[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
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
remove_circle_outline

Search Results (1,550)

Search Parameters:
Keywords = circadian rhythm

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
12 pages, 758 KiB  
Article
New Insights into Health Conditions Related to Malfunctions in Clock Genes
by Kaja Majewska, Mikołaj Seremak, Katarzyna Podhorodecka, Maria Derkaczew, Bartosz Kędziora, Paulina Boniecka, Kamila Zglejc-Waszak, Agnieszka Korytko, Małgorzata Pawłowicz and Joanna Wojtkiewicz
Biomolecules 2024, 14(10), 1282; https://doi.org/10.3390/biom14101282 - 11 Oct 2024
Viewed by 367
Abstract
Chronotypes play a crucial role in regulating sleep–wake cycles and overall health. The aim of this study was to investigate chronotype, sleep quality, polymorphisms of clock genes and the level of leptin in serum. We used standardized questionnaires to assess chronotype and sleep [...] Read more.
Chronotypes play a crucial role in regulating sleep–wake cycles and overall health. The aim of this study was to investigate chronotype, sleep quality, polymorphisms of clock genes and the level of leptin in serum. We used standardized questionnaires to assess chronotype and sleep quality. Genetic analysis was performed to determine the selected clock gene polymorphism. Serum leptin level was measured by the Elisa method. The results showed that serum leptin concentration was elevated in women, as well as in men who had a high waist-to-hip ratio (WHR) and body mass index (BMI). The evidence indicated that younger students (<22 years old) were most likely to experience poor sleep quality. Nevertheless, our multivariate analysis revealed that young age and a morning-oriented chronotype were associated with better sleep quality. We noted that clock gene polymorphisms were present in 28.6% of the participants. Moreover, polymorphisms of PER1 c.2247C>T (rs2735611) and PER2 c.-12C>G (rs2304672) genes were associated with serum leptin level and chronotype, respectively. These findings provide insights into the relationships between chronotype, sleep quality, clock gene polymorphisms and obesity risk in biomedical students. Understanding these factors can contribute to better sleep management and potential interventions to improve health outcomes in humans. Full article
Show Figures

Figure 1

Figure 1
<p>The figure shows risk factors associated with poor sleep quality (* age &lt; 22).</p>
Full article ">
28 pages, 1968 KiB  
Review
The Influence of Circadian Rhythms on DNA Damage Repair in Skin Photoaging
by Zhi Su, Qianhua Hu, Xiang Li, Zirun Wang and Ying Xie
Int. J. Mol. Sci. 2024, 25(20), 10926; https://doi.org/10.3390/ijms252010926 - 11 Oct 2024
Viewed by 1080
Abstract
Circadian rhythms, the internal timekeeping systems governing physiological processes, significantly influence skin health, particularly in response to ultraviolet radiation (UVR). Disruptions in circadian rhythms can exacerbate UVR-induced skin damage and increase the risk of skin aging and cancer. This review explores how circadian [...] Read more.
Circadian rhythms, the internal timekeeping systems governing physiological processes, significantly influence skin health, particularly in response to ultraviolet radiation (UVR). Disruptions in circadian rhythms can exacerbate UVR-induced skin damage and increase the risk of skin aging and cancer. This review explores how circadian rhythms affect various aspects of skin physiology and pathology, with a special focus on DNA repair. Circadian regulation ensures optimal DNA repair following UVR-induced damage, reducing mutation accumulation, and enhancing genomic stability. The circadian control over cell proliferation and apoptosis further contributes to skin regeneration and response to UVR. Oxidative stress management is another critical area where circadian rhythms exert influence. Key circadian genes like brain and muscle ARNT-like 1 (BMAL1) and circadian locomotor output cycles kaput (CLOCK) modulate the activity of antioxidant enzymes and signaling pathways to protect cells from oxidative stress. Circadian rhythms also affect inflammatory and immune responses by modulating the inflammatory response and the activity of Langerhans cells and other immune cells in the skin. In summary, circadian rhythms form a complex defense network that manages UVR-induced damage through the precise regulation of DNA damage repair, cell proliferation, apoptosis, inflammatory response, oxidative stress, and hormonal signaling. Understanding these mechanisms provides insights into developing targeted skin protection and improving skin cancer prevention. Full article
(This article belongs to the Special Issue Molecular Mechanism in DNA Replication and Repair)
Show Figures

Figure 1

Figure 1
<p>Key pathways in circadian modulation of UVR-induced skin damage.</p>
Full article ">Figure 2
<p>Circadian regulation of cytokine and chemokine production in keratinocytes.</p>
Full article ">Figure 3
<p>Overview of circadian rhythms in UVR-induced skin damage and repair.</p>
Full article ">
22 pages, 4935 KiB  
Systematic Review
The Effects of Time-Restricted Eating on Fat Loss in Adults with Overweight and Obese Depend upon the Eating Window and Intervention Strategies: A Systematic Review and Meta-Analysis
by Yixun Xie, Kaixiang Zhou, Zhangyuting Shang, Dapeng Bao and Junhong Zhou
Nutrients 2024, 16(19), 3390; https://doi.org/10.3390/nu16193390 - 5 Oct 2024
Viewed by 2012
Abstract
Time-restricted eating (TRE) is a circadian rhythm-based intermittent fasting intervention that has been used to treat obesity. However, the efficacy and safety of TRE for fat loss have not been comprehensively examined and the influences of TRE characteristics on such effects are unknown. [...] Read more.
Time-restricted eating (TRE) is a circadian rhythm-based intermittent fasting intervention that has been used to treat obesity. However, the efficacy and safety of TRE for fat loss have not been comprehensively examined and the influences of TRE characteristics on such effects are unknown. This systematic review and meta-analysis comprehensively characterized the efficacy and safety of TRE for fat loss in adults with overweight and obese, and it explored the influence of TRE characteristics on this effect. Methods: A search strategy based on the PICOS principle was used to find relevant publications in seven databases. The outcomes were body composition, anthropometric indicators, and blood lipid metrics. Twenty publications (20 studies) with 1288 participants, covering the period from 2020 to 2024, were included. Results: Compared to the control group, TRE safely and significantly reduced body fat percentage, fat mass, lean mass, body mass, BMI, and waist circumference (MDpooled = −2.14 cm, 95% CI = −2.88~−1.40, p < 0.001), and increased low-density lipoprotein (LDL) (MDpooled = 2.70, 95% CI = 0.17~5.22, p = 0.037), but it did not alter the total cholesterol, high-density lipoprotein, and triglycerides (MDpooled = −1.09~1.20 mg/dL, 95% CI −4.31~5.47, p > 0.05). Subgroup analyses showed that TRE only or TRE-caloric restriction with an eating window of 6 to 8 h may be appropriate for losing body fat and overall weight. Conclusions: This work provides moderate to high evidence that TRE is a promising dietary strategy for fat loss. Although it may potentially reduce lean mass and increase LDL, these effects do not pose significant safety concerns. This trial was registered with PROSPERO as CRD42023406329. Full article
(This article belongs to the Section Nutrition and Obesity)
Show Figures

Figure 1

Figure 1
<p>Flow chart of the publication screening.</p>
Full article ">Figure 2
<p>Risk of bias assessment in the RCTs in the included studies. A total of 20 publications [<a href="#B10-nutrients-16-03390" class="html-bibr">10</a>,<a href="#B11-nutrients-16-03390" class="html-bibr">11</a>,<a href="#B12-nutrients-16-03390" class="html-bibr">12</a>,<a href="#B13-nutrients-16-03390" class="html-bibr">13</a>,<a href="#B21-nutrients-16-03390" class="html-bibr">21</a>,<a href="#B28-nutrients-16-03390" class="html-bibr">28</a>,<a href="#B29-nutrients-16-03390" class="html-bibr">29</a>,<a href="#B30-nutrients-16-03390" class="html-bibr">30</a>,<a href="#B31-nutrients-16-03390" class="html-bibr">31</a>,<a href="#B32-nutrients-16-03390" class="html-bibr">32</a>,<a href="#B33-nutrients-16-03390" class="html-bibr">33</a>,<a href="#B34-nutrients-16-03390" class="html-bibr">34</a>,<a href="#B35-nutrients-16-03390" class="html-bibr">35</a>,<a href="#B36-nutrients-16-03390" class="html-bibr">36</a>,<a href="#B37-nutrients-16-03390" class="html-bibr">37</a>,<a href="#B38-nutrients-16-03390" class="html-bibr">38</a>,<a href="#B39-nutrients-16-03390" class="html-bibr">39</a>,<a href="#B40-nutrients-16-03390" class="html-bibr">40</a>,<a href="#B41-nutrients-16-03390" class="html-bibr">41</a>,<a href="#B42-nutrients-16-03390" class="html-bibr">42</a>] were incorporated into the review.</p>
Full article ">Figure 3
<p>Meta−analysis results of (<b>A</b>) body fat percentage [<a href="#B12-nutrients-16-03390" class="html-bibr">12</a>,<a href="#B13-nutrients-16-03390" class="html-bibr">13</a>,<a href="#B21-nutrients-16-03390" class="html-bibr">21</a>,<a href="#B28-nutrients-16-03390" class="html-bibr">28</a>,<a href="#B29-nutrients-16-03390" class="html-bibr">29</a>,<a href="#B30-nutrients-16-03390" class="html-bibr">30</a>,<a href="#B31-nutrients-16-03390" class="html-bibr">31</a>,<a href="#B33-nutrients-16-03390" class="html-bibr">33</a>,<a href="#B34-nutrients-16-03390" class="html-bibr">34</a>,<a href="#B35-nutrients-16-03390" class="html-bibr">35</a>,<a href="#B36-nutrients-16-03390" class="html-bibr">36</a>,<a href="#B37-nutrients-16-03390" class="html-bibr">37</a>,<a href="#B38-nutrients-16-03390" class="html-bibr">38</a>,<a href="#B39-nutrients-16-03390" class="html-bibr">39</a>,<a href="#B42-nutrients-16-03390" class="html-bibr">42</a>], (<b>B</b>) fat mass [<a href="#B10-nutrients-16-03390" class="html-bibr">10</a>,<a href="#B11-nutrients-16-03390" class="html-bibr">11</a>,<a href="#B12-nutrients-16-03390" class="html-bibr">12</a>,<a href="#B13-nutrients-16-03390" class="html-bibr">13</a>,<a href="#B21-nutrients-16-03390" class="html-bibr">21</a>,<a href="#B28-nutrients-16-03390" class="html-bibr">28</a>,<a href="#B29-nutrients-16-03390" class="html-bibr">29</a>,<a href="#B30-nutrients-16-03390" class="html-bibr">30</a>,<a href="#B31-nutrients-16-03390" class="html-bibr">31</a>,<a href="#B34-nutrients-16-03390" class="html-bibr">34</a>,<a href="#B35-nutrients-16-03390" class="html-bibr">35</a>,<a href="#B36-nutrients-16-03390" class="html-bibr">36</a>,<a href="#B37-nutrients-16-03390" class="html-bibr">37</a>,<a href="#B39-nutrients-16-03390" class="html-bibr">39</a>,<a href="#B40-nutrients-16-03390" class="html-bibr">40</a>,<a href="#B41-nutrients-16-03390" class="html-bibr">41</a>,<a href="#B42-nutrients-16-03390" class="html-bibr">42</a>], and (<b>C</b>) lean mass [<a href="#B10-nutrients-16-03390" class="html-bibr">10</a>,<a href="#B11-nutrients-16-03390" class="html-bibr">11</a>,<a href="#B12-nutrients-16-03390" class="html-bibr">12</a>,<a href="#B13-nutrients-16-03390" class="html-bibr">13</a>,<a href="#B21-nutrients-16-03390" class="html-bibr">21</a>,<a href="#B29-nutrients-16-03390" class="html-bibr">29</a>,<a href="#B30-nutrients-16-03390" class="html-bibr">30</a>,<a href="#B31-nutrients-16-03390" class="html-bibr">31</a>,<a href="#B34-nutrients-16-03390" class="html-bibr">34</a>,<a href="#B35-nutrients-16-03390" class="html-bibr">35</a>,<a href="#B36-nutrients-16-03390" class="html-bibr">36</a>,<a href="#B37-nutrients-16-03390" class="html-bibr">37</a>,<a href="#B40-nutrients-16-03390" class="html-bibr">40</a>,<a href="#B41-nutrients-16-03390" class="html-bibr">41</a>,<a href="#B42-nutrients-16-03390" class="html-bibr">42</a>].</p>
Full article ">Figure 4
<p>Meta−analysis results for (<b>A</b>) body mass [<a href="#B11-nutrients-16-03390" class="html-bibr">11</a>,<a href="#B12-nutrients-16-03390" class="html-bibr">12</a>,<a href="#B13-nutrients-16-03390" class="html-bibr">13</a>,<a href="#B21-nutrients-16-03390" class="html-bibr">21</a>,<a href="#B28-nutrients-16-03390" class="html-bibr">28</a>,<a href="#B29-nutrients-16-03390" class="html-bibr">29</a>,<a href="#B30-nutrients-16-03390" class="html-bibr">30</a>,<a href="#B31-nutrients-16-03390" class="html-bibr">31</a>,<a href="#B33-nutrients-16-03390" class="html-bibr">33</a>,<a href="#B34-nutrients-16-03390" class="html-bibr">34</a>,<a href="#B35-nutrients-16-03390" class="html-bibr">35</a>,<a href="#B36-nutrients-16-03390" class="html-bibr">36</a>,<a href="#B37-nutrients-16-03390" class="html-bibr">37</a>,<a href="#B38-nutrients-16-03390" class="html-bibr">38</a>,<a href="#B39-nutrients-16-03390" class="html-bibr">39</a>,<a href="#B40-nutrients-16-03390" class="html-bibr">40</a>,<a href="#B41-nutrients-16-03390" class="html-bibr">41</a>,<a href="#B42-nutrients-16-03390" class="html-bibr">42</a>], (<b>B</b>) body mass index [<a href="#B12-nutrients-16-03390" class="html-bibr">12</a>,<a href="#B13-nutrients-16-03390" class="html-bibr">13</a>,<a href="#B30-nutrients-16-03390" class="html-bibr">30</a>,<a href="#B31-nutrients-16-03390" class="html-bibr">31</a>,<a href="#B33-nutrients-16-03390" class="html-bibr">33</a>,<a href="#B34-nutrients-16-03390" class="html-bibr">34</a>,<a href="#B35-nutrients-16-03390" class="html-bibr">35</a>,<a href="#B37-nutrients-16-03390" class="html-bibr">37</a>,<a href="#B38-nutrients-16-03390" class="html-bibr">38</a>,<a href="#B39-nutrients-16-03390" class="html-bibr">39</a>,<a href="#B40-nutrients-16-03390" class="html-bibr">40</a>,<a href="#B42-nutrients-16-03390" class="html-bibr">42</a>], and (<b>C</b>) waist circumference [<a href="#B11-nutrients-16-03390" class="html-bibr">11</a>,<a href="#B12-nutrients-16-03390" class="html-bibr">12</a>,<a href="#B13-nutrients-16-03390" class="html-bibr">13</a>,<a href="#B21-nutrients-16-03390" class="html-bibr">21</a>,<a href="#B31-nutrients-16-03390" class="html-bibr">31</a>,<a href="#B33-nutrients-16-03390" class="html-bibr">33</a>,<a href="#B34-nutrients-16-03390" class="html-bibr">34</a>,<a href="#B35-nutrients-16-03390" class="html-bibr">35</a>,<a href="#B38-nutrients-16-03390" class="html-bibr">38</a>,<a href="#B39-nutrients-16-03390" class="html-bibr">39</a>,<a href="#B40-nutrients-16-03390" class="html-bibr">40</a>,<a href="#B42-nutrients-16-03390" class="html-bibr">42</a>].</p>
Full article ">Figure 5
<p>Meta−analysis results for (<b>A</b>) total cholesterol [<a href="#B11-nutrients-16-03390" class="html-bibr">11</a>,<a href="#B12-nutrients-16-03390" class="html-bibr">12</a>,<a href="#B13-nutrients-16-03390" class="html-bibr">13</a>,<a href="#B21-nutrients-16-03390" class="html-bibr">21</a>,<a href="#B28-nutrients-16-03390" class="html-bibr">28</a>,<a href="#B31-nutrients-16-03390" class="html-bibr">31</a>,<a href="#B35-nutrients-16-03390" class="html-bibr">35</a>,<a href="#B37-nutrients-16-03390" class="html-bibr">37</a>,<a href="#B40-nutrients-16-03390" class="html-bibr">40</a>,<a href="#B42-nutrients-16-03390" class="html-bibr">42</a>], (<b>B</b>) high−density lipoprotein cholesterol [<a href="#B10-nutrients-16-03390" class="html-bibr">10</a>,<a href="#B11-nutrients-16-03390" class="html-bibr">11</a>,<a href="#B12-nutrients-16-03390" class="html-bibr">12</a>,<a href="#B13-nutrients-16-03390" class="html-bibr">13</a>,<a href="#B28-nutrients-16-03390" class="html-bibr">28</a>,<a href="#B29-nutrients-16-03390" class="html-bibr">29</a>,<a href="#B31-nutrients-16-03390" class="html-bibr">31</a>,<a href="#B35-nutrients-16-03390" class="html-bibr">35</a>,<a href="#B37-nutrients-16-03390" class="html-bibr">37</a>,<a href="#B40-nutrients-16-03390" class="html-bibr">40</a>,<a href="#B42-nutrients-16-03390" class="html-bibr">42</a>], (<b>C</b>) low−density lipoprotein cholesterol [<a href="#B10-nutrients-16-03390" class="html-bibr">10</a>,<a href="#B11-nutrients-16-03390" class="html-bibr">11</a>,<a href="#B12-nutrients-16-03390" class="html-bibr">12</a>,<a href="#B13-nutrients-16-03390" class="html-bibr">13</a>,<a href="#B28-nutrients-16-03390" class="html-bibr">28</a>,<a href="#B29-nutrients-16-03390" class="html-bibr">29</a>,<a href="#B35-nutrients-16-03390" class="html-bibr">35</a>,<a href="#B37-nutrients-16-03390" class="html-bibr">37</a>,<a href="#B40-nutrients-16-03390" class="html-bibr">40</a>,<a href="#B42-nutrients-16-03390" class="html-bibr">42</a>], and (<b>D</b>) triglycerides [<a href="#B10-nutrients-16-03390" class="html-bibr">10</a>,<a href="#B11-nutrients-16-03390" class="html-bibr">11</a>,<a href="#B12-nutrients-16-03390" class="html-bibr">12</a>,<a href="#B13-nutrients-16-03390" class="html-bibr">13</a>,<a href="#B28-nutrients-16-03390" class="html-bibr">28</a>,<a href="#B29-nutrients-16-03390" class="html-bibr">29</a>,<a href="#B35-nutrients-16-03390" class="html-bibr">35</a>,<a href="#B37-nutrients-16-03390" class="html-bibr">37</a>,<a href="#B40-nutrients-16-03390" class="html-bibr">40</a>,<a href="#B42-nutrients-16-03390" class="html-bibr">42</a>].</p>
Full article ">
18 pages, 3288 KiB  
Article
The Change Rate of the Fbxl21 Gene and the Amino Acid Composition of Its Protein Correlate with the Species-Specific Lifespan in Placental Mammals
by Vassily A. Lyubetsky, Gregory A. Shilovsky, Jian-Rong Yang, Alexandr V. Seliverstov and Oleg A. Zverkov
Biology 2024, 13(10), 792; https://doi.org/10.3390/biology13100792 - 2 Oct 2024
Viewed by 876
Abstract
This article proposes a methodology for establishing a relationship between the change rate of a given gene (relative to a given taxon) together with the amino acid composition of the proteins encoded by this gene and the traits of the species containing this [...] Read more.
This article proposes a methodology for establishing a relationship between the change rate of a given gene (relative to a given taxon) together with the amino acid composition of the proteins encoded by this gene and the traits of the species containing this gene. The methodology is illustrated based on the mammalian genes responsible for regulating the circadian rhythms that underlie a number of human disorders, particularly those associated with aging. The methods used are statistical and bioinformatic ones. A systematic search for orthologues, pseudogenes, and gene losses was performed using our previously developed methods. It is demonstrated that the least conserved Fbxl21 gene in the Euarchontoglires superorder exhibits a statistically significant connection of genomic characteristics (the median of dN/dS for a gene relative to all the other orthologous genes of a taxon, as well as the preference or avoidance of certain amino acids in its protein) with species-specific lifespan and body weight. In contrast, no such connection is observed for Fbxl21 in the Laurasiatheria superorder. This study goes beyond the protein-coding genes, since the accumulation of amino acid substitutions in the course of evolution leads to pseudogenization and even gene loss, although the relationship between the genomic characteristics and the species traits is still preserved. The proposed methodology is illustrated using the examples of circadian rhythm genes and proteins in placental mammals, e.g., longevity is connected with the rate of Fbxl21 gene change, pseudogenization or gene loss, and specific amino acid substitutions (e.g., asparagine at the 19th position of the CRY-binding domain) in the protein encoded by this gene. Full article
Show Figures

Figure 1

Figure 1
<p>Molecular mechanism of circadian rhythms. The processes going on in the nucleus are shown at the top, and the processes in the cytoplasm are shown at the bottom. Dashed lines with arrows, translocation of the corresponding proteins into the nucleus; solid lines with arrows, direct effect, including catalysis; solid line with a blunt end, inhibition; circle with letter “P”, phosphate groups attached to the proteins; broken arrows, promoters; the diagonal crosses indicate protein degradation. Bmal1, basic helix-loop-helix ARNT like 1; CK1, casein kinase 1; Clock, circadian locomotor output cycles kaput protein; Cry, cryptochrome protein; GSK3, glycogen synthase kinase-3; Per, period protein; β-TrCP, beta-transducin repeat containing E3 ubiquitin–protein ligase.</p>
Full article ">Figure 2
<p>In the <span class="html-italic">Fbxl21</span> and <span class="html-italic">Fbxl3</span> genes from the Euarchontoglires superorder, for the species-specific maximal reported lifespan (MRLS) and the common logarithm of body weight, linear regressions (the lines with their estimated uncertainty) for the median and MRLS (<b>a</b>), as well as for the median and the common logarithm of body weight (<b>b</b>), are shown. Each dot represents an <span class="html-italic">Fbxl21</span> or <span class="html-italic">Fbxl3</span> gene from one animal species. The correlation between the considered species traits and the genomic characteristic—the median, which for <span class="html-italic">Fbxl3</span> does not seem reliable enough due to the ambiguity of the regression line and small interval of change in the median, is shown.</p>
Full article ">Figure 2 Cont.
<p>In the <span class="html-italic">Fbxl21</span> and <span class="html-italic">Fbxl3</span> genes from the Euarchontoglires superorder, for the species-specific maximal reported lifespan (MRLS) and the common logarithm of body weight, linear regressions (the lines with their estimated uncertainty) for the median and MRLS (<b>a</b>), as well as for the median and the common logarithm of body weight (<b>b</b>), are shown. Each dot represents an <span class="html-italic">Fbxl21</span> or <span class="html-italic">Fbxl3</span> gene from one animal species. The correlation between the considered species traits and the genomic characteristic—the median, which for <span class="html-italic">Fbxl3</span> does not seem reliable enough due to the ambiguity of the regression line and small interval of change in the median, is shown.</p>
Full article ">Figure 3
<p>Scatter plots of the median and MRLS (<b>a</b>) and the common logarithm of body weight (<b>b</b>) for the <span class="html-italic">Fbxl21</span> and <span class="html-italic">Fbxl3</span> genes from the Euarchontoglires. Each dot represents an <span class="html-italic">Fbxl21</span> or <span class="html-italic">Fbxl3</span> gene from one animal species. The taxa are denoted by shape and color and are listed in the legend. The correlation between these species traits and the median is shown; see the description in <a href="#biology-13-00792-f002" class="html-fig">Figure 2</a>.</p>
Full article ">Figure 3 Cont.
<p>Scatter plots of the median and MRLS (<b>a</b>) and the common logarithm of body weight (<b>b</b>) for the <span class="html-italic">Fbxl21</span> and <span class="html-italic">Fbxl3</span> genes from the Euarchontoglires. Each dot represents an <span class="html-italic">Fbxl21</span> or <span class="html-italic">Fbxl3</span> gene from one animal species. The taxa are denoted by shape and color and are listed in the legend. The correlation between these species traits and the median is shown; see the description in <a href="#biology-13-00792-f002" class="html-fig">Figure 2</a>.</p>
Full article ">Figure 4
<p>Scatter plots of the median and MRLS for the <span class="html-italic">Fbxl21</span> and <span class="html-italic">Fbxl3</span> genes in Cetartiodactyla (as a part of the Laurasiatheria superorder). Each dot represents an <span class="html-italic">Fbxl21</span> or <span class="html-italic">Fbxl3</span> gene from one animal species. The taxa are denoted with the shape and color and are listed in the legend. The correlation between the data on these species traits and the median is shown; see the description from <a href="#biology-13-00792-f002" class="html-fig">Figure 2</a>.</p>
Full article ">Figure 5
<p>The protein tree of Fbxl21 in the anthropoid apes and Cercopithecidae (all species represented in RefSeq that have this protein). In the common chimpanzee and bonobo, the Fbxl21 protein sequences are identical. The node corresponding to the root of the phylogenetic tree is marked with a cross. The “0.03” above the bar represents the scale of the branch lengths, which corresponds to the expected number of substitutions per site. Specifically, a branch length of 0.03 indicates an expected 3% difference in amino acid sequence between the nodes connected by that branch. The peculiarity of <span class="html-italic">Fbxl21</span> gene evolution in apes is shown.</p>
Full article ">Figure 6
<p>(<b>a</b>) The occurrence (%) of amino acids in the Fbxl21 protein, averaged by species. It is similar for Euarchontoglires, Laurasiatheria, and Placentalia. 20 amino acids are listed in the first row. (<b>b</b>) For the three superorders, Euarchontoglires, Laurasiatheria, and all placental mammals, the Pearson correlation coefficient <span class="html-italic">r</span> and the <span class="html-italic">p</span>-value of the amino acid occurrence in Fbxl21 with the species-specific lifespan are shown. (<b>c</b>) shows the common logarithm of the body weight; see (B) for a description. Correlation characteristics of the relationship between the frequency of each amino acid in the Fbxl21 protein in the indicated taxa and the species traits were obtained.</p>
Full article ">Figure 7
<p>Alignment of Cry-binding domain. Almost all species in the list of mammal taxa include the Fbxl21 protein indicated in the rows. The table shows the consensus (54 positions) of the Fbxl21 Cry-binding domain and below the number of species in taxa with the amino acid distinct from the consensus amino acid. The upper part of the table shows all amino acid substitutions; the lower part shows only radical substitutions. The rightmost column shows the number of conservative positions where no substitution occurs, or no radical substitution in a taxon occurs, respectively. The 19th position of the domain is the least conservative.</p>
Full article ">
28 pages, 1447 KiB  
Review
Melatonin’s Impact on Wound Healing
by Eun-Hwa Sohn, Su-Nam Kim and Sung-Ryul Lee
Antioxidants 2024, 13(10), 1197; https://doi.org/10.3390/antiox13101197 - 2 Oct 2024
Viewed by 527
Abstract
Melatonin (5-methoxy-N-acetyltryptamine) is an indoleamine compound that plays a critical role in the regulation of circadian rhythms. While melatonin is primarily synthesized from the amino acid tryptophan in the pineal gland of the brain, it can also be produced locally in various tissues, [...] Read more.
Melatonin (5-methoxy-N-acetyltryptamine) is an indoleamine compound that plays a critical role in the regulation of circadian rhythms. While melatonin is primarily synthesized from the amino acid tryptophan in the pineal gland of the brain, it can also be produced locally in various tissues, such as the skin and intestines. Melatonin’s effects in target tissues can be mediated through receptor-dependent mechanisms. Additionally, melatonin exerts various actions via receptor-independent pathways. In biological systems, melatonin and its endogenous metabolites often produce similar effects. While injuries are common in daily life, promoting optimal wound healing is essential for patient well-being and healthcare outcomes. Beyond regulating circadian rhythms as a neuroendocrine hormone, melatonin may enhance wound healing through (1) potent antioxidant properties, (2) anti-inflammatory actions, (3) infection control, (4) regulation of vascular reactivity and angiogenesis, (5) analgesic (pain-relieving) effects, and (6) anti-pruritic (anti-itch) effects. This review aims to provide a comprehensive overview of scientific studies that demonstrate melatonin’s potential roles in supporting effective wound healing. Full article
Show Figures

Figure 1

Figure 1
<p>The chemical structure of melatonin. Melatonin structure is composed of an indole ring with an acetamide side chain (−CH<sub>2</sub>CONH<sub>2</sub>) and a methoxy side chain (−OCH<sub>3</sub>) on a benzene ring.</p>
Full article ">Figure 2
<p>Biosynthesis of melatonin and its degrading pathway. L-Tryptophan and serotonin are precursors of melatonin. Melatonin synthesis from L-tryptophan sequentially involves four enzyme-catalyzed reactions: hydroxylation, decarboxylation, acetylation, and methylation. The oxidation of melatonin can lead to the formation of various oxidized forms and metabolites. Melatonin undergoes secondary metabolism in the kidney and is excreted in the urine as an inactive metabolite 6-OHM. Abbreviations: AAD: aromatic amino acid decarboxylase; AANAT: Aralkylamine N-acetyltransferase; ASMT: Acetylserotonin O-methyltransferase; AFMK: N1-Acetyl-N2-formyl-5-methoxykynuramine; AMK: N1-Acetyl-5-methoxykynuramine; HIMOT: Hydroxy-O-methyltransferase; TPH1/2: Tryptophan Hydroxylase 1/2; NAT: N-Acetyltransferase; 5-HT: 5-hydroxytryptamine; 6-OHM: 6-Hydroxymelatonin; OH*: hydroxy radical.</p>
Full article ">Figure 3
<p>Potential mechanisms of melatonin action in relation to local concentrations.</p>
Full article ">Figure 4
<p>The roles of melatonin in antioxidant defense, anti-inflammatory response, and mitochondrial protection.</p>
Full article ">Figure 5
<p>The roles of melatonin in vascular function, angiogenesis, neurological regulation, infection control, and oncostatic activity.</p>
Full article ">Figure 6
<p>Melatonin for optimal support of wound healing. Beyond regulating circadian rhythm, melatonin exhibits (1) powerful antioxidant capacities, (2) anti-inflammatory actions, (3) infection control, (4) regulation of vascular reactivity and angiogenesis, (5) analgesic (pain-relieving) effects, (6) anti-itch (anti-pruritic) effects, and (7) oncostatic effects.</p>
Full article ">
16 pages, 3067 KiB  
Article
Response of Native and Non-Native Subarctic Plant Species to Continuous Illumination by Natural and Artificial Light
by Tatjana G. Shibaeva, Elena G. Sherudilo, Alexandra A. Rubaeva, Natalya Yu. Shmakova and Alexander F. Titov
Plants 2024, 13(19), 2742; https://doi.org/10.3390/plants13192742 - 30 Sep 2024
Viewed by 409
Abstract
This study addressed the following questions: How does continuous lighting (CL) impact plant physiology, and photosynthetic and stress responses? Does the impact of CL depend on the source of the light and other environmental factors (natural vs. artificial)? Do responses to CL differ [...] Read more.
This study addressed the following questions: How does continuous lighting (CL) impact plant physiology, and photosynthetic and stress responses? Does the impact of CL depend on the source of the light and other environmental factors (natural vs. artificial)? Do responses to CL differ for native and non-native plant species in the subarctic region and, if differences exist, what physiological reasons might they be associated with them? Experiments were conducted with three plants native to the subarctic region (Geranium sylvaticum L., Geum rivale L., Potentilla erecta (L.) Raeusch.) and three non-native plant species (Geranium himalayense Klotzsch, Geum coccineum Sibth. and Sm., Potentilla atrosanguinea Loddiges ex D. Don) introduced in the Polar-Alpine Botanic Garden (KPABG, 67°38′ N). The experimental groups included three species pairs exposed to (1) a natural 16 h photoperiod, (2) natural CL, (3) an artificial 16 h photoperiod and (4) artificial CL. In the natural environment, measurements of physiological and biochemical parameters were carried out at the peak of the polar day (at the end of June), when the plants were illuminated continuously, and in the second week of August, when the day length was about 16 h. Th experiments with artificial lighting were conducted in climate chambers where plants were exposed to 16 h or 24 h photoperiods for two weeks. Other parameters (light intensity, spectrum composition, temperature and air humidity) were held constant. The obtained results have shown that plants lack specific mechanisms of tolerance to CL. The protective responses are non-specific and induced by developing photo-oxidative stress. In climate chambers, under constant environmental conditions artificial CL causes leaf injuries due to oxidative stress, the main cause of which is circadian asynchrony. In nature, plants are not photodamaged during the polar day, as endogenous rhythms are maintained due to daily fluctuations of several environmental factors (light intensity, spectral distribution, temperature and air humidity). The obtained data show that among possible non-specific protective mechanisms, plants use flavonoids to neutralize the excess ROS generated under CL. In local subarctic plants, their photoprotective role is significantly higher than in non-native introduced plant species. Full article
(This article belongs to the Section Plant Response to Abiotic Stress and Climate Change)
Show Figures

Figure 1

Figure 1
<p>Location of the study site. Polar-Alpine Botanic Garden (KPABG), Kirovsk, Russia (67°38′ N).</p>
Full article ">Figure 2
<p>Hours of daylight and twilight in Kirovsk (67°38′ N) © WeatherSpark.com. The number of hours during which the sun is visible (black line). From bottom (most yellow) to top (most gray) the color bands indicate full daylight, twilight (civil, nautical and astronomical) and full night.</p>
Full article ">Figure 3
<p>Plants under study. Native-to-subarctic plant species: (<b>a</b>) <span class="html-italic">Geranium sylvaticum</span> L., (<b>b</b>) <span class="html-italic">Geum rivale</span> L., (<b>c</b>) <span class="html-italic">Potentilla erecta</span> (L.) Raeusch.; non-native plant species: (<b>d</b>) <span class="html-italic">Geranium himalayense</span> Klotzsch, (<b>e</b>) <span class="html-italic">Geum coccineum</span> Sibth. and Sm., and (<b>f</b>) <span class="html-italic">Potentilla atrosanguinea</span> Loddiges ex D. Don.</p>
Full article ">Figure 4
<p>The daily course of stomatal conductance of A-16 and A-CL leaves of <span class="html-italic">Potentilla atrosanguinea</span>.</p>
Full article ">Figure 5
<p>Leaves of plants exposed to A-16 (top on (<b>a</b>,<b>c</b>–<b>e</b>) and left on (<b>b</b>,<b>f</b>)) and A-CL (bottom on (<b>a</b>,<b>c</b>–<b>e</b>) and right on (<b>b</b>,<b>f</b>)) for 2 weeks. (<b>a</b>) <span class="html-italic">Geranium sylvaticum</span>, (<b>b</b>) <span class="html-italic">Geum rivale</span>, (<b>c</b>) <span class="html-italic">Potentilla erecta</span>, (<b>d</b>) <span class="html-italic">Geranium himalayense</span>, (<b>e</b>) <span class="html-italic">Geum coccineum</span> and (<b>f</b>) <span class="html-italic">Potentilla atrosanguinea</span>.</p>
Full article ">Figure 6
<p>Leaf color pattern in aging N-16 h leaves. (<b>a</b>) <span class="html-italic">Geranium sylvaticum</span>, (<b>b</b>) <span class="html-italic">Geum rivale</span>, (<b>c</b>) <span class="html-italic">Potentilla erecta</span>, (<b>d</b>) <span class="html-italic">Geranium himalayense</span>, (<b>e</b>) <span class="html-italic">Geum coccineum</span> and (<b>f</b>) <span class="html-italic">Potentilla atrosanguinea</span>.</p>
Full article ">Figure 7
<p>Photosynthesis active radiation (PAR) measured during the polar day on 24–26 June 2024 at Kirovsk, Russia (67°38′ N).</p>
Full article ">
30 pages, 2014 KiB  
Systematic Review
Postnatal Development of the Circadian Rhythmicity of Human Pineal Melatonin Synthesis and Secretion (Systematic Review)
by Ekkehart Paditz
Children 2024, 11(10), 1197; https://doi.org/10.3390/children11101197 - 29 Sep 2024
Viewed by 562
Abstract
Introduction: According to current knowledge, at birth, the pineal gland and melatonin receptors are already present and the suprachiasmatic nucleus is largely functional, and noradrenaline, the key pineal transmitter, can be detected in the early foetal period. It is still unclear why the [...] Read more.
Introduction: According to current knowledge, at birth, the pineal gland and melatonin receptors are already present and the suprachiasmatic nucleus is largely functional, and noradrenaline, the key pineal transmitter, can be detected in the early foetal period. It is still unclear why the pineal gland is not able to start its own pulsatile synthesis and secretion of melatonin in the first months of life, and as a result, infants during this time are dependent on an external supply of melatonin. Method: The causes and consequences of this physiological melatonin deficiency in human infancy are examined in a systematic review of the literature, in which 40 of 115 initially selected publications were evaluated in detail. The references of these studies were checked for relevant studies on this topic. References from previous reviews by the author were taken into account. Results: The development and differentiation of the pineal gland, the pinealocytes, as the site of melatonin synthesis, and the development and synaptic coupling of the associated predominantly noradrenergic neural pathways and vessels and the associated Lhx4 homebox only occurs during the first year of life. Discussion: The resulting physiological melatonin deficiency is associated with sleep disorders, infant colic, and increased crying in babies. Intervention studies indicate that this deficiency should be compensated for through breastfeeding, the administration of nonpooled donor milk, or through industrially produced chrononutrition made from nonpooled cow’s milk with melatonin-poor day milk and melatonin-rich night milk. Full article
(This article belongs to the Special Issue Current Advances in Paediatric Sleep Medicine)
Show Figures

Figure 1

Figure 1
<p>Flow chart with declaration of the selection procedure applied here. List of excluded items, see <a href="#app1-children-11-01197" class="html-app">Supplement S1</a>. Source of the flow chart: Page MJ, et al. BMJ 2021;372:n71. doi: 10.1136/bmj.n71. This work is licensed under CC BY 4.0.</p>
Full article ">Figure 2
<p>Melatonin concentrations in human breast milk with circadian rhythmicity in colostrum (daytime = blue) and in the first six months (daytime = grey) and at night in colostrum (orange) and in the first six months (yellow). Melatonin concentration in pg/mL. The data were taken from the review by Oliveira et al., 2024 and are presented graphically in aggregate form. The individual measured values for colostrum are taken from the studies by Qin (1), Aparici-Gonzalo (2), Pontes 2007 (3), Pontes 2006 (4), Honori-Franca (5), Silva (6), and Illnerova (7), as well as for mature milk from Kimata (1), Aparici-Gonzalo (2), Cohen Engler (3), Molad (4), and Silva (5) [<a href="#B20-children-11-01197" class="html-bibr">20</a>,<a href="#B37-children-11-01197" class="html-bibr">37</a>,<a href="#B38-children-11-01197" class="html-bibr">38</a>,<a href="#B39-children-11-01197" class="html-bibr">39</a>,<a href="#B142-children-11-01197" class="html-bibr">142</a>,<a href="#B143-children-11-01197" class="html-bibr">143</a>,<a href="#B144-children-11-01197" class="html-bibr">144</a>,<a href="#B145-children-11-01197" class="html-bibr">145</a>,<a href="#B147-children-11-01197" class="html-bibr">147</a>,<a href="#B178-children-11-01197" class="html-bibr">178</a>,<a href="#B179-children-11-01197" class="html-bibr">179</a>].</p>
Full article ">Figure 3
<p>Analysis of relative optical density (O.D.) of inducible cAMP early repressor (ICER) hybridization signal in rat pineal gland (solid line) and superimposed pineal melatonin values (dashed line). ICER night-time values start to be significantly different from daytime values from P8 onward (P8: <span class="html-italic">p</span> &lt; 0.05; P10, P15, P20, adult (Ad): <span class="html-italic">p</span> &lt; 0.01). Night-time melatonin values start to be significantly different from daytime values at P8 (P8: <span class="html-italic">p</span> &lt; 0.05; P15: <span class="html-italic">p</span> &lt; 0.01) [<a href="#B189-children-11-01197" class="html-bibr">189</a>]. With kind permission.</p>
Full article ">Figure 4
<p>Representation of the signal transduction chain for the activation of pulsatile pineal synthesis and secretion, with indication of the localisation of M1 and M2 melatonin receptors in the brain that are associated with sleep [<a href="#B204-children-11-01197" class="html-bibr">204</a>,<a href="#B205-children-11-01197" class="html-bibr">205</a>]. Light is converted into chemical impulses in the photosensitive ganglion cells of the retina, which stimulate the retinal formation of melanopsin. Melanopsin inhibits the synthesis of melatonin [<a href="#B206-children-11-01197" class="html-bibr">206</a>]. The central master clock SCN is activated via the retinohypothalamic tract. Via several sympathetic ganglia (paraventricular nucleus (PVN), upper thoracic medulla, cervical ganglion), melatonin synthesis in the pineal gland is inhibited by light or activated by darkness in the evening. The pineal gland secretes melatonin directly into the cerebrospinal fluid and via venous effluents into the jugular vein. Melatonin exerts its effects in particular via two receptor types (M1, M2), which stimulate the switch from wakefulness to sleep in the frontal pre-cortex (M1), after melatonin has induced the transition from wakefulness to NREM sleep via feedback mechanisms to the SCN (M1, M2). The thalamus, as the ‘gateway to consciousness’, is sent into NREM sleep via M2 receptors and is opened in this state for the transfer of verbal information from short-term memory to long-term memory in the hippocampus; the consolidation of memory content takes place to a large extent during undisturbed sleep. The transition from NREM to REM sleep is induced by M2 receptors in the ventrolateral periaqueductal grey matter. During REM sleep, REM muscle atonia is generated via several neural switching points, and motor information can now be stored (‘You learn to ride a bike in your sleep’). The basal forebrain is involved in these processes (M2). A highly simplified overview based on [<a href="#B79-children-11-01197" class="html-bibr">79</a>,<a href="#B204-children-11-01197" class="html-bibr">204</a>,<a href="#B207-children-11-01197" class="html-bibr">207</a>,<a href="#B208-children-11-01197" class="html-bibr">208</a>]. Slightly modified according to Paditz [<a href="#B209-children-11-01197" class="html-bibr">209</a>], with kind permission.</p>
Full article ">Figure 5
<p>Expression of Lhx4 in the developing rat pineal gland. The arrow points to the pineal gland. Scale bar, 1 mm; E, embryonic day; P, postnatal day; ZT, zeitgeber time. Radiochemical in situ hybridisation for detection of Lhx4 mRNA in coronal sections of the brain from rats sacrificed at ZT6 (<b>left</b>) and ZT18 (<b>middle</b>) at the indicated developmental stages (<b>one per row</b>) ranging from E15 to P30. ZT18 sections were counterstained in cresyl violet for comparison (<b>right</b>). Reproduced from Hertz [<a href="#B198-children-11-01197" class="html-bibr">198</a>], with kind permission.</p>
Full article ">
11 pages, 263 KiB  
Article
Circadian Rhythm Genes and Their Association with Sleep and Sleep Restriction
by Marcin Sochal, Marta Ditmer, Aleksandra Tarasiuk-Zawadzka, Agata Binienda, Szymon Turkiewicz, Adam Wysokiński, Filip Franciszek Karuga, Piotr Białasiewicz, Jakub Fichna and Agata Gabryelska
Int. J. Mol. Sci. 2024, 25(19), 10445; https://doi.org/10.3390/ijms251910445 - 27 Sep 2024
Viewed by 369
Abstract
Deprivation of sleep (DS) and its effects on circadian rhythm gene expression are not well understood despite their influence on various physiological and psychological processes. This study aimed to elucidate the changes in the expression of circadian rhythm genes following a night of [...] Read more.
Deprivation of sleep (DS) and its effects on circadian rhythm gene expression are not well understood despite their influence on various physiological and psychological processes. This study aimed to elucidate the changes in the expression of circadian rhythm genes following a night of sleep and DS. Their correlation with sleep architecture and physical activity was also examined. The study included 81 participants who underwent polysomnography (PSG) and DS with actigraphy. Blood samples were collected after PSG and DS. Expression levels of brain and muscle ARNT-like 1 (BMAL1), circadian locomotor output cycles kaput (CLOCK), neuronal PAS domain protein 2 (NPAS2), period 1 (PER1), cryptochrome 1 (CRY1) and nuclear receptor subfamily 1 group D member 1 (NR1D1) were analyzed using qRT-PCR. DS decreased the expression of CLOCK and BMAL1 while increasing PER1. PER1 expression correlated positively with total sleep time and non-rapid-eye-movement (NREM) sleep duration and negatively with sleep latency, alpha, beta and delta waves in the O1A2 lead. Physical activity during DS showed positive correlations with CLOCK, BMAL1, and CRY1. The findings highlight the role of PER1 in modulating sleep patterns, suggesting potential targets for managing sleep-related disorders. Further research is essential to deepen the understanding of these relationships and their implications. Full article
(This article belongs to the Section Molecular Genetics and Genomics)
13 pages, 3056 KiB  
Article
Adaptive Differences in Cellular and Behavioral Responses to Circadian Disruption between C57BL/6 and BALB/c Strains
by Changxiao Ma, Haonan Li, Wenyu Li, Guangrui Yang and Lihong Chen
Int. J. Mol. Sci. 2024, 25(19), 10404; https://doi.org/10.3390/ijms251910404 - 27 Sep 2024
Viewed by 439
Abstract
The regulation of the mammalian circadian clock is largely dependent on heredity. In model animals for circadian rhythm studies, C57BL/6 and BALB/c mice exhibit considerable differences in their adaptation to circadian disruption, yet deeper comparisons remain unexplored. Here, we have established embryonic fibroblast [...] Read more.
The regulation of the mammalian circadian clock is largely dependent on heredity. In model animals for circadian rhythm studies, C57BL/6 and BALB/c mice exhibit considerable differences in their adaptation to circadian disruption, yet deeper comparisons remain unexplored. Here, we have established embryonic fibroblast cells derived from C57BL/6 mice (MEF) and BALB/c (BALB/3T3) mice, which have been transfected with the Bmal1 promoter-driven luciferase (Bmal1-Luc) reporter gene. Next, dexamethasone was applied for various cyclic stimulations, which revealed that Bmal1 bioluminescence of MEF cells was entrained to 24 to 26 h cycles, whereas BALB/3T3 cells have a wider range (22 to 28 h) with lower amplitudes. Behaviorally, BALB/c mice swiftly adapted to a 6-h advance light/dark cycle, unlike C57BL/6 mice. Furthermore, we found the expression of the circadian rhythm gene Npas2 in BALB/c mice is significantly lower than that in C57BL/6 mice. This observation is consistent with the differentially expressed genes (DEGs) in the intestine and lung tissues of C57BL/6 and BALB/c mice, based on the RNA-seq datasets downloaded from the Gene Expression Omnibus (GEO). In summary, our study uncovers that BALB/c mice possess greater resilience in circadian rhythm than C57BL/6 mice, both cellular and behaviorally, identifying potential genes underlying this difference. Full article
(This article belongs to the Section Molecular Microbiology)
Show Figures

Figure 1

Figure 1
<p>Measurement of the adaptability of MEF and BALB/3T3 cellular circadian rhythms to DEX T cycles &lt; 24 h. (<b>A</b>) Free run; (<b>B</b>) T16; (<b>C</b>) T18; (<b>D</b>) T20; (<b>E</b>) T22. (<b>F</b>) The amplitude of <span class="html-italic">Bmal1</span> in MEF and BALB/3T3 cells under different T cycles (n = 3; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 2
<p>Measurement of the adaptability of MEF and BALB/3T3 cellular circadian rhythms to DEX T cycles ≥ 24 h. (<b>A</b>) T24; (<b>B</b>) T26; (<b>C</b>) T28; (<b>D</b>) T30. (<b>E</b>) The amplitude of <span class="html-italic">Bmal1</span> in MEF and BALB/3T3 cells under different T cycles (n = 3; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 3
<p>The basal behavioral rhythm in C57BL/6 and BALB/c mice. (<b>A</b>) Representative double-plotted actograms of wheel running activity of C57BL/6 and BALB/c mice in LD cycle. (<b>B</b>) The average 10-day diurnal activity profile in LD cycles. (<b>C</b>) Daytime, nighttime, and total activity levels. (<b>D</b>) Double-plotted actograms of wheel running activity of C57BL/6 and BALB/c mice in DD condition. (<b>E</b>) Totally activity in DD. (<b>F</b>) FRP. (<b>G</b>) The FFT in LD and DD (n = 14; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 4
<p>The entrainment of C57BL/6 and BALB/c mice to the +6 shift. (<b>A</b>) Representative double-plotted actograms of wheel running activity of mice treated with the + 6 shift. The white bar represents the day, and the gray bar represents the nighttime. (<b>B</b>) Quantified activity onsets. (<b>C</b>) PS50 values are calculated by activity onset. (<b>D</b>) Phase angle of entrainment. (<b>E</b>) FFT relative power for 7 days before and after the shift. (<b>F</b>) The amplitude of circadian behavior revealed by FFT relative power (n = 14; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 5
<p>The internal behavioral rhythm of C57BL/6 and BALB/c mice in the +6 shift. (<b>A</b>) Representative double-plotted actograms of running-wheel activity of mice in the +6 shift followed by DD. (<b>B</b>) Quantification of activity onset. (<b>C</b>) The magnitude of the phase shifts after DD (n = 14; *** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 6
<p>Expression rhythms of core clock genes in MEF and BALB/3T3 cells. (<b>A</b>) <span class="html-italic">Bmal1</span>; (<b>B</b>) <span class="html-italic">Per1</span>; (<b>C</b>) <span class="html-italic">Per2</span>; (<b>D</b>) <span class="html-italic">Clock</span>; (<b>E</b>) <span class="html-italic">Cry1</span>; (<b>F</b>) <span class="html-italic">Cry2</span>; (<b>G</b>) <span class="html-italic">Nr1d1</span>; (<b>H</b>) <span class="html-italic">Npas2</span> (n = 3; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 7
<p>Expression of core clock genes in C57BL/6 and BALB/c mice tissue. (<b>A</b>) mRNA (n = 4) and (<b>B</b>) western blotting (n = 3) and corresponding data statistics for expression of core clock genes in the hypothalamus, intestine, and lung (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">
18 pages, 4374 KiB  
Article
Hepatocellular Carcinoma in Mice Affects Neuronal Activity and Glia Cells in the Suprachiasmatic Nucleus
by Mona Yassine, Soha A. Hassan, Lea Aylin Yücel, Fathima Faiba A. Purath, Horst-Werner Korf, Charlotte von Gall and Amira A. H. Ali
Biomedicines 2024, 12(10), 2202; https://doi.org/10.3390/biomedicines12102202 - 27 Sep 2024
Viewed by 538
Abstract
Background: Chronic liver diseases such as hepatic tumors can affect the brain through the liver–brain axis, leading to neurotransmitter dysregulation and behavioral changes. Cancer patients suffer from fatigue, which can be associated with sleep disturbances. Sleep is regulated via two interlocked mechanisms: [...] Read more.
Background: Chronic liver diseases such as hepatic tumors can affect the brain through the liver–brain axis, leading to neurotransmitter dysregulation and behavioral changes. Cancer patients suffer from fatigue, which can be associated with sleep disturbances. Sleep is regulated via two interlocked mechanisms: homeostatic regulation and the circadian system. In mammals, the hypothalamic suprachiasmatic nucleus (SCN) is the key component of the circadian system. It generates circadian rhythms in physiology and behavior and controls their entrainment to the surrounding light/dark cycle. Neuron–glia interactions are crucial for the functional integrity of the SCN. Under pathological conditions, oxidative stress can compromise these interactions and thus circadian timekeeping and entrainment. To date, little is known about the impact of peripheral pathologies such as hepatocellular carcinoma (HCC) on SCN. Materials and Methods: In this study, HCC was induced in adult male mice. The key neuropeptides (vasoactive intestinal peptide: VIP, arginine vasopressin: AVP), an essential component of the molecular clockwork (Bmal1), markers for activity of neurons (c-Fos), astrocytes (GFAP), microglia (IBA1), as well as oxidative stress (8-OHdG) in the SCN were analyzed by immunohistochemistry at four different time points in HCC-bearing compared to control mice. Results: The immunoreactions for VIP, Bmal1, GFAP, IBA1, and 8-OHdG were increased in HCC mice compared to control mice, especially during the activity phase. In contrast, c-Fos was decreased in HCC mice, especially during the late inactive phase. Conclusions: Our data suggest that HCC affects the circadian system at the level of SCN. This involves an alteration of neuropeptides, neuronal activity, Bmal1, activation of glia cells, and oxidative stress in the SCN. Full article
(This article belongs to the Special Issue Understanding Diseases Affecting the Central Nervous System)
Show Figures

Figure 1

Figure 1
<p>Vasoactive intestinal peptide (VIP) immunoreaction (IR) in the SCN. Representative fluorescent microphotographs showing the immunoreaction (IR) of VIP (red) in the suprachiasmatic nucleus (SCN) of (<b>A</b>–<b>D</b>) PB control mice and (<b>E</b>–<b>H</b>) HCC mice. The white asterisk indicates the ventral core region, while white arrowhead indicates the dorsal shell region of SCN. 3v: third ventricle. Scale bar = 100 μm. Mice were sacrificed at different time points at 6 h intervals starting at Zeitgeber time (ZT) 02 = 2 h after the light on. (<b>I</b>) Quantification of VIP-immunoreaction (IR) in arbitrary unit (A.U.) in the SCN. White and black bar indicates light/dark phase, respectively. Two-way ANOVA followed by Sidak’s multiple comparisons test. Total of 12 mice per group, n = 3 mice at each time point. *: <span class="html-italic">p</span> &lt; 0.05 between PB control and HCC.</p>
Full article ">Figure 2
<p>Arginine vasopressin (AVP) immunoreaction (IR) in the SCN. Representative fluorescent microphotographs showing the immunoreaction (IR) of AVP (green) in the SCN of (<b>A</b>–<b>D</b>) PB control mice and (<b>E</b>–<b>H</b>) HCC mice. 3v: third ventricle. Scale bar = 100 μm. Mice were sacrificed at different time points at 6 h intervals starting at ZT02 = 2 h after the light was on. (<b>I</b>) Quantification of AVP-immunoreaction (IR) in arbitrary unit (A.U.) in the SCN. White and black bar indicates light/dark phase, respectively. Two-way ANOVA followed by Sidak’s multiple comparisons test. Total of 12 mice per group, n = 3 mice at each time point. *: <span class="html-italic">p</span> &lt; 0.05 between PB control and HCC.</p>
Full article ">Figure 3
<p>Orexin-immunoreactive (ir) cells in the lateral hypothalamus. Representative fluorescent microphotographs showing orexin-ir cells (green) in the lateral hypothalamus (LH) of (<b>A</b>–<b>D</b>) PB control mice and (<b>E</b>–<b>H</b>) HCC-bearing mice. Scale bar = 200 μm. Mice were sacrificed at different time points at 6 h intervals starting at ZT02 = 2 h after the light was on. (<b>I</b>) Quantification of number of orexin-immunoreactive cells per mm<sup>2</sup> in the lateral hypothalamus. White and black bar indicates light/dark phase, respectively. Two-way ANOVA followed by Sidak’s multiple comparisons test. Total of 12 mice per group, n = 3 mice at each time point.</p>
Full article ">Figure 4
<p>Bmal1-immunoreaction (IR) in the SCN. Representative bright-field photomicrographs showing the Bmal1-immunoreactive cells (brown staining) in the SCN of (<b>A</b>–<b>D</b>) PB control mice and (<b>E</b>–<b>H</b>) in HCC-bearing mice. 3v: third ventricle. OC: optic chiasma. Scale bar = 100 μm. Mice were sacrificed at different time points at 6 h intervals starting at ZT02 = 2 h after the light on. (<b>I</b>) Quantification of Bmal1-IR in arbitrary units (A.U.) in the SCN at individual time points at 6 h intervals. White and black bar indicates light/dark phase, respectively. Two-way ANOVA followed by Sidak’s multiple comparisons test. Total of 12 mice per group, n = 3 mice at each time point. **: <span class="html-italic">p</span> &lt; 0.01 between PB control and HCC.</p>
Full article ">Figure 5
<p>c-Fos-immunoreactive (ir) cells in the SCN. Representative bright-field photomicrograph showing the positively stained c-Fos cells (brown staining) in the SCN of (<b>A</b>–<b>D</b>) PB control mice and (<b>E</b>–<b>H</b>) in HCC-bearing mice. 3v: third ventricle. OC: optic chiasma. Scale bar = 100 μm. Mice were sacrificed at different time points at 6 h intervals starting at ZT02 = 2 h after the light was on. (<b>I</b>) Quantification of c-Fos-ir cells/mm<sup>2</sup> in the SCN. White and black bar indicates for light/dark phase, respectively. Two-way ANOVA followed by Sidak’s multiple comparisons test. Total of 12 mice per group, n = 3 mice at each time point. *: <span class="html-italic">p</span> &lt; 0.05 between PB control and HCC.</p>
Full article ">Figure 6
<p>Astrocytic marker GFAP-immunoreaction (IR) in SCN. Representative fluorescent microphotographs showing GFAP immunoreaction (IR) (green) in the SCN of (<b>A</b>–<b>D</b>) PB control mice and (<b>E</b>–<b>H</b>) HCC mice. 3v: third ventricle. Scale bar = 150 μm. Mice were sacrificed at different time points at 6 h intervals starting at ZT02 = 2 h after the light was on. (<b>I</b>) Quantification of GFAP-immunoreaction (IR) in arbitrary unit (A.U.) in the SCN. White and black bar indicates light/dark phase, respectively. Two-way ANOVA followed by Sidak’s multiple comparisons test. Total of 12 mice per group n = 3 mice at each time point. **: <span class="html-italic">p</span> &lt; 0.01 between PB control and HCC.</p>
Full article ">Figure 7
<p>Microglial marker IBA-1-immunoreaction (IR) in SCN. Representative fluorescent photomicrographs showing IBA-1-immunoreaction (IR) (red) in the SCN of (<b>A</b>–<b>D</b>) PB control mice and of (<b>E</b>–<b>H</b>) HCC-bearing mice. 3v: third ventricle. Scale bar = 150 μm. Mice were sacrificed at different time points at 6 h intervals starting at ZT02 = 2 h after the light was on. (<b>I</b>) Quantification of IBA-1-immunoreaction (IR) in arbitrary unit (A.U.) in the SCN. White and black bar indicates for light/dark phase, respectively. Two-way ANOVA followed by Sidak’s multiple comparisons test. Total of 12 mice per group, n = 3 mice at each time point. **: <span class="html-italic">p</span> &lt; 0.01 between control and HCC.</p>
Full article ">Figure 8
<p>Oxidative stress marker 8-OHdG-immunoreaction in (IR) SCN. Representative fluorescent photomicrographs showing the immunoreaction (IR) of 8- hydroxydeoxyguanosine (8-OHdG) (red) in the SCN of (<b>A</b>–<b>D</b>) PB control mice and of (<b>E</b>–<b>H</b>) HCC-bearing mice. 3v: third ventricle. OC: optic chiasma. Scale bar = 100 μm. Mice were sacrificed at different time points at 6 h intervals starting at ZT02 = 2 h after the light was on. (<b>I</b>) Quantification of 8-OHdG-immunoreaction (IR) in arbitrary unit (A.U.) in the SCN. White and black bar indicates light/dark phase, respectively. Two-way ANOVA followed by Sidak’s multiple comparisons test. Total of 12 mice per group, n = 3 mice at each time point. *: <span class="html-italic">p</span> &lt; 0.05 between PB control and HCC.</p>
Full article ">
13 pages, 1718 KiB  
Article
Melatonin Supplementation Alleviates Impaired Spatial Memory by Influencing Aβ1-42 Metabolism via γ-Secretase in the icvAβ1-42 Rat Model with Pinealectomy
by Irina Georgieva, Jana Tchekalarova, Zlatina Nenchovska, Lidia Kortenska and Rumiana Tzoneva
Int. J. Mol. Sci. 2024, 25(19), 10294; https://doi.org/10.3390/ijms251910294 - 24 Sep 2024
Viewed by 681
Abstract
In the search for Alzheimer’s disease (AD) therapies, most animal models focus on familial AD, which accounts for a small fraction of cases. The majority of AD cases arise from stress factors, such as oxidative stress, leading to neurological changes (sporadic AD). Early [...] Read more.
In the search for Alzheimer’s disease (AD) therapies, most animal models focus on familial AD, which accounts for a small fraction of cases. The majority of AD cases arise from stress factors, such as oxidative stress, leading to neurological changes (sporadic AD). Early in AD progression, dysfunction in γ-secretase causes the formation of insoluble Aβ1-42 peptides, which aggregate into senile plaques, triggering neurodegeneration, cognitive decline, and circadian rhythm disturbances. To better model sporadic AD, we used a new AD rat model induced by intracerebroventricular administration of Aβ1-42 oligomers (icvAβ1-42) combined with melatonin deficiency via pinealectomy (pin). We validated this model by assessing spatial memory using the radial arm maze test and measuring Aβ1-42 and γ-secretase levels in the frontal cortex and hippocampus with ELISA. The icvAβ1-42 + pin model experienced impaired spatial memory and increased Aβ1-42 and γ-secretase levels in the frontal cortex and hippocampus, effects not seen with either icvAβ1-42 or the pin alone. Chronic melatonin treatment reversed memory deficits and reduced Aβ1-42 and γ-secretase levels in both structures. Our findings suggest that our icvAβ1-42 + pin model is extremely valuable for future AD research. Full article
(This article belongs to the Special Issue Molecular and Cellular Mechanisms of Apoptosis and Senescence)
Show Figures

Figure 1

Figure 1
<p>The effect of pinealectomy (pin) and chronic treatment with melatonin (mel) on icvAβ<sub>1-42</sub> -related effect on (<b>A</b>) mean working memory errors (WMEs) during each trial 1st to 5th, (<b>B</b>) the average WMEs, (<b>C</b>) the average double WMEs (DWMEs), and (<b>D</b>) the average time to fulfill the criterion. Data are presented as mean ± S.E.M. (<b>A</b>) *** <span class="html-italic">p</span> &lt; 0.001, C-pin-veh vs. C-sham-veh group; * <span class="html-italic">p</span> = 0.026, Aβ-sham-veh vs. C-sham-veh; ** <span class="html-italic">p</span> &lt; 0.01, Aβ-pin-veh vs. C-sham-veh; *** <span class="html-italic">p</span> &lt; 0.001, C-pin-mel vs. C-pin-veh; ** <span class="html-italic">p</span> = 0.005, Aβ-pin-mel vs. Aβ-pin-veh; (<b>B</b>) * <span class="html-italic">p</span> = 0.018, C-pin-veh vs. C-sham-veh; * <span class="html-italic">p</span> = 0.048, Aβ-sham-veh vs. C-sham; <span class="html-italic">p</span> = 0.05, Aβ-pin-veh vs. C-sham; <span class="html-italic">p</span> = 0.031, Aβ-pin-mel vs. Aβ-pin-veh; (<b>C</b>) *** <span class="html-italic">p</span> &lt; 0.001, Aβ-sham-veh and Aβ-pin-veh vs. C-sham-veh group; *** <span class="html-italic">p</span> &lt; 0.001, Aβ-sham-mel vs. Aβ-sham-veh; * <span class="html-italic">p</span> = 0.016, Aβ-pin-mel vs. Aβ-pin-veh; (<b>D</b>) * <span class="html-italic">p</span> = 0.05, C-pin-veh vs. C-sham-veh; ** <span class="html-italic">p</span> = 0.007, Aβ-sham-veh vs. C-sham-veh; ** <span class="html-italic">p</span> = 0.01, Aβ-pin-veh vs. C-sham-veh; * <span class="html-italic">p</span> &lt; 0.05, C-pin-mel vs. C-pin-veh; ** <span class="html-italic">p</span> = 0.01, Aβ-sham-mel vs. Aβ-sham-veh.</p>
Full article ">Figure 2
<p>Aβ<sub>1-42</sub> levels (pg/mL) under the influence of icvAβ<sub>1-42</sub> (Aβ) and/or pinealectomy (pin) and chronic treatment with melatonin (mel) in the (<b>A</b>) frontal cortex and (<b>B</b>) hippocampus. Data are presented as mean ± S.E.M. (<b>A</b>) *** <span class="html-italic">p</span> &lt; 0.001, C-sham-veh vs. Aβ-pin-veh group and Aβ-pin-veh vs. Aβ-pin-mel group. (<b>B</b>) *** <span class="html-italic">p</span> &lt; 0.001, C-sham-veh vs. Aβ-pin-veh group; Aβ-pin-veh vs. Aβ-pin-mel group.</p>
Full article ">Figure 3
<p>The influence of icvAβ<sub>1-42</sub> (Aβ) and/or pinealectomy (pin) and chronic treatment with melatonin (mel) on γ-secretase (GS) levels (ng/L) in the (<b>A</b>) frontal cortex and (<b>B</b>) hippocampus. Data are presented as mean ± S.E.M. (<b>A</b>) *** <span class="html-italic">p</span> &lt; 0.001, C-sham-veh compared to Aβ-pin-veh group and Aβ-pin-veh compared to Aβ-pin-mel group. (<b>B</b>) * <span class="html-italic">p</span> &lt; 0.05, C-sham-veh compared to C-pin-veh; C-sham-veh compared to Aβ-sham-veh; C-sham-veh compared to Aβ-pin-veh group; *** <span class="html-italic">p</span> &lt; 0.001, C-pin-veh compared to C-pin-mel group.</p>
Full article ">Figure 4
<p>Schematic diagram of the experimental protocol. Brain structures used to determine biochemical markers were taken from the cohort of rats studied in the radial arm maze test.</p>
Full article ">
14 pages, 425 KiB  
Article
Prospective Evaluation of Transsphenoidal Pituitary Surgery in Patients with Cushing’s Disease: Delayed Remission and the Role of Postsurgical Cortisol as a Predictive Factor
by Athanasios Saratziotis, Maria Baldovin, Claudia Zanotti, Sara Munari, Diego Cazzador, Enrico Alexandre, Luca Denaro, Jiannis Hajiioannou and Enzo Emanuelli
Healthcare 2024, 12(18), 1900; https://doi.org/10.3390/healthcare12181900 - 22 Sep 2024
Viewed by 588
Abstract
Background. Transsphenoidal surgery is the treatment of choice for Cushing’s disease. Successful surgery is associated with subnormal postoperative serum cortisol concentrations and cortisoluria levels, which may guide decisions regarding immediate reoperation. Remission is defined as the biochemical reversal of hypercortisolism with the re-emergence [...] Read more.
Background. Transsphenoidal surgery is the treatment of choice for Cushing’s disease. Successful surgery is associated with subnormal postoperative serum cortisol concentrations and cortisoluria levels, which may guide decisions regarding immediate reoperation. Remission is defined as the biochemical reversal of hypercortisolism with the re-emergence of diurnal circadian rhythm. Methods. A single-center prospective cohort study was conducted among thirty-three patients who underwent transsphenoidal pituitary surgery for Cushing’s disease. Postoperative surgical outcomes, daily morning cortisolemia, and 24 h urinary-free cortisol from the first to the fifth morning were evaluated. Results. All patients underwent surgery, with a remission rate of 81.2%. Of the 26 patients who achieved early remission, 92% remained in remission. Two patients (7.7%) showed recurrence of Cushing’s disease during a mean follow-up of 81.7 months. Early postoperative hypocortisolism suggests complete removal of the tumor, correlating with high rates of remission (p < 0.001). Also, in 12.5% of patients with early cortisol values >138 nmol/L, there was a gradual late remission. Conclusions. In our cohort of patients, the endoscopic transsphenoidal approach was safe and effective in the treatment of Cushing’s disease. We demonstrated that serum and urinary cortisol concentrations did not experience significant fluctuations from the first to the fifth day. This constitutes an accurate predictor of durable remission, comprising a distinctive finding in the intermediate term by our team. Full article
Show Figures

Figure 1

Figure 1
<p>Evaluation of all the examined parameters (blood, urine) from the first to the fifth day from the same patients. (<b>a</b>) Postsurgery cortisolemia in nmol/L on the y-axis, and from the first to the fifth day on the x-axis. (<b>b</b>) Postsurgery cortisoluria (nmol/24 h) on the y-axis, and from the first to the fifth day on the x-axis. Pod, Postoperative day.</p>
Full article ">
15 pages, 7401 KiB  
Article
Longitudinal Study of Changes in Ammonia, Carbon Dioxide, Humidity and Temperature in Individually Ventilated Cages Housing Female and Male C57BL/6N Mice during Consecutive Cycles of Weekly and Bi-Weekly Cage Changes
by Martina Andersson, Karin Pernold, Niklas Lilja, Rafael Frias-Beneyto and Brun Ulfhake
Animals 2024, 14(18), 2735; https://doi.org/10.3390/ani14182735 - 21 Sep 2024
Viewed by 698
Abstract
Housing conditions are essential for ensuring animal welfare and high-quality research outcomes. In this study, we continuously monitored air quality—specifically ammonia, carbon dioxide, relative humidity, and temperature—in Individually Ventilated Cages (IVCs) housing five female or male C57BL/6N mice. The cages were cleaned either [...] Read more.
Housing conditions are essential for ensuring animal welfare and high-quality research outcomes. In this study, we continuously monitored air quality—specifically ammonia, carbon dioxide, relative humidity, and temperature—in Individually Ventilated Cages (IVCs) housing five female or male C57BL/6N mice. The cages were cleaned either weekly or bi-weekly, and the data were collected as the mice aged from 100 to 348 days. The survival rate remained above 96%, with body weight increasing by 35–52% during the study period. The ammonia levels rose throughout the cleaning cycle, but averaged below 25 ppm. However, in the older, heavier mice with bi-weekly cage cleaning, the ammonia levels reached between 25 and 75 ppm, particularly in the males. While circadian rhythms influenced the ammonia concentration only to a small extent, the carbon dioxide levels varied between 800 and 3000 ppm, increasing by 30–50% at night and by 1000 ppm with body weight. Humidity also correlated primarily with the circadian rhythms (10% higher at night) and, to a lesser extent, with body weight, reaching ≥70% in the middle-aged mice. The temperature variations remained minimal, within a 1 °C range. We conclude that air quality assessments in IVCs should be conducted during animals’ active periods, and both housing density and biomass must be considered to optimise welfare. Full article
(This article belongs to the Special Issue Care and Well-Being of Laboratory Animals)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) This micrograph shows the gas sampling probe with several holes at the distal end. The probe passed through the hole on the rear of the lid, where it was secured by a rubber seal. Individual cables connected the probes of the cages to the sensors located in the adjacent room. A pump sucked air from the cages at a constant rate of 1000 mL h<sup>−1</sup> per cage. (<b>B</b>) The mice tolerated the probe well and often nested near it. (<b>C</b>) The airflow dynamics of the IVC GM500 cage, with blue arrows indicating incoming air and red arrows indicating outgoing air. The manufacturer’s recommended ventilation rate is 75 ACH. (<b>D</b>) This image shows the IVC, as in C, with the most common locations of the nest(s) (blue area) or latrine(s) (red areas) and the position of the sampling probe (orange profile).</p>
Full article ">Figure 2
<p>(<b>A</b>) This shows the survival of the mice used in this study for each sex and housing condition. The colour code at the top of panel B indicates the housing conditions. The recording periods are marked with black lines. (<b>B</b>) This shows the average and SEM of body weight (in grams) for the female and male mice housed with weekly (7 d) and bi-weekly (14 d) cage changes, respectively. The colour code at the top of panel A indicates sex and the different housing conditionsThe periods of the first and second recordings are marked with black lines.</p>
Full article ">Figure 3
<p>Four panels showing the location of the latrines (toilets) and nests in the cages with the female and male mice and weekly or bi-weekly CCs as indicated. The percentage of latrine (red) and nest (blue) locations noted is shown for each of the nine areas into which the floor was divided. The row totals are shown on the left and right, respectively. The location of the sampling probe is indicated in each panel.</p>
Full article ">Figure 4
<p>(<b>A</b>–<b>D</b>) Plots of average (blue line) and 95% CI (light grey error bar) of NH3 ppm every hour (left side of each panel) or every fourth hour (right side of each panel) during consecutive cycles of weekly (<b>A</b>,<b>C</b>) and bi-weekly (<b>B</b>,<b>D</b>) CCs. A-B panels show female mice data, while (<b>C</b>,<b>D</b>) are panels showing male mice data. Period 1 (age 100–126 days) and period 2 (age 322–348 days) are marked (on the abscissa) to reflect animal age and experimental phase. Date and time format on the abscissa is month:day:h:min.</p>
Full article ">Figure 5
<p>(<b>A</b>–<b>D</b>) The plots of the average (blue line) and 95% CI (light blue error bar) of CO<sub>2</sub> ppm every hour (left side of each panel) or every fourth hour (right side of each panel) during the first and last two weeks of recordings to resolve the circadian pattern of variations in the cages with female (<b>A</b>,<b>B</b>) and male (<b>C</b>,<b>D</b>) mice having weekly (panels to the left) and bi-weekly (panels to the right) CCs. Period 1 (100–114 days) and period 2 (334–348 days) are indicated on the abscissa. Date and time format on the abscissa is month:day:h:min. The ordinates show ppm CO<sub>2</sub>. The background levels of CO<sub>2</sub> ppm are indicated by orange lines in each panel and were typically 350–400 ppm.</p>
Full article ">Figure 6
<p>(<b>A</b>–<b>D</b>) The plots of the average (blue line) and 95% CI (light blue error bar) of relative humidity (%; ordinates) at every hour (left side of each panel) or every fourth hour (right side of each panel) during the first and last two weeks of the recordings to resolve the circadian pattern of variations in cages humidity with female (<b>A</b>,<b>B</b>) and male (<b>C</b>,<b>D</b>) mice having weekly (panels to the left) and bi-weekly (panels to the right) CCs. Period 1 (100–114 days) and period 2 (334–348 days) are indicated on the abscissa. Date and time format on the abscissa is month:day:h:min.The background humidity is indicated by an orange line in each panel. The average difference between the humidity in the cage and the background humidity is indicated by a red line (lower trace in each panel) together with the 95% CI. The high-frequency fluctuations in the humidity values (background values and measurements inside the cage) in the latter part of period 1 were due to a technical error.</p>
Full article ">Figure 7
<p>(<b>A</b>–<b>D</b>) The plots of the average (blue line) and 95% CI (light blue error bar) of temperature (°C; ordinates) at every hour (left side of each panel) or every fourth hour (right side of each panel) during the first and last two weeks of the recordings to resolve the circadian pattern of variations in cages temperature in cages with female (<b>A</b>,<b>B</b>) and male (<b>C</b>,<b>D</b>) mice having weekly (panels to the left) and bi-weekly (panels to the right) CCs. Period 1 (100–114 days) and period 2 (334–348 days) are indicated on the abscissa. Format of date and time on the abscissa is month:day:h:min. The background temperature is shown as an orange line in each panel. The average difference between the temperature in the cage and the background temperature is indicated by a red line together with the 95% CI. The left ordinates show humidity in %, and the right ordinates show the temperature difference between the cage and the background.</p>
Full article ">Figure 8
<p>This diagram shows the effects of interrupted cage ventilation (red arrow) on carbon dioxide ppm (left ordinate), humidity, and ammonia ppm in the cages (on the right ordinate). When ventilation was resumed (blue arrow), the sensor readings returned to normal values. Format of date and time on the abscissa is month:day:h:min.</p>
Full article ">
16 pages, 9827 KiB  
Article
The Transcriptome Characterization of the Hypothalamus and the Identification of Key Genes during Sexual Maturation in Goats
by Qing Li, Tianle Chao, Yanyan Wang, Rong Xuan, Yanfei Guo, Peipei He, Lu Zhang and Jianmin Wang
Int. J. Mol. Sci. 2024, 25(18), 10055; https://doi.org/10.3390/ijms251810055 - 19 Sep 2024
Viewed by 458
Abstract
Sexual maturation in goats is a dynamic process regulated precisely by the hypothalamic–pituitary–gonadal axis and is essential for reproduction. The hypothalamus plays a crucial role in this process and is the control center of the reproductive activity. It is significant to study the [...] Read more.
Sexual maturation in goats is a dynamic process regulated precisely by the hypothalamic–pituitary–gonadal axis and is essential for reproduction. The hypothalamus plays a crucial role in this process and is the control center of the reproductive activity. It is significant to study the molecular mechanisms in the hypothalamus regulating sexual maturation in goats. We analyzed the serum hormone profiles and hypothalamic mRNA expression profiles of female goats during sexual development (1 day old (neonatal, D1, n = 5), 2 months old (prepuberty, M2, n = 5), 4 months old (sexual maturity, M4, n = 5), and 6 months old (breeding period, M6, n = 5)). The results indicated that from D1 to M6, serum hormone levels, including FSH, LH, progesterone, estradiol, IGF1, and leptin, exhibited an initial increase followed by a decline, peaking at M4. Furthermore, we identified a total of 508 differentially expressed genes in the hypothalamus, with a total of four distinct expression patterns. Nuclear receptor subfamily 1, group D, member 1 (NR1D1), glucagon-like peptide 1 receptor (GLP1R), and gonadotropin-releasing hormone 1 (GnRH-1) may contribute to hormone secretion, energy metabolism, and signal transduction during goat sexual maturation via circadian rhythm regulation, ECM receptor interactions, neuroactive ligand–receptor interactions, and Wnt signaling pathways. This investigation offers novel insights into the molecular mechanisms governing the hypothalamic regulation of goat sexual maturation. Full article
Show Figures

Figure 1

Figure 1
<p>Changes in serum hormone levels in Jining grey goats during sexual maturation. (<b>A</b>). Follicle-stimulating hormone, FSH. (<b>B</b>). Luteinizing hormones, LH. (<b>C</b>). Progesterone, P. (<b>D</b>). Estradiol, E2. (<b>E</b>). Insulin-like growth factor 1, IGF-1. (<b>F</b>). Leptin, LEP. Different lowercase letters indicate significant differences in phenotypic indicators between different days of age (<span class="html-italic">p</span> &lt; 0.05). All data are presented as mean ± standard error. D1, M2, M4, and M6 represent 1-day-old, 2-month-old, 4-month-old, and 6-month-old goats, respectively.</p>
Full article ">Figure 2
<p>Overview of transcriptome analysis based on FPKM values at four stages of goat hypothalamus development. (<b>A</b>). PCA of 20 hypothalamic transcriptome data at four developmental stages of goats. (<b>B</b>). Histogram of the number of DEGs. (<b>C</b>). UpSet plots of the DEG. (<b>D</b>). The clustering heatmap of the DEGs in the 20 samples.</p>
Full article ">Figure 3
<p>Go enrichment analysis of DEGs in different comparison groups during sexual maturation in Jining grey goats. The results of GO enrichment analysis for M2 vs. D1 (<b>A</b>), M4 vs. D1 (<b>B</b>), M4 vs. M2 (<b>C</b>), M6 vs. D1 (<b>D</b>), M6 vs. M4 (<b>E</b>), and M6 vs. M2 (<b>F</b>) are presented respectively. BP, biological process; CC, cellular component; MF, molecular function.</p>
Full article ">Figure 4
<p>KEGG enrichment analysis of DEGs in different comparison groups during sexual maturation in Jining grey goats. (<b>A</b>) M2 vs. D1, (<b>B</b>) M4 vs. M2, (<b>C</b>) M6 vs. M4, (<b>D</b>) M4 vs. D1, and (<b>E</b>,<b>F</b>) M6 vs. D1.</p>
Full article ">Figure 5
<p>Expression pattern cluster analysis based on DEGs. (<b>A</b>): (<b>a</b>–<b>d</b>) Four different expression patterns of DEGs identified using Mfuzz. (<b>B</b>): (<b>a</b>–<b>d</b>) Results of GO analysis in the 4 clusters. (<b>C</b>): (<b>a</b>–<b>d</b>) Results of KEGG enrichment of the top 20 DEGs in the 4 clusters.</p>
Full article ">Figure 6
<p>PPI network modules within the hypothalamus. Four significant modules were constructed using the PPI network of DEGs. (<b>A</b>): Module 1 (MCODE score = 17.89) (<b>B</b>): Module 2 (MCODE score = 9.56) (<b>C</b>): Module 2 (MCODE score = 4) (<b>D</b>): Module 4 (MCODE score = 4). Different colors represent DEGs with different modes of expression.</p>
Full article ">Figure 7
<p>Quantitative validation of goat hypothalamus transcriptome data. (<b>A</b>–<b>M</b>): The <span class="html-italic">X</span>-axis represents different comparison groups, and the <span class="html-italic">Y</span>-axis represents log2(FoldChange) of qRT-PCR and RNA-seq. (<b>N</b>): The Correlation between qRT-PCR and RNA-seq (log2(FoldChange)).</p>
Full article ">
15 pages, 938 KiB  
Review
A New Frontier in Cystic Fibrosis Pathophysiology: How and When Clock Genes Can Affect the Inflammatory/Immune Response in a Genetic Disease Model
by Annalucia Carbone, Pamela Vitullo, Sante Di Gioia, Stefano Castellani and Massimo Conese
Curr. Issues Mol. Biol. 2024, 46(9), 10396-10410; https://doi.org/10.3390/cimb46090618 - 18 Sep 2024
Viewed by 1218
Abstract
Cystic fibrosis (CF) is a monogenic syndrome caused by variants in the CF Transmembrane Conductance Regulator (CFTR) gene, affecting various organ and systems, in particular the lung, pancreas, sweat glands, liver, gastrointestinal tract, vas deferens, and vascular system. While for some [...] Read more.
Cystic fibrosis (CF) is a monogenic syndrome caused by variants in the CF Transmembrane Conductance Regulator (CFTR) gene, affecting various organ and systems, in particular the lung, pancreas, sweat glands, liver, gastrointestinal tract, vas deferens, and vascular system. While for some organs, e.g., the pancreas, a strict genotype-phenotype occurs, others, such as the lung, display a different pathophysiologic outcome in the presence of the same mutational asset, arguing for genetic and environmental modifiers influencing severity and clinical trajectory. CFTR variants trigger a pathophysiological cascade of events responsible for chronic inflammatory responses, many aspects of which, especially related to immunity, are not ascertained yet. Although clock genes expression and function are known modulators of the innate and adaptive immunity, their involvement in CF has been only observed in relation to sleep abnormalities. The aim of this review is to present current evidence on the clock genes role in immune-inflammatory responses at the lung level. While information on this topic is known in other chronic airway diseases (chronic obstructive pulmonary disease and asthma), CF lung disease (CFLD) is lacking in this knowledge. We will present the bidirectional effect between clock genes and inflammatory factors that could possibly be implicated in the CFLD. It must be stressed that besides sleep disturbance and its mechanisms, there are not studies directly addressing the exact nature of clock genes’ involvement in inflammation and immunity in CF, pointing out the directions of new and deepened studies in this monogenic affection. Importantly, clock genes have been found to be druggable by means of genetic tools or pharmacological agents, and this could have therapeutic implications in CFLD. Full article
(This article belongs to the Special Issue Complex Molecular Mechanism of Monogenic Diseases: 3rd Edition)
Show Figures

Figure 1

Figure 1
<p>CFTR mutation classes, mechanisms, and principal examples. CFTR variants are reported as HGVS nomenclature (legacy). The left part has been adapted from Wikimedia Commons [<a href="#B16-cimb-46-00618" class="html-bibr">16</a>]. While class I mutations are associated with the annihilation of CFTR-mediated chloride transport, chloride transport gradually increases through the remaining five classes with the greatest activity being observed in class IV–VI mutations [<a href="#B17-cimb-46-00618" class="html-bibr">17</a>]. The examples are reported as HGVS (Human Genome Variation Society) and legacy nomenclatures. The dashed arrow indicates reduced CFTR synthesis in class V mutations. Red crosses identify mutation classes where the CFTR synthesis, processing or gating is abolished or reduced. The red triangle depicts the decrease in total CFTR mediated chloride transport and is referred to the right part of the Figure (i.e., class mutations).</p>
Full article ">Figure 2
<p>Alteration of circadian clock genes and derived phenotypes in CF mice and cells. CF (<span class="html-italic">F508del</span>/<span class="html-italic">F508del</span>) mice show dysregulation of clock genes expression in lungs and other organs (brain, colon, fat, jejunum, and skeletal muscle tissues). Due to a still-unknown mechanism (denoted by blue dashed arrows), reductions in locomotor activities, phase shift, and decreased serum melatonin levels occur. On the other hand, CF cells in cultures are denoted by microtubule destabilization with further consequences, such as a high perinuclear accumulation of cholesterol and increased pro-inflammatory signaling. Hdac6 KO or inhibition relieves all these phenotypes in vivo and in vitro, suggesting that microtubule dynamics are a modulator of circadian rhythm regulation.</p>
Full article ">
Back to TopTop