurodegenerative diseases. However, the sequence of events driving ischemia-induced functional changes in a cell-specific manner is unclear. Methods: To address this gap in knowledge, we used the bilateral common carotid artery stenosis (BCAS) mouse model, [...] Read more.
Background: Chronic hypoperfusion is a risk factor for neurodegenerative diseases. However, the sequence of events driving ischemia-induced functional changes in a cell-specific manner is unclear. Methods: To address this gap in knowledge, we used the bilateral common carotid artery stenosis (BCAS) mouse model, and evaluated progressive functional changes to neurons, arterioles, astrocytes, and microglial cells at 14 and 28 days post-BCAS surgery. To assess the neuro-glio-vascular response to an acute ischemic insult, brain slices were superfused with low O2 conditions. Using whole-cell patch-clamp electrophysiology, we measured basic membrane properties (e.g., resting membrane potential, capacitance, input resistance) in cortical pyramidal neurons. The activity of astrocytes was evaluated by monitoring Ca2+ from Aldh1l1-CreERT2; R26-lsl-GCaMP6f mice. Vascular reactivity to low O2 from the BCAS mice was also assessed ex vivo. Results: Our data showed no changes to the basic membrane properties of cortical pyramidal neurons. On the other hand, astrocyte activity was characterized by a progressive increase in the resting Ca2+. Notably, at 14 and 28 days post-BCAS, there was an increased expression of anti-inflammatory-related markers (IL-10, S100A10, TRPA1, and Nrf2). These data suggest that, in young mice, BCAS-induced increases in resting Ca2+ were associated with the expression of neuroprotective signals. Contrary to observations in glial cells, vascular function was impaired post-BCAS surgery, as shown by a blunted vasodilatory response to low O2 and the vasodilatory signal, adenosine. Conclusions: Together, these data suggest that, in young mice, BCAS leads to vascular dysfunction (e.g., impaired vasodilation in parenchymal arterioles), and in the absence of neuronal dysfunction, mild ischemia is associated with the activation of glial-derived neuroprotective signals. Full article
Show Figures

Figure 1

Figure 1
<p>Basic membrane properties from cortical pyramidal neurons post-BCAS surgery. Whole-cell patch-clamp recordings from sham, BCAS 14d, and BCAS 28d neurons. (<b>A</b>) Summary data corresponding to resting membrane potential. (<b>B</b>) Summary data corresponding to cell capacitance. (<b>C</b>) Summary data corresponding to input resistance. (<b>D</b>) Input–output function. One-way ANOVA followed by Dunnett’s multiple comparison test (MCT) (<span class="html-italic">n</span> = 27 neurons for sham, <span class="html-italic">n</span> = 38 neurons for BCAS 14d, and <span class="html-italic">n</span> = 30 neurons for BCAS 28d). Data expressed as mean SEM, * <span class="html-italic">p</span> &lt; 0.05 vs. sham.</p>
Full article ">Figure 2
<p>Low O<sub>2</sub>-induced changes in membrane properties of cortical pyramidal neurons from sham, 14, and 28 days post-BCAS surgery mice. (<b>A</b>) Low O<sub>2</sub>-induced changes in resting membrane potential. (<b>B</b>) Proportion of neurons showing depolarization vs. hyperpolarization in response to low O<sub>2</sub>. (<b>C</b>–<b>E</b>) Delta membrane potential resulting from low O<sub>2</sub> exposure to sham, BCAS 14d, and BCAS 28d brain slices. (<b>F</b>,<b>G</b>) Number of action potentials (AP) at various step currents from depolarizing (K) and hyperpolarizing (L) cortical neurons. (<b>A</b>,<b>B</b>) Two-way ANOVA followed by Sidak’s MCT (n-27 sham, <span class="html-italic">n</span> = 38 BCAS 14d, <span class="html-italic">n</span> = 30 BCAS 28d). (<b>C</b>–<b>E</b>) One-way ANOVA, followed by Dunnett’s MCT (C, <span class="html-italic">n</span> = 27 sham, <span class="html-italic">n</span> = 38 BCAS 14, <span class="html-italic">n</span> = 30 BCAS 28d; D/E, <span class="html-italic">n</span> = 18/9 sham, <span class="html-italic">n</span> = 30/8 BCAS 14, <span class="html-italic">n</span> = 24/6 BCAS 28d) (<b>F</b>,<b>G</b>) The mixed-effects model was followed by Dunnett’s MCT (<span class="html-italic">n</span> = 16/11 sham, <span class="html-italic">n</span> = 30/8 BCAS 14d, <span class="html-italic">n</span> = 30/6 BCAS 28d). Data expressed as mean SEM. * <span class="html-italic">p</span> &lt; 0.05, ** or <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01, *** or <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001, ns = not significant. Symbols for groups showing significances are * BCAS 14d, and <sup>#</sup> BCAS 28d.</p>
Full article ">Figure 3
<p>Low O<sub>2</sub>-induced changes in parenchymal arteriole vasoreactivity. (<b>A</b>) Differential interference contrast image of a cannulated and perfused parenchymal arteriole in a brain slice preparation. (<b>B</b>) Vascular reactivity to bath applied low O<sub>2</sub> treatment in sham, BCAS 14d, and BCAS 28d mice. Two-way ANOVA followed by Sidak’s MCT (<span class="html-italic">n</span> = 6 sham, <span class="html-italic">n</span> = 7 BCAS 14d, <span class="html-italic">n</span> = 7 BCAS 28d). Data expressed as mean SEM. *** <span class="html-italic">p</span> &lt; 0.001 and **** <span class="html-italic">p</span> &lt; 0.0001. Scale bar = 20 µm.</p>
Full article ">Figure 4
<p>Parenchymal arteriole responses to adenosine and 10 mM K<sup>+</sup>. (<b>A</b>) Summary data showing percent (%) relaxation of pressurized parenchymal arterioles responses to increasing concentrations of adenosine in sham (<span class="html-italic">n</span> = 6), BCAS 14d (<span class="html-italic">n</span> = 8), and BCAS 28d (<span class="html-italic">n</span> = 7) mice. Two-way ANOVA followed by Dunnett’s MCT (between group comparisons * sham vs BCAS 14d, τ sham vs. BCAS 28d). (<b>B</b>) Summary data showing percent (%) relaxation of pressurized parenchymal arterioles to K<sup>+</sup> in sham (<span class="html-italic">n</span> = 6), BCAS 14d (<span class="html-italic">n</span> = 8), and BCAS 28d (<span class="html-italic">n</span> = 6) mice. One-way ANOVA followed by Dunnett’s MCT vs. sham. Data expressed as means ± SEM. * <span class="html-italic">p</span> &lt; 0.05, **<sup>/ττ</sup> <span class="html-italic">p</span> &lt; 0.01 and <sup>ττττ</sup> <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 5
<p>Spontaneous and low O<sub>2</sub>-induced cortical astrocytic Ca<sup>2+</sup> dynamics post-BCAS ex vivo. (<b>A</b>) Summary data showing astrocyte Ca<sup>2+</sup> events in response to low O<sub>2</sub> in sham, BCAS 14d, and BCAS 28d mice. (<b>B</b>) Proportion of astrocytes responding with an activation or inhibition of Ca<sup>2+</sup> events to low O<sub>2</sub>. (<b>C</b>,<b>D</b>) Summary data showing low O2-induced changes in Ca<sup>2+</sup> events for activated (<b>C</b>) and inhibited (<b>D</b>) astrocytes. (<b>E</b>,<b>F</b>) Summary data showing maximum delta F/F<sub>0</sub> (<b>E</b>) and average delta F/F<sub>0</sub> (<b>F</b>). (<b>A</b>,<b>B</b>,<b>E</b>,<b>F</b>) Two-way ANOVA followed by Sidak’s MCT (sham (<span class="html-italic">n</span> = 32), BCAS 14d (<span class="html-italic">n</span> = 30) and BCAS 28d (<span class="html-italic">n</span> = 18)). (<b>C</b>,<b>D</b>) Two-way ANOVA followed by Sidak’s MCT (sham (<span class="html-italic">n</span> = 9/23), BCAS 14d (<span class="html-italic">n</span> = 10/19) and BCAS 28d (<span class="html-italic">n</span> = 6/12)). Data expressed as mean SEM. * or <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05, ** or <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 and **** <span class="html-italic">p</span> &lt; 0.0001. (*) Within-group comparisons, (#) between-group comparisons.</p>
Full article ">Figure 6
<p>Astrocyte network Ca<sup>2+</sup> activity changes in response to low O<sub>2</sub> post-BCAS surgery. (<b>A</b>) Representative confocal image of multiple GCaMP6f labeled astrocytes in a brain slice. (<b>B</b>) Summary data showing spatial density changes before and after low O<sub>2</sub> treatment. (<b>C</b>,<b>D</b>) Summary data showing temporal density (<b>C</b>) and temporal density with similar size events (<b>D</b>) before and after low O<sub>2</sub> treatment in sham (<span class="html-italic">n</span> = 32), BCAS 14d (<span class="html-italic">n</span> = 29), and BCAS 28d (n-18) mice. Two-way ANOVA followed by Sidak’s MCT. Data expressed as means ± SEM. <sup>#</sup> <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. (*) Within-group comparisons, (#) between-group comparisons.</p>
Full article ">Figure 7
<p>Microglia morphological changes post-BCAS surgery. (<b>A</b>) Representative immunofluorescence confocal images of cortical microglia labeled with Iba1 before and after skeleton analysis used for structural quantification. (<b>B</b>) Summary data showing the quantification of microglia arborization properties in sham (<span class="html-italic">n</span> = 12), BCAS 14d (<span class="html-italic">n</span> = 12), and BCAS 28d (<span class="html-italic">n</span> = 12) mice brain slices. One-way ANOVA followed by Holm–Sidak’s MCT. Data expressed as means ± SEM. * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 8
<p>Quantitative mRNA expression of inflammatory markers in BCAS brains. (<b>A</b>) Summary data showing fold changes in various inflammatory markers for astrocytes and microglia. (<b>B</b>) Summary data showing fold changes in TRPA1 expression level. (<b>C</b>) Summary data showing fold changes in Nrf2 expression level. One sample t and Wilcoxon test (<span class="html-italic">n</span> = 5 per group). Data expressed as means ± SEM. ** <span class="html-italic">p</span> &lt; 0.014. ns = not significant.</p>
Full article ">
15 pages, 1937 KiB  
Article
Improving the Performance of Electrotactile Brain–Computer Interface Using Machine Learning Methods on Multi-Channel Features of Somatosensory Event-Related Potentials
by Marija Novičić, Olivera Djordjević, Vera Miler-Jerković, Ljubica Konstantinović and Andrej M. Savić
Sensors 2024, 24(24), 8048; https://doi.org/10.3390/s24248048 - 17 Dec 2024
Viewed by 558
Abstract
Traditional tactile brain–computer interfaces (BCIs), particularly those based on steady-state somatosensory–evoked potentials, face challenges such as lower accuracy, reduced bit rates, and the need for spatially distant stimulation points. In contrast, using transient electrical stimuli offers a promising alternative for generating tactile BCI [...] Read more.
Traditional tactile brain–computer interfaces (BCIs), particularly those based on steady-state somatosensory–evoked potentials, face challenges such as lower accuracy, reduced bit rates, and the need for spatially distant stimulation points. In contrast, using transient electrical stimuli offers a promising alternative for generating tactile BCI control signals: somatosensory event-related potentials (sERPs). This study aimed to optimize the performance of a novel electrotactile BCI by employing advanced feature extraction and machine learning techniques on sERP signals for the classification of users’ selective tactile attention. The experimental protocol involved ten healthy subjects performing a tactile attention task, with EEG signals recorded from five EEG channels over the sensory–motor cortex. We employed sequential forward selection (SFS) of features from temporal sERP waveforms of all EEG channels. We systematically tested classification performance using machine learning algorithms, including logistic regression, k-nearest neighbors, support vector machines, random forests, and artificial neural networks. We explored the effects of the number of stimuli required to obtain sERP features for classification and their influence on accuracy and information transfer rate. Our approach indicated significant improvements in classification accuracy compared to previous studies. We demonstrated that the number of stimuli for sERP generation can be reduced while increasing the information transfer rate without a statistically significant decrease in classification accuracy. In the case of the support vector machine classifier, we achieved a mean accuracy over 90% for 10 electrical stimuli, while for 6 stimuli, the accuracy decreased by less than 7%, and the information transfer rate increased by 60%. This research advances methods for tactile BCI control based on event-related potentials. This work is significant since tactile stimulation is an understudied modality for BCI control, and electrically induced sERPs are the least studied control signals in reactive BCIs. Exploring and optimizing the parameters of sERP elicitation, as well as feature extraction and classification methods, is crucial for addressing the accuracy versus speed trade-off in various assistive BCI applications where the tactile modality may have added value. Full article
Show Figures

Figure 1

Figure 1
<p>Feature maps following the feature selection process. Figures (<b>A</b>–<b>E</b>) represent the number of times each feature was selected across all subjects and tested values of <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>N</mi> </mrow> <mrow> <mi>a</mi> <mi>v</mi> <mi>g</mi> </mrow> </msub> </mrow> </semantics></math>, with darker colors indicating higher selection frequencies. Figure (<b>F</b>) shows the number of classifiers that selected each feature, where white indicates no selection and black indicates selection by all classifiers.</p>
Full article ">Figure 2
<p>Accuracy of each classifier across all subjects in relation to the value <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>N</mi> </mrow> <mrow> <mi mathvariant="normal">a</mi> <mi mathvariant="normal">v</mi> <mi mathvariant="normal">g</mi> </mrow> </msub> </mrow> </semantics></math>. The dashed line represents the mean value of accuracy for all subjects, and the shaded area indicates the standard deviation interval.</p>
Full article ">Figure 3
<p>Intervals of statistical difference of each classifier across all subjects in relation to the value <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>N</mi> </mrow> <mrow> <mi mathvariant="normal">a</mi> <mi mathvariant="normal">v</mi> <mi mathvariant="normal">g</mi> </mrow> </msub> </mrow> </semantics></math>. Green squares represent intervals where there is a statistically significant difference, and black squares are in correlation with intervals with no significant difference.</p>
Full article ">Figure 4
<p>Information transfer rate in relation to <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>N</mi> </mrow> <mrow> <mi mathvariant="normal">a</mi> <mi mathvariant="normal">v</mi> <mi mathvariant="normal">g</mi> </mrow> </msub> </mrow> </semantics></math>.</p>
Full article ">
11 pages, 3096 KiB  
Article
The Impact of Selective Spatial Attention on Auditory–Tactile Integration: An Event-Related Potential Study
by Weichao An, Nan Zhang, Shengnan Li, Yinghua Yu, Jinglong Wu and Jiajia Yang
Brain Sci. 2024, 14(12), 1258; https://doi.org/10.3390/brainsci14121258 - 15 Dec 2024
Viewed by 588
Abstract
Background: Auditory–tactile integration is an important research area in multisensory integration. Especially in special environments (e.g., traffic noise and complex work environments), auditory–tactile integration is crucial for human response and decision making. We investigated the influence of attention on the temporal course and [...] Read more.
Background: Auditory–tactile integration is an important research area in multisensory integration. Especially in special environments (e.g., traffic noise and complex work environments), auditory–tactile integration is crucial for human response and decision making. We investigated the influence of attention on the temporal course and spatial distribution of auditory–tactile integration. Methods: Participants received auditory stimuli alone, tactile stimuli alone, and simultaneous auditory and tactile stimuli, which were randomly presented on the left or right side. For each block, participants attended to all stimuli on the designated side and detected uncommon target stimuli while ignoring all stimuli on the other side. Event-related potentials (ERPs) were recorded via 64 scalp electrodes. Integration was quantified by comparing the response to the combined stimulus to the sum of the responses to the auditory and tactile stimuli presented separately. Results: The results demonstrated that compared to the unattended condition, integration occurred earlier and involved more brain regions in the attended condition when the stimulus was presented in the left hemispace. The unattended condition involved a more extensive range of brain regions and occurred earlier than the attended condition when the stimulus was presented in the right hemispace. Conclusions: Attention can modulate auditory–tactile integration and show systematic differences between the left and right hemispaces. These findings contribute to the understanding of the mechanisms of auditory–tactile information processing in the human brain. Full article
(This article belongs to the Section Cognitive, Social and Affective Neuroscience)
Show Figures

Figure 1

Figure 1
<p>Experimental paradigm. (<b>a</b>) Sequence of events and their duration. (<b>b</b>) Stimulus conditions: There were 12 stimulus conditions in total: 8 unisensory stimuli (4 standard stimuli and 4 target stimuli) and 4 multisensory stimuli (2 standard stimuli and 2 target stimuli). The black dots in the tactile stimulus column indicate that the probe was raised.</p>
Full article ">Figure 2
<p>Event-related potentials (ERPs) of standard stimulus. The figure shows the grand average of ERPs for unisensory stimulus summation (blue trace) and simultaneous auditory and somatosensory stimulation (red trace) in 20 participants. ERPs are shown at central electrode positions (AFz, FCz, and CPz) and lateral electrode positions (C3, CP5, C4, and CP6). (<b>a</b>) Stimulus presented in the left half-space; (<b>b</b>) stimulus presented in the right half-space. The shaded gray part indicates a significant difference in stimulus type.</p>
Full article ">Figure 3
<p>Influence of attention on topographic voltage distribution. The topographic voltage distributions of the grand average event-related potential (ERP) components for the standard stimulus in the 4 time windows of interest for attended and unattended stimuli when the stimuli were presented in the left and right hemispaces, respectively. The maps show the mean voltage of AT − (A + T) within the corresponding time windows (70–90 ms, 90–110 ms, 110–130 ms, and 180–220 ms). A, auditory; T, tactile; AT, auditory–tactile. Black circles indicate electrodes with significant differences in stimulus type.</p>
Full article ">
19 pages, 871 KiB  
Article
Stirred Not Shaken: A Longitudinal Pilot Study of Head Kinematics and Cognitive Changes in Horseracing
by Emma Edwards, Bert Bond, Timothy P. Holsgrove, Jerry Hill, Ryan Baker and Genevieve K. R. Williams
Vibration 2024, 7(4), 1171-1189; https://doi.org/10.3390/vibration7040060 - 27 Nov 2024
Viewed by 901
Abstract
The purpose of this longitudinal pilot study was to add to the body of research relating to head kinematics/vibration in sport and their potential to cause short-term alterations in brain function. In horseracing, due to the horse’s movement, repeated low-level accelerations are transmitted [...] Read more.
The purpose of this longitudinal pilot study was to add to the body of research relating to head kinematics/vibration in sport and their potential to cause short-term alterations in brain function. In horseracing, due to the horse’s movement, repeated low-level accelerations are transmitted to the jockey’s head. To measure this, professional jockeys (2 male, 2 female) wore an inertial measurement unit (IMU) to record their head kinematics while riding out. In addition, a short battery of tests (Stroop, Trail Making Test B, choice reaction time, manual dexterity, and visual function) was completed immediately before and after riding. Pre- and post-outcome measures from the cognitive test battery were compared using descriptive statistics. The average head kinematics measured across all jockeys and days were at a low level: resultant linear acceleration peak = 5.82 ± 1.08 g, mean = 1.02 ± 0.01 g; resultant rotational velocity peak = 10.37 ± 3.23 rad/s, mean = 0.85 ± 0.15 rad/s; and resultant rotational acceleration peak = 1495 ± 532.75 rad/s2, mean = 86.58 ± 15.54 rad/s2. The duration of an acceleration event was on average 127.04 ± 17.22 ms for linear accelerations and 89.42 ± 19.74 ms for rotational accelerations. This was longer than those noted in many impact and non-impact sports. Jockeys experienced high counts of linear and rotational head accelerations above 3 g and 400 rad/s2, which are considered normal daily living levels (average 300 linear and 445 rotational accelerations per hour of riding). No measurable decline in executive function or dexterity was found after riding; however, a deterioration in visual function (near point convergence and accommodation) was seen. This work lays the foundation for future large-scale research to monitor the head kinematics of riders, measure the effects and understand variables that might influence them. Full article
(This article belongs to the Special Issue Vibrations in Sports)
Show Figures

Figure 1

Figure 1
<p>Schematic of the typical workload in a year and a typical week during the season for a UK professional flat jockey.</p>
Full article ">Figure 2
<p>Position of the inertial measurement unit attached on the mastoid process behind a jockey’s right ear.</p>
Full article ">Figure 3
<p>Heat maps of average daily exposure to head accelerations at the three time points, yellow asterisks (*) mark the peak values measured on those days. (<b>a</b>) Linear head acceleration counts above 3 g where the x-axis is the acceleration in g and the y-axis is each day of the data collection; (<b>b</b>) Rotational head acceleration counts above 400 rad/s, where the x-axis is rotational acceleration in rad/s<sup>2</sup> and the y-axis is each day of the data collection.</p>
Full article ">
21 pages, 1716 KiB  
Article
AI-Driven Neuro-Monitoring: Advancing Schizophrenia Detection and Management Through Deep Learning and EEG Analysis
by Elena-Anca Paraschiv, Lidia Băjenaru, Cristian Petrache, Ovidiu Bica and Dragoș-Nicolae Nicolau
Future Internet 2024, 16(11), 424; https://doi.org/10.3390/fi16110424 - 16 Nov 2024
Cited by 1 | Viewed by 1217
Abstract
Schizophrenia is a complex neuropsychiatric disorder characterized by disruptions in brain connectivity and cognitive functioning. Continuous monitoring of neural activity is essential, as it allows for the detection of subtle changes in brain connectivity patterns, which could provide early warnings of cognitive decline [...] Read more.
Schizophrenia is a complex neuropsychiatric disorder characterized by disruptions in brain connectivity and cognitive functioning. Continuous monitoring of neural activity is essential, as it allows for the detection of subtle changes in brain connectivity patterns, which could provide early warnings of cognitive decline or symptom exacerbation, ultimately facilitating timely therapeutic interventions. This paper proposes a novel approach for detecting schizophrenia-related abnormalities using deep learning (DL) techniques applied to electroencephalogram (EEG) data. Using an openly available EEG dataset on schizophrenia, the focus is on preprocessed event-related potentials (ERPs) from key electrode sites and applied transfer entropy (TE) analysis to quantify the directional flow of information between brain regions. TE matrices were generated to capture neural connectivity patterns, which were then used as input for a hybrid DL model, combining convolutional neural networks (CNNs) and Bidirectional Long Short-Term Memory (BiLSTM) networks. The model achieved a performant accuracy of 99.94% in classifying schizophrenia-related abnormalities, demonstrating its potential for real-time mental health monitoring. The generated TE matrices revealed significant differences in connectivity between the two groups, particularly in frontal and central brain regions, which are critical for cognitive processing. These findings were further validated by correlating the results with EEG data obtained from the Muse 2 headband, emphasizing the potential for portable, non-invasive monitoring of schizophrenia in real-world settings. The final model, integrated into the NeuroPredict platform, offers a scalable solution for continuous mental health monitoring. By incorporating EEG data, heart rate, sleep patterns, and environmental metrics, NeuroPredict facilitates early detection and personalized interventions for schizophrenia patients. Full article
(This article belongs to the Special Issue eHealth and mHealth)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Heatmap representation of directional connectivity between brain regions based on TE values (unitless)<span class="html-italic">:</span> (<b>a</b>) schizophrenia and (<b>b</b>) HCs.</p>
Full article ">Figure 2
<p>Training and validation accuracy (<b>a</b>) and loss (<b>b</b>) plots for the proposed model.</p>
Full article ">Figure 3
<p>The classification report.</p>
Full article ">Figure 4
<p>The proposed integration of the DL model into the NeuroPredict platform.</p>
Full article ">
12 pages, 3245 KiB  
Article
Enhancing Epilepsy Seizure Detection Through Advanced EEG Preprocessing Techniques and Peak-to-Peak Amplitude Fluctuation Analysis
by Muawiyah A. Bahhah and Eyad Talal Attar
Diagnostics 2024, 14(22), 2525; https://doi.org/10.3390/diagnostics14222525 - 12 Nov 2024
Viewed by 898
Abstract
Objectives: Naturally, there are several challenges, such as muscular artifacts, ocular movements and electrical interferences that depend on precise diagnosis and classification, which hamper exact epileptic seizure detection. This study has been conducted to improve seizure detection accuracy in epilepsy patients using an [...] Read more.
Objectives: Naturally, there are several challenges, such as muscular artifacts, ocular movements and electrical interferences that depend on precise diagnosis and classification, which hamper exact epileptic seizure detection. This study has been conducted to improve seizure detection accuracy in epilepsy patients using an advanced preprocessing technique that could remove such noxious artifacts. Methods: In the frame of this paper, the core tool in the area of epilepsy, EEG, will be applied to record and analyze the electrical patterns of the brain. The dataset includes recordings of seven epilepsy patients taken by the Unit of Neurology and Neurophysiology, University of Siena. The preprocessing techniques employed include advanced artifact removal and signal enhancement methods. We introduced Peak-to-Peak Amplitude Fluctuation (PPAF) to assess amplitude variability within Event-Related Potential (ERP) waveforms. This approach was applied to data from patients experiencing 3–5 seizures, categorized into three distinct groups. Results: The results indicated that the frontal and parietal regions, particularly the electrode areas Cz, Pz and Fp2, are the main contributors to epileptic seizures. Additionally, the implementation of the PPAF metric enhanced the effectiveness of seizure detection and classification algorithms, achieving accuracy rates of 99%, 98% and 95% for datasets with three, four and five seizures, respectively. Conclusions: The present research extends the epilepsy diagnosis with clues on brain activity during seizures and further demonstrates the effectiveness of advanced preprocessing techniques. The introduction of PPAF as a metric could have promising potential in improving both the accuracy and reliability of epilepsy seizure detection algorithms. These observations provide important implications for control and treatment both in focal and in generalized epilepsy. Full article
(This article belongs to the Section Medical Imaging and Theranostics)
Show Figures

Figure 1

Figure 1
<p>EEG processing stages in the study.</p>
Full article ">Figure 2
<p>EEG real (−, time acquisition: Power spectrum envelopes and ERP activity of eight most significant independent components.</p>
Full article ">Figure 3
<p>Epilepsy detection for patient with four seizures by examining the IC ERP envelope plots with topographical maps.</p>
Full article ">Figure 4
<p>This figure displays the power spectrum and ERP activity of the most significant independent components. The black traces represent ERP envelopes, with key IC contributions highlighted in color. The scalp topography maps illustrate the spatial distribution of the ICs during seizure events.</p>
Full article ">Figure 5
<p>The ERP and power spectrum for a patient with five seizures. Key independent components (IC 3, 15, 16 and 17) are represented along with their spatial projections.</p>
Full article ">Figure 6
<p>Mean PPAF values for different seizure categories. The percentage of amplitude fluctuations indicates seizure waveform stability across the 3-, 4- and 5-seizure datasets.</p>
Full article ">
15 pages, 1322 KiB  
Systematic Review
Audiovestibular Dysfunction Related to Anti-Phospholipid Syndrome: A Systematic Review
by Jiann-Jy Chen, Chih-Wei Hsu, Yen-Wen Chen, Tien-Yu Chen, Bing-Yan Zeng and Ping-Tao Tseng
Diagnostics 2024, 14(22), 2522; https://doi.org/10.3390/diagnostics14222522 - 11 Nov 2024
Cited by 1 | Viewed by 1468
Abstract
Background: Anti-phospholipid syndrome (APS) has emerged as a significant issue in autoimmune diseases over recent decades. Its hallmark feature is thromboembolic events, potentially affecting any vascularized area including the microcirculation of the inner ear. Since the first case report of APS-related audiovestibular dysfunction [...] Read more.
Background: Anti-phospholipid syndrome (APS) has emerged as a significant issue in autoimmune diseases over recent decades. Its hallmark feature is thromboembolic events, potentially affecting any vascularized area including the microcirculation of the inner ear. Since the first case report of APS-related audiovestibular dysfunction described in 1993, numerous reports have explored the association between APS-related antibodies and audiovestibular dysfunction. These studies indicate a higher prevalence of APS-related antibodies in patients with sensorineural hearing loss compared to healthy controls. Unlike other idiopathic hearing loss disorders, audiovestibular dysfunction associated with APS may respond to appropriate treatments, highlighting the importance of timely recognition by clinicians to potentially achieve favorable outcomes. Therefore, this systematic review aims to consolidate current evidence on the characteristics, pathophysiology, assessment, and management of audiovestibular dysfunction linked to APS. Methods: This systematic review utilized electronic searches of the PubMed, Embase, ClinicalKey, Web of Science, and ScienceDirect online platforms. The initial search was performed on 27 January 2024, with the final update search completed on 20 June 2024. Results: Based on theoretical pathophysiology, anticoagulation emerges as a pivotal treatment strategy. Additionally, drawing from our preliminary data, we propose a modified protocol combining anticoagulants, steroids, and non-invasive brain stimulation to offer clinicians a novel therapeutic approach for managing these symptoms. Conclusions: Clinicians are encouraged to remain vigilant about the possibility of APS and its complex audiovestibular manifestations, as prompt intervention could stabilize audiovestibular function effectively. Full article
(This article belongs to the Special Issue Etiology, Diagnosis, and Treatment of Congenital Hearing Loss)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Flowchart of the whole systematic review procedure.</p>
Full article ">Figure 2
<p>Schematic diagram of the physiopathology of anti-phospholipid syndrome in audiovestibular dysfunction.</p>
Full article ">Figure 3
<p>Flowchart of the modified anticoagulants plus steroid and non-invasive brain stimulation treatment protocol to manage anti-phospholipid syndrome-related audiovestibular dysfunction. Note: this is a proposal of a future study protocol.</p>
Full article ">
18 pages, 2289 KiB  
Article
Expert and Novice Teachers’ Cognitive Neural Differences in Understanding Students’ Classroom Action Intentions
by Yishan Lin, Rui Li, Jesús Ribosa, David Duran and Binghai Sun
Brain Sci. 2024, 14(11), 1080; https://doi.org/10.3390/brainsci14111080 - 29 Oct 2024
Viewed by 969
Abstract
Objectives: Teachers’ intention understanding ability reflects their professional insight, which is the basis for effective classroom teaching activities. However, the cognitive process and brain mechanism of how teachers understand students’ action intention in class are still unclear. Methods: This study used event-related potential [...] Read more.
Objectives: Teachers’ intention understanding ability reflects their professional insight, which is the basis for effective classroom teaching activities. However, the cognitive process and brain mechanism of how teachers understand students’ action intention in class are still unclear. Methods: This study used event-related potential (ERP) technology to explore the cognitive neural differences in intention understanding ability among teachers with different levels of knowledge and experience. The experiment used the comic strips paradigm to examine the ability of expert and novice teachers to understand students’ normative and non-normative classroom actions under different text prompts (“how” and “why”). Results: The results revealed that in the late time window, expert teachers induced larger P300 and LPC amplitudes when they understood students’ classroom action intentions, while the N250 amplitudes induced by novice teachers in the early time window were significantly larger. In addition, for both types of teachers, when understanding the intentions behind students’ normative actions, the N250 amplitude was the most significant, while the P300 and LPC amplitudes were more significant for non-normative actions. Conclusions: This study found that teachers at varying professional development stages had different time processing processes in intention understanding ability, which supported teachers’ brain electrophysiological activities related to social ability. Full article
(This article belongs to the Section Cognitive, Social and Affective Neuroscience)
Show Figures

Figure 1

Figure 1
<p>Examples of experimental materials: (<b>a</b>) normative action; (<b>b</b>) non-normative action.</p>
Full article ">Figure 2
<p>Experimental procedure used in the study. Each trial consisted of a specific sequence: a fixation point was presented for 500 ms, followed by either a “how” or “why” text prompt, which was displayed for 1500 ms in a random order across trials. Subsequently, photographs of students engaged in normative and non-normative actions were presented for 2000 ms. After viewing the photographs, participants made a key press judgment regarding the actions and then rated the comprehensibility of both the text prompt and the student photographs.</p>
Full article ">Figure 3
<p>Expertise level and action type of differential wave in different electrode points.</p>
Full article ">Figure 4
<p>Expertise level and text prompt of differential wave in different electrode points.</p>
Full article ">Figure 5
<p>Action type and text prompt of differential wave in different expertise levels.</p>
Full article ">Figure 6
<p>Topographical maps of expertise level and action type in N250, P300, and LPC. (<b>A</b>) Topographic map of expert teacher under normative action, (<b>B</b>) Topographic map of expert teacher under non-normative action, (<b>C</b>) Topographic map of novice teacher under normative action, (<b>D</b>) Topographic map of novice teacher under non-normative action.</p>
Full article ">
15 pages, 3317 KiB  
Article
Musicianship Modulates Cortical Effects of Attention on Processing Musical Triads
by Jessica MacLean, Elizabeth Drobny, Rose Rizzi and Gavin M. Bidelman
Brain Sci. 2024, 14(11), 1079; https://doi.org/10.3390/brainsci14111079 - 29 Oct 2024
Viewed by 808
Abstract
Background: Many studies have demonstrated the benefits of long-term music training (i.e., musicianship) on the neural processing of sound, including simple tones and speech. However, the effects of musicianship on the encoding of simultaneously presented pitches, in the form of complex musical [...] Read more.
Background: Many studies have demonstrated the benefits of long-term music training (i.e., musicianship) on the neural processing of sound, including simple tones and speech. However, the effects of musicianship on the encoding of simultaneously presented pitches, in the form of complex musical chords, is less well established. Presumably, musicians’ stronger familiarity and active experience with tonal music might enhance harmonic pitch representations, perhaps in an attention-dependent manner. Additionally, attention might influence chordal encoding differently across the auditory system. To this end, we explored the effects of long-term music training and attention on the processing of musical chords at the brainstem and cortical levels. Method: Young adult participants were separated into musician and nonmusician groups based on the extent of formal music training. While recording EEG, listeners heard isolated musical triads that differed only in the chordal third: major, minor, and detuned (4% sharper third from major). Participants were asked to correctly identify chords via key press during active stimulus blocks and watched a silent movie during passive blocks. We logged behavioral identification accuracy and reaction times and calculated information transfer based on the behavioral chord confusion patterns. EEG data were analyzed separately to distinguish between cortical (event-related potential, ERP) and subcortical (frequency-following response, FFR) evoked responses. Results: We found musicians were (expectedly) more accurate, though not faster, than nonmusicians in chordal identification. For subcortical FFRs, responses showed stimulus chord effects but no group differences. However, for cortical ERPs, whereas musicians displayed P2 (~150 ms) responses that were invariant to attention, nonmusicians displayed reduced P2 during passive listening. Listeners’ degree of behavioral information transfer (i.e., success in distinguishing chords) was also better in musicians and correlated with their neural differentiation of chords in the ERPs (but not high-frequency FFRs). Conclusions: Our preliminary results suggest long-term music training strengthens even the passive cortical processing of musical sounds, supporting more automated brain processing of musical chords with less reliance on attention. Our results also suggest that the degree to which listeners can behaviorally distinguish chordal triads is directly related to their neural specificity to musical sounds primarily at cortical rather than subcortical levels. FFR attention effects were likely not observed due to the use of high-frequency stimuli (>220 Hz), which restrict FFRs to brainstem sources. Full article
(This article belongs to the Section Sensory and Motor Neuroscience)
Show Figures

Figure 1

Figure 1
<p><b>Stimulus paradigm to elicit neural and behavioral responses.</b> We used a clustered stimulus paradigm [<a href="#B23-brainsci-14-01079" class="html-bibr">23</a>,<a href="#B27-brainsci-14-01079" class="html-bibr">27</a>,<a href="#B28-brainsci-14-01079" class="html-bibr">28</a>] to elicit FFRs and ERPs within a single trial. For a given trial, the stimulus (a major, minor, or detuned chord) was presented in a rapid stimulus train used to elicit the FFR, followed by a 1500 ms gap of silence and then a single presentation of the chord which elicits an ERP. Following the ERP-evoking stimulus, participants were asked to behaviorally identify the chord via keyboard press.</p>
Full article ">Figure 2
<p><b>Behavioral performance in chord identification for musicians and nonmusicians (active condition).</b> (<b>A</b>) Musicians outperformed nonmusicians in identification accuracy across the board. Nonmusicians showed more gradient performance and were better at identifying minor chords relative to the major and detuned chords. (<b>B</b>) Mean behavioral confusion matrices per group. Cells denote the proportion of responses labeled as a given chord relative to the actual stimulus category presented. Diagonals show correct responses. Chance level = 33%. NMs show more perceptual confusions than Ms, especially between major and detuned chords. (<b>C</b>) Information transfer (IT) derived from the perceptual confusion matrices. IT represents the degree to which listeners’ responses can be accurately predicted given the known input stimulus [<a href="#B52-brainsci-14-01079" class="html-bibr">52</a>]. Values approaching 100% indicate perfect prediction of the response given the input; values approaching 0 indicate total independence of the stimulus and response. IT is near the ceiling and much larger in Ms, indicating higher fidelity differentiation of musical stimuli. Error bars = ±1 S.E.M.</p>
Full article ">Figure 3
<p><b>Group-averaged FFRs to chordal stimuli.</b> (<b>A</b>) Time waveforms to chordal stimuli in musicians and nonmusicians (active condition only). (<b>B</b>) Response spectra (FFTs). FFRs were not influenced by music training. Peaks for the chordal root (1; 220 Hz), third (3*; 262–287 Hz), and fifth (5; 330 Hz) are demarcated. Only the chordal third (*) differed across stimuli. FFRs (collapsed across groups) did not differ between active and passive conditions in the time (<b>C</b>) or frequency (<b>D</b>) domains.</p>
Full article ">Figure 4
<p><b>FFR latency and amplitude characteristics.</b> (<b>A</b>) FFR onset amplitude varied with chord stimulus but not music training. An interaction between condition and stimulus showed stronger amplitudes in the active relative to the passive condition for the major stimulus only (not shown). (<b>B</b>) FFR onset latency and (<b>C</b>) FFR F0 amplitude were invariant both within and between participants. Error bars = ±1 S.E.M.</p>
Full article ">Figure 5
<p><b>Group-averaged ERP waveforms during active and passive conditions</b>. (<b>A</b>) Musicians’ P2 latencies did not differ with attention. (<b>B</b>) Nonmusicians showed stronger P2 responses in the passive compared to active condition.</p>
Full article ">Figure 6
<p><b>ERP latency and amplitude characteristics.</b> (<b>A</b>) P2 latency and (<b>B</b>) amplitudes varied with attention and group. (<b>C</b>) N1 latency was invariant. (<b>D</b>) N1 amplitudes showed a stronger negativity with attention.</p>
Full article ">Figure 7
<p><b>Brain–behavior relation in the differentiation of musical chords.</b> (<b>A</b>) ERP distance between chords (Euclidian distance between all pairwise N1-P2 amplitudes) is larger in musicians, indicating more salient neural responses across triads. (<b>B</b>) ERP distance between chords correlates with behavioral IT, indicating a brain–behavior correspondence between perceptual and neural (cortical) differentiation. No such brain–behavior correspondence was observed for brainstem FFRs. Error bars: ±1 S.E.M.</p>
Full article ">
31 pages, 3352 KiB  
Review
Using Zebrafish to Study Multiciliated Cell Development and Disease States
by Thanh Khoa Nguyen, Sophia Baker, John-Michael Rodriguez, Liana Arceri and Rebecca A. Wingert
Cells 2024, 13(21), 1749; https://doi.org/10.3390/cells13211749 - 23 Oct 2024
Viewed by 1243
Abstract
Multiciliated cells (MCCs) serve many important functions, including fluid propulsion and chemo- and mechanosensing. Diseases ranging from rare conditions to the recent COVID-19 global health pandemic have been linked to MCC defects. In recent years, the zebrafish has emerged as a model to [...] Read more.
Multiciliated cells (MCCs) serve many important functions, including fluid propulsion and chemo- and mechanosensing. Diseases ranging from rare conditions to the recent COVID-19 global health pandemic have been linked to MCC defects. In recent years, the zebrafish has emerged as a model to investigate the biology of MCCs. Here, we review the major events in MCC formation including centriole biogenesis and basal body docking. Then, we discuss studies on the role of MCCs in diseases of the brain, respiratory, kidney and reproductive systems, as well as recent findings about the link between MCCs and SARS-CoV-2. Next, we explore why the zebrafish is a useful model to study MCCs and provide a comprehensive overview of previous studies of genetic components essential for MCC development and motility across three major tissues in the zebrafish: the pronephros, brain ependymal cells and nasal placode. Taken together, here we provide a cohesive summary of MCC research using the zebrafish and its future potential for expanding our understanding of MCC-related disease states. Full article
(This article belongs to the Collection Feature Papers in ‘Cellular Pathology’)
Show Figures

Figure 1

Figure 1
<p>Different pathways of MCC formation. There are three known pathways of MCC formation: (1) Parental centriole pathway: the parent centriole serves as a template for two to eight centrioles arising from it. These centrioles will eventually mature, disengage and dock to the apical membrane to become basal bodies for MCC formation. (2) Deuterosome-dependent pathway: centrioles arise from the deuterosomes instead of parental centrioles. (3) <span class="html-italic">De novo</span> pathway: neither parental centriole nor deuterosomes are needed for centriole formation. The centriole is thought to arise and mature from the pericentriolar material cloud (PCM). Created with <a href="http://BioRender.com" target="_blank">BioRender.com</a>.</p>
Full article ">Figure 2
<p>MCCs in human diseases. There are many pathological conditions across human tissues that are defective in MCC development and motility or have been linked to MCC-regulating genes. Pathological conditions involving MCCs were recorded in several tissues, such as the upper and lower respiratory systems, nervous system, kidney, oviduct and efferent duct. In the kidney, pathological conditions have been linked to ectopic development of MCCs. Created with <a href="http://BioRender.com" target="_blank">BioRender.com</a>.</p>
Full article ">Figure 3
<p>Zebrafish as a model to study development and disease. Zebrafish host many important traits, allowing them to be a great biomedical model to study embryogenesis and model disease states. (1) Transparent embryos: live zebrafish embryos are optically transparent and can be maneuvered to maintain transparency, allowing for development studies. (<b>A</b>) The zebrafish embryo at 24 hpf with almost no pigment across the body. Scale bar = 100 μm. (<b>B</b>) Graph depicts the migration of fluorescent 40 kDa Dextran conjugate in the proximal tubule of live embryo from 24 h post-injection (hpi) to 48 hpi. Scale bar = 100 μm. (2) Simple architecture: zebrafish have simpler organization than other animal models, while still maintaining complexity. For example, the zebrafish pronephros contains only two nephrons while being fully segmented, allowing for effective renal studies. (3) Rapid, external development: zebrafish development and organogenesis happen rapidly, allowing for organogenesis studies. (4) Robust regeneration: zebrafish can regenerate many tissues, allowing for recovery and regeneration studies. (5) Genetic conservation: zebrafish share about 70% of their genes with humans, allowing for studies of genetic diseases. (6) High fecundity: zebrafish are highly productive and can breed year-round in good conditions. Created with <a href="http://BioRender.com" target="_blank">BioRender.com</a>.</p>
Full article ">Figure 4
<p>MCC tissues in the zebrafish. MCCs appear in many tissues in the zebrafish, such as the brain ependymal cells, nasal placode and pronephros. (1) Pronephros: the zebrafish pronephros is segmented into the glomerulus (G), neck (N), proximal convoluted tubule (PCT), proximal straight tubule (PST), distal early (DE), corpuscles of Stannius (CS), distal late (DL) and pronephric duct (PD). At 24 hpf or 28 ss, MCCs are distributed in a salt-and-pepper manner from the caudal end of the PCT to the rostral region of the DE. MCCs were detected as early as 20 ss. (<b>A</b>) 24 hpf WISH with MCC marker <span class="html-italic">odf3b</span> to demonstrate MCCs. Scale bar = 100 μm, inset = 50 μm (<b>B</b>) 28 hpf whole-mount immunofluorescence for acetylated α-tubulin (cilia, green), γ-tubulin (basal bodies, red) and DAPI (nucleus, blue) in the proximal pronephros of WT embryos. The white dash depicts MCC bundles. Scale bar = 50 μm. (2) Nasal placode: MCCs are located in the lateral rim of the nasal placode. The signal of Gmnc, the master regulator of MCC formation, was detected in the nasal placode as early as 18 hpf. (3) Brain ependymal cells: MCCs are detected around 28 hpf and highly enriched near the midline of the tela choroida, the epithelial layer above the dorsal telencephalon (Tel) and forebrain choroid plexus (ChP). Colored bands depict the time range in which MCCs were reported in each tissue in literature. Created with <a href="http://BioRender.com" target="_blank">BioRender.com</a>.</p>
Full article ">Figure 5
<p>Genetic map of MCC development across zebrafish tissues. Map showing genetic interaction across different genetic factors in MCC development that were reported in the zebrafish. Interactions include activation (black arrows), inhibition (black inhibition arrow) and cooperation between factors (black double-headed arrows). Brown arrows denote factors that have been shown to influence MCC organization. Created with <a href="http://BioRender.com" target="_blank">BioRender.com</a>.</p>
Full article ">Figure 6
<p>Essential genetic factors for MCC motility in zebrafish. Map showing genetic factors that were reported in literature to help regulate aspects of MCC motility, such as beating pattern, amplitude or frequency in zebrafish. Boxes denote factors that were shown to be essential in MCC motility. Dashed box denotes a factor that was found to be important for dynein arm development and potentially important for MCC motility. Created with <a href="http://BioRender.com" target="_blank">BioRender.com</a>.</p>
Full article ">
11 pages, 1976 KiB  
Article
Neurophysiological Correlates of Expert Knowledge: An Event-Related Potential (ERP) Study about Law-Relevant Versus Law-Irrelevant Terms
by Peter Walla, Stefan Kalt and Konrad Lachmayer
Brain Sci. 2024, 14(10), 1029; https://doi.org/10.3390/brainsci14101029 - 17 Oct 2024
Viewed by 1592
Abstract
Background: The evaluation of evidence, which frequently takes the form of scientific evidence, necessitates the input of experts in relevant fields. The results are presented as expert opinions or expert evaluations, which are generally accepted as a reliable representation of the facts. A [...] Read more.
Background: The evaluation of evidence, which frequently takes the form of scientific evidence, necessitates the input of experts in relevant fields. The results are presented as expert opinions or expert evaluations, which are generally accepted as a reliable representation of the facts. A further issue that remains unresolved though is the process of evaluating the expertise and knowledge of an expert in the first instance. In general, earned certificates, grades and other objective criteria are typically regarded as representative documentation to substantiate an expert status. However, there is a possibility that these may not always be sufficiently representative. Objectives: The goal of the present study was to provide evidence that the neural processing of law-relevant and law-irrelevant terms varies significantly between participants who have received training in the field of law (experts) and those who have not (novices). Methods: To this end, changes in brain activity were recorded via electroencephalography (EEG) during visual presentations of terms belonging to five different categories (fake right, democracy, filler word, basic right and rule of law). Event-related potentials (ERPs) were subsequently averaged for each category and subjected to statistical analysis. Results: The results clearly demonstrate that participants trained in law processed fake rights and filler words in a similar manner. Furthermore, both of these conditions elicited different levels of brain activity compared to all law-relevant terms. This was not the case in participants who had not received legal training. The brains of untrained participants processed all five term categories in a strikingly similar manner. In light of prior knowledge regarding language processing, the primary focus was on two distinct electrode locations: one in the left posterior region, and the other in the left frontal region. In both locations, the most prominent differences in brain activity elicited by the aforementioned term categories in law-trained participants occurred approximately 450 milliseconds after stimulus onset. The results were further corroborated by a repeated-measures ANOVA and subsequent t-tests, which also demonstrated the absence of this effect in law-untrained participants. Conclusions: The findings of this study provide empirical evidence that brain activity measurements, in particular ERPs, can be used to distinguish between experts trained in a specific field of expertise and novices in that field. Such findings have the potential to facilitate objective assessments of expertise, enabling comparisons between experts and novices that extend beyond traditional criteria such as qualifications and experience. Instead, individuals can be evaluated based on their cognitive processes, as observed through brain activity. Full article
(This article belongs to the Special Issue EEG and Event-Related Potentials)
Show Figures

Figure 1

Figure 1
<p>Event-related potentials (ERPs) calculated for all 5 term categories for the left parietal electrode position P7 in both groups (law-trained and law-untrained participants). Statistical analysis revealed that the time window around 450 ms post stimulus (from 443 ms to 459 ms after stimulus onset) showed the most dominant effects in law-trained participants. This time window is marked with a light blue bar across all ERPs. Note that in law-trained participants, both fake rights and filler words elicited similar brain activity levels (potentials) that differ in very similar ways from those of all 3 law-relevant terms that elicited very similar activities. This pattern of brain activities is totally absent in law-untrained participants.</p>
Full article ">Figure 2
<p>Event-related potentials (ERPs) calculated for all 5 term categories for the left frontal electrode position FC5 in both groups (law-trained and law-untrained participants). Statistical analysis revealed that the time window around 450 ms post stimulus (from 443 ms to 459 ms after stimulus onset) showed the most dominant effects in law-trained participants. This time window is marked with a light blue bar across all ERPs. Note that in law-trained participants both fake rights and filler words elicited similar brain activity levels (potentials) that differ in very similar ways from those of all 3 law-relevant terms that elicited very similar activities. This pattern of brain activities is totally absent in law-untrained participants. This finding resembles the one described for the left parietal electrode position P7, with just smaller activity differences (that also started a bit later) between fake rights and filler words and the rest of the term conditions.</p>
Full article ">Figure 3
<p>Topographical maps including data from all electrode locations created for all 5 term conditions for the time point 450 ms post stimulus and for both groups. Red boxes mark the topographies for the two critical term categories that elicited similar brain activity levels in law-trained brains (fake rights and filler words).</p>
Full article ">
Back to TopTop